08.11.2014 Views

Modern Polymer Spect..

Modern Polymer Spect..

Modern Polymer Spect..

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

<strong>Modern</strong> <strong>Polymer</strong> <strong>Spect</strong>roscopy<br />

Edited by Giuseppe Zerbi<br />

@ WILEY-VCH


This Page Intentionally Left Blank


<strong>Modern</strong><br />

<strong>Polymer</strong> <strong>Spect</strong>roscopy<br />

Edited by<br />

Giuseppe Zerbi<br />

CB WILEY-VCH<br />

Weinheim * New York - Chichester<br />

Brisbane * Singapore - Toronto


Prof. G. Zerbi<br />

Dipartimento di Ingegneria Chimica e Chimica Industriale<br />

Politecnico di Milano<br />

Piazza Leoiiardo da Vinci 32<br />

20132 Milano<br />

Italy<br />

This book was carefully produced. Nevertheless, authors, editor and publisher do not warrant<br />

the information contained therein to be free of errors. Readers are advised to keep in mind<br />

that statements, data, illustrations, procedural details or other items may inadvertently be<br />

inaccurate.<br />

Cover illustration: Dr. Erica Mannucci<br />

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

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

Deutsche Bibliothek Cataloguing-in-Publication Data<br />

<strong>Modern</strong> polymer spectroscopy / ed. by Giuseppe Zerbi. - 1. AuA. - Weinheim;<br />

New York; Chichester; Brisbane; Singapore; Toronto: Wiley-VCH, 1999<br />

ISBN 3-527-29655-7<br />

8 WILEY-VCH Verlag GmbH. D-69469 Weinheim (Federal Republic of Germany), 1999<br />

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

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

be reproduced in any fomi - by photoprinting, microfilm, or any other means - nor transmitted<br />

or translated into machine language without written permission from the publishers. Registered<br />

names, trademarks, etc. used in this book, even when not specifically marked as such. are not to be<br />

considered unprotected by law.<br />

Composition: Asco Typesetters, Hong Kong.<br />

Printing: Strauss Offsetdruck Gin bH., D-69509 Morlenbach.<br />

Bookbinding: Wilhelin Osswald 8 Co. D-67433 Neustaclt.<br />

Printed in the Federal Republic of Germany.


Preface<br />

For unfortunate reasons the success of vibrational [infrared and Raman spectroscopy)<br />

in industrial and university laboratories seems to fade quickly in favor of<br />

other physical techniques which aim at chemical or structural diagnosis of unknown<br />

samples. On the other hand, engineers and instrument manufacturers in the field<br />

of vibrational spectroscopy keep producing magnificent and very sophisticated<br />

new instruments and accessories which enable the recording of vibrational spectra<br />

of samples under the most awkward experimental conditioiis which can never be<br />

attained by the other very popular new physical techniques.<br />

At present, infrared spectra can be obtained with fast and very fast FTIR interferometers<br />

with microscopes, in reflection and microreflection, in diffusion, at very<br />

low or very high temperature, in dilute solutions, etc. Ranian scattering can span a<br />

very large energy range in the excitation lines, thus reaching off-resonance or resonance<br />

conditions; spectrometers and interferometers can be used, microsampling is<br />

common, a few scattered photons can be detected with very sensitive CCD detectors<br />

and optical fibers provide a variety of new sampling procedures.<br />

Parallel to the technological development we watch the invasion (even on a<br />

coinmercial level) of theoretical and computational techniques (both ah initio or<br />

semiempirical) which, by simply pushing a button, provide vibrational frequencies<br />

and intensities and even show the animated wiggling of molecules on the screen of<br />

any personal computer.<br />

In spite of the wealth of experimental and theoretical data which are easily<br />

available for the study of molecules and of their behavior, other fields of physics<br />

and chemistry seem to have the priority in the teaching at Universities. Vibrational<br />

spectroscopy is, in general, no longer taught in detail and, at most, students are<br />

quickly exposed to molecular vibrations in some courses of analytical chemistry<br />

or structural organic chemistry. It is curious to notice that molecular dynamics is<br />

considered a very complicated mathematical machinery which must be avoided<br />

by simply mentioning briefly that inolecules ‘wiggle’ in a curious way. Then, the<br />

traditional old-fashioned structure-group frequency spectral correlations (mostly in<br />

infrared) are considered the only useful tool for the interpretation of spectra. This<br />

very limited spectroscopic culture is obviously transferred to industrial laboratories<br />

who rely on the use of vibrational spectra only for very simple chemical diagnosis<br />

and routine analytical determinations.<br />

The conclusion of this analysis is that the ratio: (number of inforniation)/(capa-


vi<br />

Prefucr<br />

bility of the experimental and theoretical techniques) turns out to be very small, in<br />

spite of the great potential offered by vibrational spectroscopy.<br />

The above analysis, shared by many spectroscopists in the field of small molecules,<br />

can be further expanded when vibrational spectroscopy is considered in the field of<br />

polymers and macromolecules in general. The wiggling of polymers adds new flavor<br />

to physics and chemistry. The translational periodicity of infinite polymers with<br />

perfect structure generates phonons and collective vibrations which give rise to<br />

absorption or Raman scattering bands that escape the interpretation based on the<br />

traditional spectroscopic correlations. The concept of collective motions forins the<br />

basis for the understanding of the vibrations of finite chain molecules which form a<br />

nonnegligible part of industrially relevant materials. On the other hand, real polymer<br />

samples never show perfect chemical, strereochemical, and confoimational<br />

structure. Symmetry is broken and new bands appear which become characteristic<br />

of specific types of disorder.<br />

If a few simple theoretical concepts of the dynamics of ordered and disordered<br />

chain molecules are taken into account it can be easily perceived that the vibrational<br />

infrared and Raman spectra contain a wealth of information essential to<br />

analytical polymer chemistry, structural chemistry, and physics.<br />

The content of this book has been planned to rejuvenate the vibrational spectroscopy<br />

of polymers. At present, the classes of polymeric materials are very many<br />

in number and range from classical bulk polymers of great industrial and technological<br />

relevance to highly sophisticated functional polymers which reach even the<br />

interest of photonics and molecular electronics. Also, the whole world of biopolymen<br />

requires great attention from spectroscopy.<br />

This book touches on a very few classes of polymeric materials which we consider<br />

representatives for introducing problems, spectroscopic techniques, and solutions<br />

prototypical of many other classes of polymers and plastics.<br />

Chapters 1 and 2 introduce new experimental techniques that provide new sets of<br />

relevant data for the study of the local and overall mobility of polymer chains. In<br />

Chapter 1, the development of two-dimensional infrared spectroscopy is described<br />

with a discussion of the mathematical principles, the description of the instrumental<br />

technique, and a detailed analysis of a few cases. In Chapter 2 the success of Fourier<br />

transform infrared polarization spectroscopy is shown for the study of segmental<br />

mobility in a polymer or a liquid-crystalline polymer under the influence of an<br />

external directional perturbation such as electric, electromagnetic, or mechanical<br />

forces. Industrial research laboratories should pay much attention to the information<br />

which can be acquired with this technique for technologically relevant polymeric<br />

materials.<br />

Chapter 3 is a guided tour of molecular dynamics and infrared and Raman frequency<br />

and intensity spectroscopy of polymethylene chains, from oligomers to polymers,<br />

in their perfect and disordered states. The reader is exposed in the easiest<br />

possible way to the basic theoretical concepts and to the numerical techniques<br />

which can be applied to such studies. Many references are provided for the spectroscopist<br />

who wants to develop his or her own independent skill and judgment.


However, theory and calculations yield concepts and data to be used also in polymer<br />

characterization, even for routine analytical work. After discussing several<br />

oligoniers or polymers the reader is guided step-by-step in the practical exercise<br />

to derive the detailed scenario of the mechanism of phase transition and melting of<br />

n-alkanes.<br />

Some of the theoretical tools laid in Chapter 3 are fully exploited in Chapter 4,<br />

which exposes the reader to the actual problems (and proposed solutions) of the<br />

chemistry and physics of modern and technologically relevant polyconjugated polymers<br />

in their intact (insulating) and in the doped (electrically conducting) states.<br />

Chapter 5 is an up-to-date review of the spectroscopic and structural problems<br />

and solutions reached by a modern approach to the dynamics of polypeptides. A<br />

clear and comprehensive discussion is presented on the force fields necessary for<br />

a reliable structural analysis through the vibrational spectra. Tools are thus becoming<br />

available for a systematic study of the structure even of complex polypeptides.<br />

The message we wish to send to the reader of this book is that modern machines<br />

provide beautiful infrared and Raman spectra of polymers full of specific, unique<br />

and detailed information which can be extracted, not just merely and lazily using<br />

group frequency correlations. We will be very pleased to find that this book has<br />

provided the reader with enough motivation to overcome the potential barrier of<br />

some theoretical technicalities in order to enjoy fully the wealth of information<br />

inherently contained in the vibrational spectra of ordered or disordered chain<br />

molecules.<br />

Milan, June 1998<br />

Giuseppe Zerbi


This Page Intentionally Left Blank


Contents<br />

1 Two-Dimensional Infrared <strong>Spect</strong>roscopy<br />

I. Noda, A. E. Downrey and C. Marcott<br />

1.1 Introduction<br />

1.2 Background<br />

1.3 Basic Properties of 2D Correlation <strong>Spect</strong>ra<br />

1.4 Instrumentation<br />

1.5 Applications<br />

2 Segmental Mobility of Liquid Crystals and Liquid-Crystalline<br />

<strong>Polymer</strong>s Under External Fields: Characterization by<br />

Fourier-Transform Infrared Polarization <strong>Spect</strong>roscopy<br />

H. W. Siesler, I. Zebger, Ch. Kulinna, S. Okretic, S. Shilov and<br />

U. Hoflinann<br />

2.1 Introduction<br />

2.2 Measurement Techniques<br />

2.3 Theory<br />

2.4 Structure Dependent Alignment of Side-Chain Liquid-Crystalline<br />

Polyacrylates on Anisotropic Surfaces<br />

2.5 Electric-Field Induced Orientation and Relaxation of<br />

Liquid-Crystalline Systems<br />

2.6 Alignment of Side-Chain Liquid-Crystalline Polyesters Under<br />

Laser Irradiation<br />

2.7 Orientation of Liquid-Crystals Under Mechanical Force<br />

2.8 Conclusions<br />

1<br />

1<br />

2<br />

10<br />

12<br />

15<br />

33<br />

33<br />

34<br />

36<br />

37<br />

38<br />

60<br />

67<br />

81<br />

3 Vibrational <strong>Spect</strong>ra as a Probe of Structural Order/Disorder in<br />

Chain Molecules and <strong>Polymer</strong>s 87<br />

G. Zerbi and M. Del Zoppo<br />

3.1 Introduction 87<br />

3.2 The Dynamical Case of Small and Symmetric Molecules 88<br />

3.3 How to Describe the Vibrations of a Molecule 94<br />

3.4 Short and Long Range Vibrational Coupling in Molecules 95<br />

3.5 Towards Larger Molecules: From Oligoiners to <strong>Polymer</strong>s 98<br />

3.6 From Dynamics to Vibrational <strong>Spect</strong>ra of One-Dimensional<br />

Lattices 107


X<br />

Corztrnts<br />

3.7 The Case of Isotactic Polypropylene: A Textbook Case<br />

3.8 Density of Vibrational States and Neutron Scattering<br />

3.9 Moving Towards Reality: From Order to Disorder<br />

3.10 What Do We Learn from Calculations<br />

3.11 A Very Simple Case: Lattice Dynamics of HCI-DCl Mixed<br />

Crystals<br />

3.12 Cis-Tram Opening of the Double Bond in the <strong>Polymer</strong>isation<br />

of Ethylene<br />

3.13 Defect Modes as Structural Probes in Polymethylene Chains<br />

3.14 Case studies:<br />

Case 1<br />

Conformational Mapping of Fatty Acids Through<br />

Mass Defects<br />

Case 2 Liquid Crystalline <strong>Polymer</strong>s: Polyesters<br />

Case 3 Chain Folding in Polyethylene Single Crystals<br />

Case 4 The Structure of the Skin and Core in Polyethylene<br />

Films (Normal and Ultradrawn)<br />

Case 5 Moving Towards More Complex Polymethylene<br />

Systems<br />

3.15 Simultaneous Configurational and Conformational Disorder.<br />

The case of Polyvynylchloride<br />

3.16 Structural Inhomogeneity and Raman <strong>Spect</strong>roscopy of<br />

LAM Modes<br />

3.17 Fermi Resonances<br />

3.18 Band Broadening and Conformational Flexibility<br />

3.19 A Worked Out Example: Froin N-Alkanes to Polyethylene.<br />

Structure and Dynamics<br />

4 Vibrational <strong>Spect</strong>roscopy of Intact and Doped Conjugated <strong>Polymer</strong>s<br />

and Their Models<br />

Y Furukawa and M. Tasuiiii<br />

4.1 Introduction<br />

4.2 Materials<br />

4.3 Geometry of Intact <strong>Polymer</strong>s<br />

4.4 Geometrical Changes Introduced by Doping<br />

4.5 Methodology of Raman Studies of Polarons, Bipolarons<br />

and Solitons<br />

4.6 Near Infrared Raman <strong>Spect</strong>roscopy<br />

4.7 Poly(p-phenylene)<br />

4.8 Other <strong>Polymer</strong>s<br />

4.9 Electronic Absorption and ESR <strong>Spect</strong>roscopies and Theory<br />

4.10 Mechanism of Charge Transport<br />

4.1 1 Summary<br />

5 Vibrational <strong>Spect</strong>roscopy of Polypeptides<br />

S. Kvinim<br />

5.1 Introduction<br />

113<br />

1 I7<br />

122<br />

128<br />

131<br />

138<br />

141<br />

144<br />

146<br />

147<br />

150<br />

152<br />

156<br />

158<br />

159<br />

163<br />

172<br />

180<br />

207<br />

207<br />

208<br />

209<br />

210<br />

215<br />

216<br />

217<br />

228<br />

230<br />

23 1<br />

234<br />

239<br />

239


Con ten t~<br />

xi<br />

Index<br />

5.2 Force Fields<br />

5.3 Amide Modes<br />

5.4 Polypeptides<br />

5.5 Summary<br />

240<br />

249<br />

257<br />

280<br />

287


This Page Intentionally Left Blank


Contributors<br />

M. Del Zoppo<br />

Dipartimento di Chimica Industriale<br />

Politecnico di Milaiio<br />

Piazza L. Da Vinci 32<br />

20133 Milano<br />

Italy<br />

A. E. Do?;rey<br />

The Procter and Gamble Company<br />

Miami Valley Laboratories<br />

P.O. Box 398707<br />

Cincinnati, OH 45239-8707<br />

USA<br />

Y. Furukawa<br />

Department of Chemistry<br />

School of Science and Engineering<br />

Waseda University<br />

Shinjuku-ku<br />

169 Tokyo<br />

Japan<br />

U. Hoffmann<br />

Bruins Instruments<br />

D 82178 Puchheim<br />

Germany<br />

S. Krimm<br />

Biophysics Research Division and<br />

Department of Physics<br />

University of Michigan<br />

Ann Arbor, MI 48109<br />

USA<br />

Ch. Kulinna<br />

Bayer AG<br />

D 40789, Monheim<br />

Germany<br />

C. Marcott<br />

The Procter and Gamble Company<br />

Miami Valley Laboratories<br />

P.O. Box 398707<br />

Cincinnati, OH 45239-8707<br />

USA<br />

I. Noda<br />

The Procter and Gamble Company<br />

Miami Valley Laboratories<br />

P.O. Box 398707<br />

Cincinnati? OH 45239-8707<br />

USA<br />

S. Okretic<br />

Fachhochschule Niederrhein<br />

Fachbereich 04<br />

D 7805 Krefeld<br />

Germany<br />

S. Shilov<br />

Institute of Macromolecular<br />

Compounds<br />

199004 St. Petersburg<br />

Russia<br />

H. W. Siesler<br />

Department of Physical Chemistry<br />

University of Essen<br />

D 45 1 17 Esseii<br />

Gerinaii y


xiv<br />

Coiitribirtovs<br />

M. Tasumi<br />

Department of Chemistry<br />

Faculty of Science<br />

Saitama University<br />

Urawa<br />

Saitama 338<br />

Japan<br />

I. Zebger<br />

Department of Physical Chemistry<br />

University of Essen<br />

D 451 17 Essen<br />

Germany<br />

G. Zerbi<br />

Dipartiniento di Chiinica Iiidustriale<br />

Politecnico di Milano<br />

Piazza L. Da Vinci 32<br />

20133 Milano<br />

Italy


1 Two-Dimensional Infrared (2D IR)<br />

<strong>Spect</strong>roscopy<br />

1.1 Introduction<br />

Two-dimensional infrared (2D IR) correlation spectroscopy [ 1-81 is a relatively new<br />

addition to the spectroscopic methods used for the characterization of polymers. In<br />

2D IR, infrared spectral intensity is obtained as a function of two independent<br />

wavenuinbers as shown in Figure 1-1. The well-recognized strength of IR spectroscopy<br />

arises from the specificity of an IR probe toward individual molecular vibrations<br />

which are strongly influenced by the local molecular structure and environment<br />

[9-121. Because of such specificity toward the submolecular state of a sample,<br />

surprisingly useful information about a complex polymeric system is provided by<br />

expanding IR analysis to the second spectral dimension.<br />

The initial idea of generating two-dimensional correlation spectra was introduced<br />

several decades ago in the field of NMR spectroscopy [13-161. Since then, numerous<br />

successful applications of multidimensional resonance spectroscopy techniques<br />

have been reported, including many different types of studies of polymeric materials<br />

by 2D NMR [17-191. However, until now the propagation of this powerful concept<br />

of inultidimensional spectroscopy in other areas of spectroscopy, especially vibrational<br />

spectroscopies such as IR and Rainan, has been surprisingly slow.<br />

One of the reasons why the two-dimensional correlation approach as applied in<br />

NMR spectroscopy was not readily incorporated into the field of IR spectroscopy<br />

was the relatively short characteristic times (on the order of picoseconds) associated<br />

with typical molecular vibrations probed by IR. Such a time scale is many orders<br />

of magnitude shorter than the relaxation times usually encountered in NMR. Consequently,<br />

the standard approach used so successfully in 2D NMR, i.e., multiplepulse<br />

excitations of a system followed by the detection and subsequent double<br />

Fourier transformation of a series of free induction decay signals, is not readily<br />

applicable to conventional IR experiments. A very different experimental approach,<br />

therefore, needed to be developed in order to produce 2D IR spectra useful for the<br />

characterization of polymers using an ordinary IR spectrometer.


Figure 1-1. A fishnet plot of a<br />

synchronous two-dimensional<br />

infrared (2D IR) correlation<br />

spectrum of atactic polystyrene<br />

in tlie CH-sti-etching vibration<br />

region at room temperature.<br />

1.2 Background<br />

1.2.1 Perturbation-Induced Dynamic <strong>Spect</strong>ra<br />

The basic scheme adapted for generating 2D IR spectra 13, 81 is shown in Figure<br />

1-2. In a typical optical spectroscopy experiment, an electromagnetic probe (eg,<br />

IR, X-ray, UV or visible light) is applied to the system of interest, and physical<br />

or chemical information about the system is obtained in the form of a spectrum<br />

representing a characteristic transformation (eg, absorption, retardation, and<br />

scattering) of the electromagnetic probe by the system constituents. In a 2D IR<br />

experiment, an external physical perturbation is applied to the system [2, 31 with the<br />

incident IR beam used as a probe for spectroscopic observation. Such a perturbation<br />

often induces time-dependent fluctuations of the spectral intensity, known as<br />

the dyrzamic spectrum, superposed onto the normal static IR spectrum.<br />

There are, of course, many different types of physical stimuli which could induce<br />

such dynamic variations in the spectral intensities of polymeric samples. Possible<br />

sources of perturbations include electrical, thermal, magnetic, acoustic, chemical,<br />

optical, and mechanical stimuli. The waveform or specific time signature of the<br />

perturbation may also vary from a simple step function or short pulse, to more<br />

complex ones, including highly multiplexed signals and even random noises. In this<br />

chapter, dynamic spectra generated by a simple sinusoidal mechanical perturbation<br />

applied to polymers will be discussed.<br />

Mechanical. eleclncal.<br />

opllcal. lliermal. elc<br />

Eleclro.magneltc<br />

probe (eg IR. UV)<br />

D y n a m I c<br />

Figure 1-2. A generalized experimental scheme for<br />

2D correlation spectroscopy based on perturbationinduced<br />

dynamic spectral signals [S).


1.2 Brick~qrorrnrl 3<br />

.<br />

Figure 1-3. Schematic diagram of the dynamic<br />

infrared linear dichroism I DIRLDi experiinent<br />

1231. A small-amplitude sinusoidal strain is<br />

applied to a sample. and submolecular level<br />

reorientation responses of cheniical moieties are<br />

monitored with a polarized IR beam.<br />

DETECTOR<br />

1.2.2 Dynamic IR Linear Dichroism (DIRLD)<br />

Figure 1-3 shows a schematic diagram of a dynamic IR linear dichroism (DIRLD)<br />

experiment [20-251 which provided the foundation for the 2D IR analysis of polymers.<br />

In DIRLD spectroscopy, a small-amplitude oscillatory strain ica. 0.1% of the<br />

sample dimension) with an acoustic-range frequency is applied to a thin polymer<br />

film. The submolecular-level response of individual chemical constituents induced<br />

by the applied dynamic strain is then monitored by using a polarized IR probe as<br />

a function of deformation frequency and other variables such as temperature.<br />

The macroscopic stress response of the system may also be measured simultaneously.<br />

In short, a DIRLD experiment may be regarded as a combination of two<br />

well-established characterization techniques already used extensively for polymers:<br />

dyizanzic inechnnicul unalysis (DMA)[26, 271 and ijlfrured dichroisin (IRD) spectroscopy<br />

[ 10, 113.<br />

The optical anisotropy, as characterized by the difference between the absorption<br />

of IR light polarized in the directions parallel and perpendicular to the reference<br />

axis (i.e., the direction of applied strain), is known as the IR linear dichroism of the<br />

system. For a uniaxially oriented polymer system [ 10, 28-30], the dichroic dzflerence,<br />

AA(v) = A,,(Y) - Al(v), is proportional to the average orientation, i.e., the<br />

second moment of the orientation distribution function, of transition dipoles (or<br />

electric-dipole transition moments) associated with the molecular vibration occurring<br />

at frequency 1’. If the average orientation of the transition dipoles absorbing<br />

light at frequency if is in the direction parallel to the applied strain, the dichroic<br />

difference AA takes a positive value; on the other hand, the IR dichroism becomes<br />

negative if the transition dipoles are perpendicularly oriented.<br />

If a sinusoidally varying small-amplitude dynamic strain is applied to a polymeric<br />

system, a similar sinusoidal change in IR dichroic difference, as shown in Figure<br />

1-4, is usually observed [3, 231. The dynamic variation of IR dichroism arises from<br />

the time-dependent reorientation of transition dipoles induced by the applied strain.<br />

Interestingly, however, the dichroism signal often is not fully in phase with the<br />

strain. There is a finite phase difference between the two sinusoidal signals representing<br />

the externally applied macroscopic perturbation and the resulting dynamic<br />

molecular-level response of the system. This phase difference is due obviously to


I I D W l<br />

0 : 2n 4n<br />

+ p H- Phase Angle<br />

0 2n 4n<br />

Figure 1-4. A small-amplitude sinusoidal<br />

strain applied to a sample and resulting<br />

dynamic IR dichroism I DIRLDI response<br />

induced by the strain. The two siiiusoidal<br />

signals are not always in phase with each<br />

other.<br />

-<br />

c<br />

n<br />

L<br />

m<br />

58msec<br />

Figure 1-5. A time-resolved<br />

dynamic IR dichroism<br />

(DIRLD I spectrum of an<br />

atactic polystyrene film under<br />

a small-amplitude (ca. 0. I'X))<br />

sinusoidal (23-Hzl dynamic<br />

strain at room temperature.<br />

the rate-dependent nature of the reorientation processes of various submolecular<br />

constituents.<br />

The advantage of using IR spectroscopy in dynamic studies of polymers is that<br />

such a measurement can be used to examine individual submolecular constituents<br />

or chemical functional groups by simply changing the wavelength of the IR probe.<br />

The variation of IR dichroism, depicted as a simple sinusoidal signal in Figure 1-4.<br />

can actually be measured as a function of not only time but also IR wavenumber.<br />

In other words, dynamic IR dichroism is obtained as a time-resolved spectrum<br />

representing the molecular level response of polymers, as shown in Figure 1-5.<br />

For a given sinusoidal dynamic strain i(t) with a small amplitude i and fre-


1.0 ,<br />

Figure 1-6. The in-phtrse<br />

and cprrt/r.ct/irr.c components<br />

of the DIRLD spectrum<br />

shown in Figure 1-5. A<br />

normal static absorbance<br />

spectrum of the sample is<br />

also provided for reference.<br />

Wavenumber<br />

quency w,<br />

E( t) = i sin ux (1-1)<br />

a time-resolved DIRLD spectrum may be represented by<br />

where AA(v) and P(v) are. respectively, the riiagnitude and phase (loss) angle of<br />

the dynamic IR dichroism [23]. By using a simple trigonometric identity, the above<br />

expression can also be rewritten in the form<br />

Ak(i9, t) = AA’(v) sin cot + AA”(v) coswt. (1-3)<br />

The wavenumber-dependent terms, AA’( I!) and AA”(v), are known respectively as<br />

the in-phase spectrum and quadrature spectrunt of the dynamic dichroism of the<br />

system. They represent the WN/ (storage) and irnagirrury (loss) components of the<br />

time-dependent fluctuations of dichroism. Figure 1-6 shows an example of the inphase<br />

and quadrature spectral pair extracted from the continuous time-resolved<br />

spectrum shown in Figure 1-5. These two ways of representing a DIRLD spectrum<br />

contain equivalent information about the reorientation dynamics of transition<br />

dipoles. However, the orthogonal representation of the time-resolved spectrum<br />

using the in-phase and quadrature spectra is obviously more compact and easier to<br />

interpret than the stacked-trace plot of the time-resolved spectrum.<br />

1.2.3 IR Dichroism and Molecular Orientation<br />

In a traditional characterization study of polymeric materials, the IR dichroism<br />

technique is most often employed for the determination of the degree of orientation


Reference<br />

axis<br />

Molecular<br />

segment<br />

Transition<br />

dipole<br />

Figure 1-7. Uniaxial orientation of a polymei- chain segment with respect<br />

to a reference axis and corresponding alignment of a transition dipole.<br />

of molecular chain segments [lo, 28-30]. For a uniaxially oriented polymer system<br />

(Figure 1-7), the orimtatioii Jbctor P?(Q), i.e., the second moment of the orientation<br />

distribution function of polymer chain segments, is given by<br />

PZ(0) = (_?(COS? 6,) - 1)/2 (1-4)<br />

where 6, is the orientation angle between each polymer chain segment and the optical<br />

reference axis of the system. The notation (cos2 Q) indicates that the squared-cosine<br />

of orientation angles for individual segments are averaged over the entire space.<br />

The orientation factor Pz(B) is a convenient measure of the average degree of<br />

orientation of polymer chains in the system. Under an ideal orientation state where<br />

all polymer chains are aligned perfectly in the direction parallel to the reference<br />

axis, the value of P2(6,) becomes unity. On the other hand, if polymer chain segments<br />

are all perpendicularly aligned, P2(8) becomes - 1/2. An optically isotropic<br />

system gives the Pl(0) value of zero.<br />

According to the classical theory of the IR dichroisiii of polymers [lo, 28, 291, the<br />

diclzroic ratio, D(v) = AII(V)/A~(V), of the system is directly related to the orientation<br />

of polymer chain segments by<br />

D(V) - 1 Ow(.) + 2<br />

PI(@) = (1-5)<br />

D(v) +2 Dw(v) - 1<br />

The ultinzate dichroic ratio D, (v) is the value of dichroic ratio if the polymer chain<br />

segments are all perfectly aligned in the direction of the reference axis, i.e.,<br />

Pz(S) = 1. Given the local orientation angle a between the polymer chain segment<br />

and transition dipole probed at the IR wavenumber v (Figure 1-7), the ultimate<br />

dichroic ratio is given by<br />

2<br />

DN,(V) = 2cot a.<br />

(1-6)<br />

It can be easily shown that Eq. (1-5) simplifies to the form


1.2 Bcrckgroinizd 7<br />

where AA (i~) is the iilti/ncrte dichroic d~juwrce which is related to D, ( v) by<br />

The ~ t r ~ ~ absorbcince, r ~ w d A,( if) = [A (1)) + 2AI (v)]/3, is believed to be independent<br />

of the degree of molecular orientation as long as the system remains symmetric<br />

with respect to the optical axis.<br />

If one accepts that the transition dipole orientation angle ci is a fixed molecular<br />

constant unaffected by the conformation of polymer chains, then D, and AAz<br />

should also be independent of the state of chain orientation. The classical theory<br />

of IR dichroisin for polymers thus makes an interesting assertion, according to Eq.<br />

(1-7), that IR dichroic difference AA(v) is alivays linearly proportional to the average<br />

orientation Pz(0) of polymer chains regardless of the IR wavenumber v. In<br />

other words, one should be able to determine the state of orientation of the entire<br />

polymer chain by simply observing the local orientation of a transition dipole<br />

associated with the molecular vibration of any arbitrarily chosen local functional<br />

group, as long as ci is known.<br />

1.2.4 Breakdown of the Classical Theory<br />

An interesting observation often made during the DIRLD measurement of polymers<br />

is that the phase angle p(~) between the applied strain and dynamic IR<br />

dichroic diff‘erence is strongly dependent on the IR wavenumber [3, 33-25]. Timedependent<br />

dichroism intensities measured at certain wavenumbers change much<br />

faster than others. Individual dynamic IR dichroisin signals thus become out of<br />

phase with each other as depicted in Figure 1-8. As already demonstrated in Figure<br />

1-6, the shape of the in-phase spectrum also becomes quite different from that of the<br />

quadrature spectrum. The relative amount of dynamic dichroism signal appearing<br />

in the in-phase and quadrature components varies considerably as a function of<br />

wavenumber.<br />

t<br />

Figure 1-8. Tinie-dependent sinusoidal u<br />

DIRLD signals from an atactic poly- 5<br />

(methyl methacrylate) sample detected at 0”<br />

different IR wavenumbers. Signals are out<br />

of phase with each other.<br />

a:<br />

(2952 cm-1)<br />

r<br />

Time, t


Quadrature 0<br />

+<br />

3150 3<br />

Wavenumber<br />

Figure 1-9. DIRLD spectra of atactic polystyrene in the phenyl<br />

CH-stretching region. Transition dipoles for IR bands marked<br />

with + are reorienting totally in phase with the applied dynamic<br />

strain, while those with 0 are moving at rates substantially<br />

diffei-ent from the strain.<br />

This surprising discovery reveals that a inajor discrepancy exists between the<br />

classical theory of IR dichroisni and actual experimental observations made for<br />

reorientation dynamics of polymer chains. Figure 1-9, for example, indicates that<br />

the intensity of the quadrature spectrum is close to zero, i.e., signals are in phase<br />

with the applied strain for dichroism peaks marked with (t). If indeed IR dichroism<br />

signals at these wavenumbers reflect the time-dependent orieiitational state of<br />

the entire polymer chain, one would conclude that polymers are reorienting instantaneously<br />

under the dynamic deformation. On the other hand, if the dichroism signals<br />

are measured for other IR bands, for example, at the peaks inarked with (O),<br />

the polymer chain seems to reorient at a rate substantially out of phase with tlie<br />

applied strain. This second statement clearly contradicts the conclusion made as a<br />

result of the previous observation. Thus, for a certain polymer system undergoing a<br />

dynamic deformation, the well-accepted classical view of IR dichroisin (i.e.,<br />

dichroic difference must always be linearly proportional to the average orientation<br />

of polymer chains, regardless of the wavenuinber of IR probe, as described in Eq.<br />

(1-7)) is no longer valid.<br />

The logical explanation for tlie experimentally observed, wavenumber-dependent<br />

behavior of the DIRLD phase angle is that the reorientation rates of individual<br />

traiisitioii dipoles in the system are not the same. For a given macroscopic perturbation<br />

such as dynamic strain, different submolecular constituents (e.g., backbone<br />

segments, side chains, and various functional groups comprising the polymer chain)<br />

may respond at different rates, more or less independently of each other. Consequently,<br />

transition dipoles associated with the molecular vibrations of different<br />

parts of a polymer chain have independent local reorieiitational responses not fully<br />

synchronized to tlie global motioiis of tlie entire polymer chain.


1.2 Background 9<br />

1.2.5 Two-Dimensional Correlation Analysis<br />

The wavenumber-dependent orientation rates of individual transition dipoles<br />

observed for dynamically stimulated polymer systems poses several intriguing<br />

questions: What makes these transition dipoles reorient at different rates? Why do<br />

some of the transition dipoles seem to reorient at a rate similar to each other? Is<br />

there any underlying mechanism responsible for synchronization, or lack thereof, in<br />

the local reorientation processes of submolecular structures‘? To answer such questions,<br />

one must introduce an effective way of representing a measure of the siniilarity<br />

oi- difference of the reorientation rates of transition dipoles.<br />

One convenient mathematical approach to comparing the behavior of a set of<br />

variables is a correlation analysis [2, 3, 81. For a pair of dynamic dichroisni signals,<br />

Aj(v1, f) and AA(v2, t + T ), which are observed at different instances separated by<br />

a fixed correlation time z at two arbitrarily selected wavenumbers, v1 and vz, for a<br />

certain length of observation period T, a two-dimerisionul c‘ross-correlation fiiriction<br />

X(z) is defined by<br />

For a pair of sinusoidally varying signals with a fixed frequency 01, the crosscorrelation<br />

function reduces to a simple form [3],<br />

X(z) = @(vl,v2)cosoz+ Y(vl,v2)sincoz. (1-10)<br />

The real and iimqinnry components, @(vl, 1’2) and Y(v1, v2), of the cross-correlation<br />

function X(z) are referred to, respectively, as the synclzroiioiis and asyiichronoLis<br />

correkition intensitji. These quantities are related to the in-phase and quadrature<br />

spectra of dynamic dichroism by<br />

(1-1 1)<br />

The synchronous correlation intensity @( VI ,v2) represents the siniilurity between the<br />

two dynamic dichroisni signals measured at dif‘fereiit wavenuinbers. This quantity<br />

becomes significant if the two dynamic IR dichroism sigiials are changing at a similar<br />

rate but vanishes if the time dependency of the signals is very different. The<br />

asynchronous correlation intensity ‘€‘(vl, vz), on the other hand, represents the dissiniilmity<br />

between signals. It becomes significant only if the two IR dichroism signals<br />

are changing out of phase with each other but reduces to zero if they are<br />

changing together.<br />

The two-dimensional nature of our correlation analysis arises from the fact<br />

that the correlation intensities are obtained by comparing the time dependence of


Wavenumber, vI<br />

Figure 1-10. A schematic contour diagram ofa synchronous<br />

2D IR correlation spectrum 131. Shaded areas represent<br />

negative correlation intensity.<br />

I<br />

n<br />

Wavenumber, v,<br />

Figure 1-11. A schematic contour diagram of an asynchronous<br />

2D IR correlation spectrum [3]. Shaded areas represent<br />

negative correlation intensity.<br />

dynamic IR signals observed at two independently chosen wavenumbers. Plots of<br />

correlation intensities as functions of two wavenumber axes are referred to as 2D<br />

IR correlution spectra. Such spectra reveal important and useful information not<br />

readily accessible from conventional one-dimensional spectra.<br />

There are many different ways to plot a 2D IR correlation spectrum. A pseudothree-dimensional<br />

representation (so-called fishnet plot) as shown in Figure 1-1 is<br />

well suited for providing the overall features of a 2D IR spectrum. Usually, however,<br />

2D IR spectra are more conveniently displayed as contour maps to indicate<br />

clearly the location and intensity of peaks on a given spectral plane. In the following<br />

section, basic properties of and information extracted from 2D IR spectra are<br />

reviewed by using schematic contour maps (Figures 1-10 and 1-1 1).<br />

1.3 Basic Properties of 2D IR Correlation <strong>Spect</strong>ra<br />

1.3.1 Synchronous 2D IR <strong>Spect</strong>rum<br />

The correlation intensity at the diagonal position of a synchronous 2D spectrum<br />

(Figure 1-10) corresponds to the autocorrelation function of perturbation-induced


1.3 Bnsic Properties of 20 IR Correlotioii <strong>Spect</strong>ra 11<br />

dynamic fluctuations of the IR signals. Local intensity maxima along the diagonal<br />

are thus referred to as autopolis. Since the magnitude of dynamic variations of IR<br />

dichroism represents the susceptibility of transition dipoles to reorient under a given<br />

external perturbation, autopeaks indirectly reflect the local mobility of chemical<br />

groups contributing to the inolecular vibrations associated with the transition<br />

dipoles. The schematic example in Figure 1-10 indicates that functional groups<br />

contributing to the molecular vibrations with frequencies A, B, C, and D are undergoing<br />

reorientational motions induced by the applied dynamic strain.<br />

Peaks located at off-diagonal positions of a 2D correlation spectrum are called<br />

cyoss peaks. They appear when the dynamic variations of IR signals at two different<br />

wavenumbers corresponding to the spectral coordinates (111 , ~ 2 are ) correlated with<br />

each other. For a synchronous spectrum, this occurs when the two IR signals are<br />

fluctuating in phase (i.e., simultaneously) with each other. Synchronized variations<br />

of dichroism signals result from the simultaneous reorientation of transition dipoles<br />

associated with corresponding IR wavenumbers. Coordinated local motions of<br />

submolecular groups lead to such simultaneous reorientations. In turn, highly correlated<br />

local reorientations of chemical groups in response to a common external<br />

stimulus imply the possible existence of interactions or connectivity which restrict<br />

the independent motions of these submolecular structures.<br />

Functional groups which are not strongly interacting, on the other hand, can<br />

move independently of each other. The transition dipoles associated with molecular<br />

vibrations of these groups may then reorient at different rates (or somewhat out of<br />

phase with each other), resulting in a much weaker synchronous correlation. Thus,<br />

as long as the normal modes of vibrations correspond to reasonably pure group<br />

frequencies, one can use the cross peaks in a synchronous 2D IR spectrum to map<br />

out the degree of intra- and intermolecular interactions of various functional<br />

groups. In Figure 1-10, functional groups contributing to the molecular vibrations<br />

with frequencies A and C may be interacting. Likewise, the pair B and D could be<br />

connected.<br />

The signs of cross peaks indicate relative reorientation directions of transition<br />

dipoles and consequently their associated chemical groups. If the sign of a synchronous<br />

cross peak is positive, the corresponding pair of transition dipoles reorient<br />

in the same direction. If negative, on the other hand, the reorientation directions are<br />

perpendicular to each other. In Figure 1 - 10, mutually perpendicular reorientation<br />

of a pair of transition dipoles at wavenumbers A and C, as well as parallel reorientation<br />

for B and D, are observed.<br />

1.3.2 Asynchronous 2D IR <strong>Spect</strong>rum<br />

An asynchronous 2D IR spectrum (Figure 1-1 1) provides complementary information;<br />

cross peaks appear if IR signals are not synchronized to be completely in<br />

phase with each other. This feature is particularly useful. since IR bands arising<br />

from molecular vibrations of different functional groups, or even of similar group5<br />

in different local environments, may exhibit substantially different time-dependent<br />

intensity fluctuations. Thus, asynchronous 2D IR spectra can be used to differen-


tiate highly overlapped IR bands, since asynchronous cross peaks should develop<br />

among these bands.<br />

From the sign of asynchronous cross peaks, one can obtain temporal information<br />

about the perturbation-induced reorientation processes of transition dipoles and<br />

their corresponding chemical groups. A positive peak in an asynchronous spectrum<br />

indicates the transition dipole with the vibrational frequency 1’1 reorients hc;fbr.r that<br />

for 1’2. If the sign is negative, reorients ufi~r<br />

1’2. However. this temporal relationship<br />

is reversed for perpendicularly reorienting pairs of transition dipoles, i.e., the<br />

synchronous correlation intensity at the same spectral coordinate becomes negative.<br />

In Figure 1-1 1, the reorientational motions of functional groups contributing to the<br />

molecular vibrations with frequencies B and D occur before those for A and C. More<br />

detailed discussions on the properties of 2D 1R spectra are found elsewhere [3).<br />

1.4 Instrumentation<br />

Dynamic infrared spectra suitable for 2D correlation analysis can, in principle, be<br />

measured using any conventional IR <strong>Spect</strong>rometer [ 71. The spectrometer, however,<br />

must be equipped with the ability to stimulate samples by some physical means and<br />

measure the resulting time-dependent fluctuations of the IR signals. Both dispersive<br />

monochromators (231 and Fourier transform infrared (FT-IR) instrumentation 151<br />

have been successfully used to measure dynamic 2D 1R spectra. Although welldesigned<br />

dispersive spectrometers can often achieve better signal-to-noise ratios<br />

(S/N) over small spectral regions, FT-IR measurements cover much broader spectral<br />

regions in less time and are, more importantly, readily available commercially.<br />

Conventional rapid-scan FT-IR spectrometers are not well suited for these types<br />

of measurements when the dynamic strain frequencies of interest are in the range<br />

between 0.1 Hz and 10 kHz. Using a step-scanning interferometer substantially<br />

simplifies dynamic 2D FT-IR measurements. Figure 1-12 shows an example of a<br />

n<br />

Rolahng-Blade<br />

chopper. wc<br />

nred<br />

P*er<br />

Figure 1-12. A schematic view<br />

of a DIRLD spectrometer based<br />

on a small-amplitude mechanical<br />

deformation 1231.


2D IR spectrometer coupled with a dynamic rheometer capable of applying a smallamplitude<br />

mechanical perturbation [23].<br />

In addition to the time-resolution constraints on dynamic infrared measurements.<br />

sensitive instrumentation is also required. The signals of interest, which are often<br />

lo4 times smaller than the normal IR absorbance of the sample, exist only as ii<br />

result of the small dynamic strain applied to the sample.<br />

The type of perturbation applied to the sample is not restricted to simple<br />

mechanical strain. Electrical perturbations, for example, can be effectively used to<br />

induce the necessary fluctuations of IR signals to produce 2D IR spectra. The use of<br />

electrical stimuli has been especially successful in studies of neniatic liquid crystalline<br />

systems [3 1-33], where selective orientation of liquid crystals under an alternating<br />

electric field was observed. Electrochemical experiments modulated with an<br />

alternating electrical current have also been used to generate dynamic 1R spectra<br />

suitable for 2D correlation analysis [34, 351. Recently, a two-dimensional photoacoustic<br />

spectroscopy (2D PAS) experiment was conducted with a step-scanning<br />

FT-IR spectrometer to obtain depth profiles of layered samples [36, 371. The characteristic<br />

time dependence of PAS signals representing sample components located<br />

at different depths inside samples were successfully differentiated by 2D correlation<br />

analysis. Each of these applications is similar in that a dynamic perturbation is<br />

applied to the sample and the time-resolved response is measured using phasesensitive<br />

detection [ 6 -81.<br />

1.4.1 Dispersive 2D IR <strong>Spect</strong>rometer<br />

A high-optical-throughput dispersive monochromator turns out to be an excellent<br />

choice for dynamic 2D IR spectroscopy [3, 231. As shown in Figure 1-12, the incident<br />

IR beam, originating from a high-intensity source, is modulated at three separate<br />

places by a chopper, photoelastic modulator (PEM), and dynamic mechanical<br />

analyzer (i.e., polymer stretcher). The chopper labels photons originating from the<br />

source so they can be distinguished from background IR emission. The PEM, which<br />

immediately follows a fixed linear polarizer in the beam path, enables the polarization<br />

direction of the beam to be switched back and forth rapidly (-100 kHz) between<br />

directions aligned parallel and perpendicular to the sample strain direction.<br />

Sample strain frequencies used are typically between 0.1 and 100 Hz. Each niodulation<br />

frequency is thus separated from the other two modulation frequencies by at<br />

least one order of magnitude. This makes analysis of the individual signal coniponents<br />

with lock-in amplifiers straightforward [23].<br />

Figure 1-13 shows an example of a train of lock-in amplifiers used to demodulate<br />

DIRLD signals obtained with a spectrometer described in Figure 1-12. In order to<br />

obtain the time-dependent dynamic absorbance and DIRLD responses, quadrature<br />

lock-in amplifiers are used. These devices monitor signals both in phase and 90" out<br />

of phase (quadrature) with the sinusoidal strain reference signal. The monochromator<br />

is scanned one wavelength at a time through the spectrum. Data are collected<br />

on six separate channels (e.g., in-phase and quadrature dynamic dichroism, in-phase<br />

and quadrature dynamic absorbance, static dichroism, and normal IR absorbance)


Slalic dichroism<br />

In-phase dynamic dlchrolsm<br />

Quadrature dynamic dichrolsm<br />

Preamp<br />

- Lockin<br />

D<br />

-<br />

VD Slatic absorbance<br />

at each wavenumber position until an acceptable S/N is achieved. Multiple scans<br />

of the entire spectrum can be collected in order to further improve the S/N and to<br />

average out longer-tenii drift.<br />

1.4.2 Step-Scanning 2D FT-IR <strong>Spect</strong>rometer<br />

The dynamic 2D IR spectroscopy experiment performed on a step-scanning FT-IR<br />

spectrometer is in many ways similar to the way it is done on a dispersive instrument<br />

[5, 71. For example, a step-scanning interferometer measures an interferogram<br />

one mirror retardation position at a time in the same way a monochromator scans<br />

through a spectrum one wavelength at a time. The mirror remains in a fixed retardation<br />

position while data are accumulated from the output channels of lock-in<br />

amplifiers tuned to the signals of interest (e.g., in-phase and quadrature dynamic<br />

absorbance, normal 1R absorbance). As in the dispersive experiment, signals are<br />

time averaged long enough to achieve an acceptable S/N and multiple scans of the<br />

entire interferogram can be coadded to average out longer-term drift. When a scan<br />

is completed, the resulting interferogram does not depend on the moving mirror<br />

velocity, as in the case of an interferograni collected in rapid-scan mode. The<br />

Fourier frequencies in a step-scanning experiment are typically in the sub-Hz range<br />

where they are well separated from the polarization modulation and strain modulation<br />

frequencies. This is in contrast to a typical rapid-scan FT-IR experiment<br />

where each IR wavelength is modulated at a different acoustic-range frequency.


1.5 Applirtrtions 15<br />

Unlike a dispersive spectrometer which uses a mechanical light chopper, the stepscanning<br />

interferometer often performs better when a technique called phase niodu-<br />

Irr/ion is used to label the IR photons originating from the source. Phase modulation<br />

is achieved by dithering either the fixed or moving mirror of the interferometer<br />

a small amount (on the order of 1 pi) at a fixed frequency. This motion causes<br />

a modulation of the IR signal intensity due to a change in optical path difference<br />

between the two arms of the interferometer. The amplitude of the signal detected is<br />

thus the slope of the actual interferogram at the average optical path difference.<br />

Most of the early measurements of DIRLD spectra were made using a photoelastic<br />

modulator IPEM) to modulate the polarization rapidly between parallel and<br />

perpendicular to the dynamic strain direction. Although polarization modulation<br />

does improve the S/N of dynamic IR spectroscopy, by about 50% over experiments<br />

using only a fixed polarizer aligned parallel to the sample strain direction, a PEM<br />

adds significant complexity to the experiment. More lock-in amplifiers and careful<br />

optical alignment are required when using polarization modulation. Indeed the first<br />

2D IR spectra reported in the literature [2] were obtaining using unpolarized light.<br />

When a PEM is not in use, the step-scan FT-IR experiment will work equally well<br />

with either a room-temperature deuterated triglycine sulfate (DTGS) detector or<br />

a liquid-nitrogen-cooled mercury-cadmium-telluride (MCT) detector. Although<br />

MCT detectors are typically an order of magnitude more sensitive than DTGS<br />

detectors, the full throughput of a commercial FT-IR spectrometer is enough to<br />

saturate the most sensitive MCT detectors. Thus, the beam is normally attenuated<br />

by using smaller apertures, neutral-density filters, or optical filters when using an<br />

MCT detector.<br />

Although dispersive dynamic IR experiments are presently more sensitive than<br />

FT experiments over limited spectral ranges, recent introductions of commercial<br />

FT-IR spectrometers with step-scanning capability should make dynamic 2D IR<br />

spectroscopy accessible to many more laboratories.<br />

1.5 Applications<br />

2D IR spectroscopy has been applied extensively to studies of polymeric materials.<br />

A recent review of 2D IR spectroscopy cites nunierous applications in the study of<br />

polymers by this technique [6]. In this section, some representative examples of 2D<br />

IR analysis of polymers are presented. We will start our discussion with a simple<br />

homogeneous amorphous polymer then move to more complex multiphase systems,<br />

such as seinicrystalline polymers. Alloys and blends consisting of more than one<br />

polymer components are of great scientific and technical importance. Both immiscible<br />

and miscible polymer blend systems may be studied by 2D IR spectroscopy.<br />

Analysis of microphase-separated block copolymers is also possible. Finally, the<br />

possible application of 2D IR spectroscopy to the studies of natural polymers of<br />

biological origin is explored.


-<br />

,<br />

i<br />

n<br />

-2950<br />

'<br />

E, Figure<br />

- ><br />

1-14. A synchronous 2D IR<br />

spectrum of a thin film of atactic<br />

polystyrene in the CH-stretching vibration<br />

region at room temperature.<br />

Regular one-dimensional spectra of<br />

the same system are provided at the<br />

top and left of the 2D spectrum for<br />

, I I l I I I I I 315" reference. Shaded areas represent<br />

I50 2950 2750<br />

Wavenumber. VI<br />

negative correlation intensity.<br />

1.5.1 Amorphous <strong>Polymer</strong>s<br />

A homogeneous one-phase system consisting of a single. noncrystallizing homopolymer,<br />

such as atactic polystyrene [3, 38-41 ] or poly(methy1 methacrylate) [42-<br />

441, provides one of the least complicated testing grounds for the applicability of<br />

the 2D IR technique for the study of local dynamics of polymers. By using 2D IR,<br />

surprisingly rich information can be extracted even from a simple polymeric system<br />

[25, 39-41 1. Figure 1-14 shows a contour-map representation of the synchronous<br />

2D IR correlation spectrum of an atactic polystyrene film in the CH-stretching<br />

region. A three-dimensional fish-net plot of the same 2D IR spectrum corresponding<br />

to Figure 1-14 has already been shown in Figure 1-1.<br />

DIRLD data used for the 2D correlation analysis was obtained by applying a<br />

23-Hz dynamic strain with an amplitude of 0.1% to a thin solution-cast film of atactic<br />

polystyrene (MI, 150,000) at room temperature 1391. The resulting dynamic dichroism<br />

spectra have already been shown in Figures 1-5 and 1-6. The time-dependent<br />

reorientations of transition dipoles associated with the molecular vibrations of<br />

backbone and side groups are all observable in the original dynamic dichroism<br />

spectra. However, much more detailed features are recognized in the 2D correlation<br />

spectrum.<br />

Autopeaks at the diagonal positions of Figure 1-14 represent transition dipoles<br />

that are susceptible to reorientational motions in the polystyrene sample. Peaks at<br />

2855 cm-I and 2925 cm-l are, respectively, assignable to the symmetric and antisymmetric<br />

CH2-stretching vibrations of the methylene groups in the backbone of<br />

the polystyrene chain; those peaks above 3000 cm-' arise from the phenyl ring CHstretching<br />

vibrations and indicate the reorientation of the side groups. The appearance<br />

of positive synchronous cross peaks at spectral coordinates corresponding to<br />

the two methylene bands indicates the transition dipoles associated with these two


\<br />

'\<br />

Melliyleiie<br />

1'7-<br />

-3000<br />

Posilivo cross peaks<br />

(phenyl 4 rnelhylena)<br />

,<br />

L<br />

a,<br />

n<br />

-3075 5<br />

C<br />

9<br />

- 2<br />

Figure 1-15. An asynchronous 2D 1R<br />

spectrum of a atactic polystyrene film<br />

in the CH-stretching vibration region<br />

I I I I I I I I I I<br />

3150<br />

3000 2900 2800<br />

wavenumbers reorient at a similar rate. Such synchronized responses are of course<br />

expected, since both 1R bands originate from vibrations of the same submolecular<br />

structural unit.<br />

The positive sign of the methylene cross peaks suggests the transition dipoles for<br />

the symmetric and antisymmetric methylene vibrations are both realigning in the<br />

same relative direction. i.e., perpendicular to the direction of applied strain. The<br />

molecular architecture dictates that these two transition dipoles are locally aligned<br />

more or less perpendicular to the backbone of the polymer segment. It is not difficult<br />

to conclude that the molecular chain segment of polystyrene must be realigning<br />

in the direction parallel to the applied dynamic tensile strain.<br />

Negative (shaded) synchronous cross peaks between the methylene and phenyl<br />

bands indicate that some of the phenyl transition dipoles are orienting parallel to<br />

the strain direction. It is known [40, 411 that the signs of these particular cross peaks<br />

for atactic polystyrene become positive at temperatures above the glass-to-rubber<br />

transition temperature (ca. 100 "C). The inversion of the cross-peak sign, and consequently,<br />

the relative reorientation direction of the side group with respect to the<br />

main chain, indicates a dramatic change in the local mobility of the polymer segments<br />

and functional groups is occurring around the glass transition temperature.<br />

Figure 1-15 shows the asynchronous 2D IR spectrum of the same sample comparing<br />

the reorientation dynamics of transition dipoles for backbone methylene and<br />

side-group phenyl. The spectral region corresponding to the asynchronous correlation<br />

of methylene stretching vibrations below 3000 cm-' is not shown since no<br />

asynchronous cross peaks are observed. The appearance of strong asynchronous<br />

cross peaks between phenyl and inethylene bands suggests the existence of rather<br />

complex reorientation dynamics involving the side groups. The rate of reorientational<br />

motion of the side groups under dynamic strain is apparently different from<br />

that of the backbone for this polymer. The signs of asynchronous cross peaks shown


in this 2D spectrum are all positive. Coupled with the fxt that the synchronous<br />

correlation intensities at corresponding coordinates i Figure 1 - 14) are all negative,<br />

one can conclude the side groups of this glassy polystyrene sample complete the<br />

realignment induced by the dynamic strain well before the main chain.<br />

This observation cannot be explained if the average local orientation angle of the<br />

phenyl groups with respect to the chain axis of polystyrene is fixed. There apparently<br />

is substantial freedom of the side groups to temporarily realign ahead of the<br />

main chain 141 1. This finding clearly demonstrates that it is impossible to monitor<br />

the main-chain dynamics of polystyrene by simply observing the reorientation of<br />

side groups. Furthermore, since phenyl groups contribute significantly toward the<br />

refractive index of polystyrene, the above observation also casts significant doubt<br />

over the validity of directly relating birefringence results of glassy polystyrene under<br />

a dynamic deformation to main-chain orientation dynamics.<br />

The observation of highly localized reorientational motions of functional groups<br />

in glassy amorphous polymers is not limited to atactic polystyrene. Similar independent<br />

motions of different side groups are observed in many other systems, including<br />

atactic poly(methy1 methacrylate) [42-441. Different side groups attached to<br />

the same polymer chain, such as ester methyl and a-methyl groups of polyjmethyl<br />

methacrylate). exhibited completely different reorientation rates (see Figure 1-8).<br />

2D IR spectroscopy has been especially useful in elucidating submolecular-level<br />

realignment of functional groups controlling the mechanical properties of glassy<br />

amorphous polymers [4, 43).<br />

The relationship between the local mobility of functional groups and glasstransition<br />

phenomenon has also been probed successfully. New insight into the<br />

dynamics of amorphous polymers going through glass-to-rubber transition processes<br />

was provided. Specifically, the local mobility of individual functional groups<br />

was found to play a very significant role in the glass-to-rubber transition of amorphous<br />

polymers [41]. In turn, it is now possible to determine if an amorphous<br />

polymer is in glassy or rubbery state by simply observing the characteristic local<br />

reorientation dynamics of functional groups using 2D IR spectroscopy.<br />

1.5.2 Semicrystalline <strong>Polymer</strong>s<br />

Semicrystalline polymers, such as polyethylene 145-471 and polypropylene (5, 481,<br />

may also be studied by using the 2D IR technique. By taking advantage of the<br />

enhanced spectral resolution of 2D IR, overlapped IR bands assigned to the coexisting<br />

crystalline and amorphous phases of semicrystalline polymers can be easily<br />

differentiated. Such differentiation has become especially useful, for example, in the<br />

study of blends of high-density polyethylene and low-density polyethylene [47).<br />

Here it was found that blends of polyethylenes are mixed at the molecular scale<br />

only in the amorphous phase, while each component crystallizes separately. In this<br />

section. an example of a 2D IR analysis applied to a film of linear low-density<br />

polyethylene is discussed [46].<br />

Figure 1-1 6 shows the asynchronous 2D IR spectrum of a thin film of linear lowdensity<br />

polyethylene consisting mainly of ethylene repeat units with a small amount


11 Amorphous Crystalline ,...." I<br />

Figure 1-16. An asynchronous ZD IR<br />

spectrum for the main chain syminetric<br />

CH2-stretching vibration of linear<br />

low-density polyethylene [46].<br />

\ I...."<br />

I I , , 1 1 -2870<br />

2870 2855 2840<br />

Wavenumber, VI<br />

(ca. 3 mole'%,) of deuterium-substituted 1-octene comonomer units. The octene<br />

comonomer is incorporated into the polyethylene chain to produce short (sixcarbon)<br />

side branches. which reduce the melt temperature and crystallinity of the<br />

polymer. By selectively labeling the octene units with deuterium, one can unambiguously<br />

differentiate the dynamics of main chain and short side branches. The<br />

spectral region shown in Figure 1-16 represents IR signals arising exclusively from<br />

the polyethylene chain, with little contribution from the octene side branches.<br />

The presence of asynchronous peaks in the CHZ-stretching vibration of backbone<br />

methylene groups indicates there are two distinct bands in this region of the IR<br />

spectrum of the linear low density polyethylene sample. The band located near<br />

2855 cm-' can be assigned to the contribution from the crystalline phase of the<br />

sample. while the band near 2863 cm-' is due to the amorphous component of this<br />

semicrystalline polymer. The above assignment can be easily verified by heating the<br />

sample above its melt temperature to remove the IR spectral contribution from<br />

the crystalline component of the specimen. Because of the significant difference in<br />

the reorientational responses of polymer chains located in the crystalline and<br />

amorphous phase domains to a given dynamic deformation. 2D IR spectroscopy<br />

can easily differentiate IR bands associated with the two-phase domains, even<br />

though these bands are highly overlapped.<br />

Figure 1-17 shows an asynchronous 2D IR spectrum correlating the backbone of<br />

polyethylene segments located in either crystalline and amorphous domains and<br />

deuterium-substituted branches. Asynchronous cross peaks develop between the<br />

crystalline component of the symmetric CH2-stretching band of the polyethylene<br />

segments and bands for octene side branches. Such asynchronicity suggests short<br />

side branches can move independently of the polyethylene chain located in the<br />

crystalline phase of the sample. No noticeable asynchronicity. however, is observed


20<br />

I<br />

2840<br />

chain<br />

-<br />

Crystal<br />

line<br />

phous<br />

t ..<br />

Figure 1-17. An asynchronous 2D IR<br />

spectrum of linear low-density poly-<br />

~<br />

ethylene comparing the strain-induced<br />

, , , I , , , , -2870<br />

50 2150 2050 local reorientational motions of the<br />

Wavenumber, VI main chain and side chain [46].<br />

between the reorientation dynamics of side branch and amorphous components of<br />

the system.<br />

The above result reveals that the octene side branches are excluded from the<br />

crystalline lattice and are accumulating preferentially in the noncrystalline region of<br />

the polyethylene. This result agrees with the view that the depression of the melt<br />

temperature and crystallinity of linear low-density polyethylene is caused by the<br />

inability of the crystalline lattice to incorporate chain branches above a certain<br />

length. It is not possible, however, to determine conclusively from this 2D IR data<br />

if the short branches of linear low-density polyethylene are uniformly distributed<br />

within the amorphous phase domain, or preferentially accumulate near the interface<br />

between the amorphous and crystalline regions.<br />

The high-resolution capability of 2D IR spectroscopy to differentiate highly<br />

overlapped crystalline and amorphous IR bands has been successfully employed<br />

for the characterization of many other semicrystalline polymers. Biodegradable<br />

poly(hydroxyalkanoate)s, for example, were studied by this technique to show that<br />

molecular defects created by the incorporation of comonomer units tend to accumulate<br />

in the amorphous regions [49] in a manner similar to the case of linear lowdensity<br />

polyethylene. Additives, such as plasticizers, which are miscible in the<br />

molten state of semicrystalline polymers also tend to be excluded from the crystal<br />

lattice and preferentially accumulate in the amorphous phase once the polymer is<br />

brought down to a temperature below the melt temperature.<br />

1.5.3 Immiscible <strong>Polymer</strong> Blends<br />

Mixtures of polymers are often immiscible and spontaneously phase separate because<br />

of the much reduced entropy contribution to the free energy of mixing [50].


1.5 Applications<br />

21<br />

/ I<br />

1430<br />

Figure 1-18. A synchronous 7D IR<br />

spectrum in the CH deformation and<br />

aromatic ring semicircle stretching<br />

vibration region of a blend of polystyrene<br />

and low-density polyethylene<br />

at room temperature [2].<br />

I I I I I I I I<br />

1510 1470 14<br />

Wavenumber, V I<br />

c,<br />

><br />

i<br />

n<br />

E<br />

1470 2<br />

a,<br />

2<br />

1510<br />

Such phase-separated polymer mixtures serve as an excellent model system where<br />

molecular-level interactions between components are small. The system studied<br />

here is a film made of an immiscible binary mixture of atactic polystyrene and lowdensity<br />

polyethylene (21. The dynamic IR measurement was carried out as before by<br />

mechanically stimulating the system at room temperature with a 23-Hz dynamic<br />

tensile strain with an amplitude of 0.1% The time-dependent fluctuations of IR<br />

absorbance induced by the strain were recorded at a spectral resolution of 4 cm-* .<br />

Figure 1-18 shows the synchronous 2D IR spectrum of the blend. Autopeaks<br />

observed on the diagonal positions of the spectrum near 1454 and 1495 cm-'<br />

represent the strain-induced local reorientation of polystyrene phenyl rings. The<br />

1454-cm-] band also contains a contribution from CH2 deformation of the backbone<br />

of polystyrene [lo, 11, 511. A pair of intense cross peaks appear at the offdiagonal<br />

positions of the spectral plane near 1454 and 1495 cm-', indicating the<br />

existence of a strong synchronicity between the reorientation of transition dipoles<br />

associated with these two IR bands of polystyrene phenyl side groups.<br />

Similarly, autopeaks corresponding to dynamic IR intensity variations for the<br />

CH, deformations of the polyethylene are observed near 1466 and 1475 cm-'.<br />

These autopeaks arise from the reorientation of molecular chains in the amorphous<br />

and crystalline domains of polyethylene. A pair of cross peaks clearly correlate the<br />

two IR bands originating from the polyethylene component of this blend. It is important<br />

to note that there is little development of synchronous cross peaks correlating<br />

polystyrene bands to polyethylene bands.<br />

The lack of synchronous cross peaks between polystyrene and polyethylene bands<br />

indicates these polymers are reorienting independently of each other. Cross peaks<br />

appearing in the asynchronous spectrum (Figure 1-19) also verify the above conclusion.<br />

For an immiscible blend of polyethylene and polystyrene, where molecularlevel<br />

interactions between the phase-separated components are absent, the timedependent<br />

behavior of IR intensity fluctuations of one component of the sample


Figure 1-19. An asynchronous 2D IR<br />

spectrum of a blend of polystyrene and<br />

1510 1470 1430 low-density polyethylene at room tern-<br />

Wavenurnber, v(<br />

perature [Z].<br />

may become substantially different from those of the other. Therefore, 2D IR correlation<br />

spectroscopy can easily differentiate individual spectral contributions from<br />

each of the immiscible components.<br />

Another notable feature in Figure 1-19 is the development of asynchronous<br />

cross peaks correlating the 1459-cm-’ band attributed to the CHz-deformation<br />

of the polystyrene backbone to the 1454-cm-’ and 1495-cm-’ bands arising from<br />

semicircle-stretching vibrations of the phenyl side groups. The asyncronicity among<br />

IR signals from these polystyrene bands again reflects the difference in mobilities of<br />

backbone and side-group functionalities, as already discussed in Section 1.5.1. The<br />

asynchronicity expected between the crystalline and amorphous bands of the polyethylene<br />

component are not apparent due to the intense cross peaks associated with<br />

polystyrene.<br />

The assignments of the 1454-cm-’ and 1459-cm-’ bands can be independently<br />

verified by selectively substituting the hydrogen atoms of polystyrene backbone<br />

with deuterium atoms. The elimination of spectral contributions from the backbone<br />

of polystyrene leaves only the pure phenyl vibration spectrum in this region, as<br />

shown in Figure 1-20. The asynchronous 2D IR spectrum (Figure 1-19) differentiated<br />

the two IR bands even without deuterium substitution. This result clearly<br />

demonstrates the truly unique feature of 2D IR spectroscopy to enhance substantially<br />

the resolution of highly overlapped spectral regions by effectively spreading<br />

IR bands along the second dimension according to the characteristic reorientational<br />

response times of individual functional groups.<br />

1.5.4 Miscible <strong>Polymer</strong> Blends<br />

Certain classes of polymer blends are known to be truly miscible and form homogeneous,<br />

one-phase mixtures. A specific interaction at the submolecular level. e.g..


1.5 Appliccitioiis 23<br />

Methylene<br />

Figure 1-20. IR spectra of normal atactic polystyrene and<br />

backbone-deuterated polystyrene films.<br />

I I I I<br />

I 475 14: 5<br />

Wavenumber<br />

hydrogen bonding, is often involved in such blends to overcome the tendency<br />

to phase segregate. In those cases, 2D IR spectroscopy is especially suited for the<br />

investigation of the molecular origin of such specific interactions. One of the wellknown<br />

miscible systems, a blend of atactic polystyrene and poly(viny1 methyl ether)<br />

1191. was investigated using 2D IR to elucidate the molecular origin of the miscibility<br />

of these polymers [52. 531.<br />

Interestingly. polystyrene and poly(viny1 methyl ether) are very different polymers.<br />

Polystyrene is a hard hydrophobic plastic resin. while polyivinyl methyl ether)<br />

is a soft water-soluble polymer. It is quite surprising that such a pair of polymers<br />

could produce a molecularly mixed homogeneous one-phase system. The specific<br />

origin of the miscibility of this polymer pair is not well understood, although some<br />

level of interaction is speculated between the methoxyl groups of poly(viny1 methyl<br />

ether) and the phenyl groups of polystyrene [54].<br />

To simplify the spectral assignment, polystyrene which had been selectively<br />

deuterium substituted to eliminate the backbone contribution (Figure 1-21) was<br />

blended with poly(viny1 methyl ether). Special attention was paid to the symmetric<br />

CHZ-stretching peak of the methoxyl group of the poly(viny1 methyl ether) component<br />

near 2820 cm-'. Figure 1-22 shows the asynchronous 2D IR spectrum of<br />

poly(viny1 methyl ether) in this region. The presence of a pair of asynchronous cross<br />

peaks at the spectral coordinates at 2825 and 2813 cm-' indicates there are two<br />

distinct populations of methoxyl groups belonging to the poly(viny1 methyl ether)<br />

component of this blend, each having a different IR absorption band.<br />

Figure 1-23 shows that the methoxyl band of poly(viny1 methyl ether) at<br />

28 13 cm-' is synchronously correlated with the IR bands located at 3055 cm-' and<br />

3024 cm-' (assigned to the phenyl side groups of the polystyrene component of this<br />

blend). The presence of such synchronous correlation cross peaks strongly suggests<br />

the existence of a specific interaction synchronizing the local motions of the<br />

methoxyl groups of polyivinyl methyl ether) with the polystyrene phenyl groups.


3200 2950 2700<br />

Figure 1-21. IR spectra of backbone-deuterated<br />

' ' ' ' ' 2950 I ' ' ' ' 27do polystyrene and poly(viny1 methyl ether) in the<br />

Wavenumber<br />

CH-stretching vibration region.<br />

3200<br />

2825<br />

PVME melhoxyl<br />

L<br />

...'<br />

- Figure 1-22. An asynchronous ZD IR<br />

spectrum for the methoxyl vibration of<br />

a blend of backbone-deuterated poly-<br />

I<br />

2850<br />

, I O I<br />

2820<br />

, I 2850<br />

I<br />

2790 styrene and poly(viny1 methyl ether)<br />

Wavenumber, v, [571.<br />

Such a specific interaction most likely involves the lone-pair of electrons on the<br />

methoxyl oxygen atom of poly(viny1 methyl ether) and n-electrons of the polystyrene<br />

phenyl groups [54].<br />

Figure 1-23 indicates only one methoxyl band is interacting with the polystyrene<br />

phenyl group, even though there are two distinct IR contributions of poly(viny1<br />

methyl ether) methoxyl groups, as already shown in Figure 1-22. The other<br />

methoxyl band of poly(viny1 methyl ether) at 2824 cm-' is asynchronously correlated<br />

with the phenyl bands of the polystyrene component (Figure 1-24). This result<br />

suggests that while one component of polyivinyl methyl ether) methoxyl groups


1.5 Ajqdicrrtions 25<br />

3024<br />

Figure 1-23. A synchronous ZD IR<br />

spectrum of a blend comparing the<br />

reorientational motions of transition<br />

dipoles associated with the polystyrene<br />

phenyl and poly(viny1 methyl ether)<br />

methoxyl groups [57].<br />

.A<br />

t<br />

2850<br />

075 3045 3015<br />

Wavenumber, v,<br />

I<br />

U<br />

/I<br />

I<br />

d3-PS phenyl<br />

I ' 2790<br />

Figure 1-24. An asynchronous 2D IR<br />

spectrum of a blend comparing the<br />

reorientational motions of transition<br />

dipoles associated with the polystyrene<br />

phenyl and poly(viny1 methyl ether)<br />

methoxyl groups 1571.<br />

I<br />

represented by the 2813-cm-' band is strongly interacting with the polystyrene<br />

phenyl groups, the other component represented by the 2824 cm-' band is not.<br />

There are many other examples of miscible polymer blends arising from specific<br />

submolecular interactions of the components. The miscible polymer pair of polystyrene<br />

and poly(pheny1ene oxide). for example, was recently studied by the 2D FT-<br />

IR technique in an effort to understand the molecular origin of the miscibility of<br />

this system [55]. Specific interactions between polymers and various small molecules,<br />

such as additives and plasticizers, were also studied by 2D IR. The importance<br />

of localized submolecular level interactions among functional groups of<br />

mixture components was demonstrated.


1.5.5 Block Copolymers<br />

Block copolymers made of dissimilar polymer segments joined by a covalent bond<br />

are interesting systems to study by 2D IR spectroscopy. Because of the repulsive<br />

interaction between the dissimilar segments, block copolymers often phase separate<br />

at the submolecular scale to form microdomains. It has been speculated for some<br />

time [56, 571 that the composition of block segments comprising such microdomains<br />

does not change sharply at the domain interface. Instead, a diffuse interlayer is<br />

thought to exist where substantial mixing of dissimilar block segments actually takes<br />

place. Analysis based on 2D 1R was used to probe the existence of such an iiitrrphirse<br />

region between microdomains [58, 591.<br />

A thin solution-cast film of a diblock copolymer consisting of a polystyrene block<br />

(a,, 50,000) and a deuterium-substituted polyisoprene block (a,, 100,000 J was<br />

stimulated at room temperature with a 0.1%)-amplitude, 23-Hz dynamic tensile<br />

strain. In this composition range, diblock copolymers of styrene and isoprene produce<br />

a hexagonally packed regular array of rod-like polystyrene microdomains<br />

embedded in a continuous matrix of the polyisoprene block. The diameter of each<br />

polystyrene microdomain is well below 0.1 pm.<br />

Figure 1-25 shows the synchronous 2D IR spectrum of this block copolymer<br />

in the CH-stretching region. Since the polyisoprene block of this polymer is fully<br />

deuterium substituted, the responses shown in Figure 1-25 represent exclusively the<br />

contribution from the polystyrene segment. One may expect the effect of the applied<br />

mechanical perturbation on the molecular reorientation of the block copolymer will<br />

be felt predominantly by the rubbery polyisoprene segments, since the dispersed<br />

polystyrene microdomains are in the glassy state at room temperature. It is, therefore,<br />

rather surprising to find autopeaks indicating substantial molecular mobility<br />

of polystyrene segments under dynamic strain at this temperature.<br />

2800<br />

2975<br />

7 1<br />

I I 1 1 1 ’<br />

i0 2975 2E<br />

Wavenurnber, vI<br />

I<br />

3150<br />

Figure 1-25. A synchronous 2D IR<br />

spectrum in the CH-stretching vibration<br />

region of a microphase-separated diblock<br />

copolymer consisting of polystyrene and<br />

perdeuterated polyisoprene segments.


1.5 Applicutions 27<br />

Furthermore. the sign of the synchronous cross peaks correlating the motion<br />

of side-group phenyl and backbone methylene bands is positive. As discussed in<br />

Section 1.5.1, the positive cross peaks indicate the polystyrene segments giving rise<br />

to these dynamic TR signals are in the rubbery state, even though the measurement<br />

was made at room temperature. This result may be interpreted as the polystyrene<br />

segments showing rubber-like behavior being located in the diffuse interlayer region<br />

between the microdomains. The glass-transition temperature of the polystyrene<br />

segments in this region is substantially depressed due to mixing with the polyisoprene<br />

segments.<br />

This hypothesis was later verified by selective deuterium substitution of part of<br />

the polystyrene segment [58, 591. Near the junction between the two blocks (most<br />

likely to be located at the interphase region), polystyrene segments exhibited rubbery<br />

behavior even at room temperature. This part of the polystyrene segment<br />

reoriented extensively for a given dynamic deformation and contributed strongly<br />

to the dynamic IR dichroism signals. On the other hand, the tail end of the polystyrene<br />

segment (distributed more uniformly within the polystyrene microdomain)<br />

remained predominantly glassy. The local submolecular mobility of the tail end of<br />

polystyrene was noticeably limited.<br />

1.5.6 Biopolymers<br />

2D IR spectroscopy has also been used extensively to study not only synthetic<br />

polymers but also biopolymers. The exceptionally high spectral resolution of 2D IR<br />

has provided a major advantage over conventional IR analyses plagued by heavily<br />

overlapped bands of biopolymers. For example, amide I and I1 regions of IR spectra<br />

of various proteins are thought to be composed of numerous overlapping bands,<br />

each representing different protein conformation, e.g., helix, sheet, and random<br />

coils. However, it is usually very difficult to separate individual spectral contributions.<br />

2D IR analysis of the protein component of human skin has been attempted<br />

in the past [3, 211. Asynchronous 2D IR spectra suggest IR bands assignable to<br />

different protein conformations can be readily differentiated by the development of<br />

asynchronous cross peaks. Synchronous 2D IR spectra, on the other hand, can<br />

correlate different IR bands belonging to similar protein conformations. Thus, 2D<br />

IR spectroscopy may be used systematically to map out the 1R band assignments of<br />

complex protein molecules.<br />

Similar 2D IR studies have been carried out for human hair keratin [44, 60, 61).<br />

Figure 1-26 shows the asynchronous 2D IR correlation spectrum of a protein film<br />

made from solubilized hair keratin [44]. In a normal one-dimensional 1R spectrum,<br />

individual band components are highly overlapped and completely obscured. The<br />

enhanced spectral resolution is quite apparent in the 2D IR spectrum. The position<br />

of many of the asynchronous 2D IR cross peaks are in good agreement with other<br />

band assignments and normal-mode analyses [62].


1'<br />

1656<br />

1661<br />

1669<br />

1679<br />

I I I I I I I I I 1 1788<br />

BE 1650 1608<br />

Wavenumber, u,<br />

Figure 1-26. An asynchronous ZD IR spectrum in the amide I vibration region of a film made of<br />

solubilized human hair keratin 1441.<br />

1.6 Further Extension of 2D Correlation Technique<br />

The basic concept of 2D IR spectroscopy based on the correlation analysis of<br />

perturbation-induced time-dependent fluctuations of IR intensities could be readily<br />

extended to other areas of polymers spectroscopy. The 2D correlation analysis has<br />

been successfully applied to the time-dependent variations of small angle X-ray<br />

scattering intensity measurement [4]. In this study, a small amplitude dynamic<br />

strain is applied a sheet of microphase-separated styrene-butadiene-styrene triblock<br />

copolymer sample. Intensity variation of scattered X-ray beam due to the<br />

strain-induced changes in the interdomain Bragg distances coupled with the reorientation<br />

of microdomain structures is analyzed by using the 2D correlation map.<br />

Similarly, the formal approach of 2D correlation analysis to time-dependent spectral<br />

intensity fluctuations has been extended to UV, Raman [63), and near-IR<br />

spectroscopy (641. There seems no intrinsic limitation to the application of this<br />

versatile technique in polymer spectroscopy.<br />

The introduction of the generalized 2D correlation method [8] has made it possible<br />

to further extend the utility of 2D correlation analysis beyond the study of<br />

simple sinusoidally varying dynamic spectral signals as originally described in Eqs.<br />

(1-2) and (1-3). By utilizing the time-domain Fourier transform of spectral intensity<br />

variations, a formal definition of 2D correlation spectra resembling Eq. (1-9) can be


1.7 Coizclirsions 29<br />

derived in a straightforward manner. Using this generalized method, the 2D correlation<br />

analysis of any time-dependent spectral signals can be carried out. Highly<br />

nonlinear time-dependent reactions of crosslinking polymer systems, for example,<br />

are followed by the time-resolved IR measurement [65, 661. The resulting sets of<br />

time-dependent IR spectra can be transformed into 2D correlation spectra to yield<br />

valuable information to probe details of curing reactions of polymers.<br />

One of the very promising new developments in the recent 2D correlation analysis<br />

of polymers is the application of this versatile technique to the study of spectral<br />

changes recorded as functions of a general physical variable, which is no longer<br />

limited to time. For example, it is possible to apply the 2D correlation analysis<br />

to the near-IR spectral changes corresponding to the development of disorder in<br />

a hydrogen-bonded polyamide system induced by the rising temperature [67]. Of<br />

course, spectral changes of polymers induced by other variables, such as pressure,<br />

age, composition, and the like, may be analyzed in the same way.<br />

1.7 Conclusions<br />

2D IR spectroscopy based on the correlation analysis of the individual timedependent<br />

behavior of localized characteristic reorientational motions of various<br />

submolecular moieties comprising a system has been shown to be very useful for a<br />

broad range of applications in the study of complex polymeric materials. In 2D IR,<br />

the spectral resolution is substantially enhanced by spreading the overlapped IR<br />

bands along the second dimension. The presence or lack of chemical interactions or<br />

connectivity among functional groups located in various parts of the polymer system<br />

are detected. The relative reorientation directions and the order of realignment<br />

sequence of submolecular units are also provided.<br />

Notable discoveries made through the use of 2D IR spectroscopy include the fact<br />

that local dynamics of side groups can proceed independent of the polymer main<br />

chain, and there can be abrupt changes in the side group realignment mechanism<br />

above and below the glass transition temperature. It is possible to probe the niicroscopic<br />

spatial distribution of submolecular components of polymers, especially in<br />

phase-separated systems such as semicrystalline polymers and block copolymers,<br />

and the degree of interactions between various components can be estimated.<br />

Specific interactions between Components in polymer mixtures are quite effectively<br />

identified. The application of 2D IR spectroscopy certainly is not limited to the<br />

characterization of traditional synthetic polymers. Complex macromolecules of<br />

biological origin can also be readily studied by this technique. The recent extension<br />

of the 2D correlation concept to (1) spectroscopic study of polymers using probes<br />

other than IR, (2) nonsinusoidally varying spectral changes, and 13) dependence<br />

on physical variables other than time has greatly expanded the scope of possible<br />

spectroscopic applications of this technique to the study of polymeric materials.


30 I T~.vo-Dii7iensioiiLiI Infmr-ed (2D IR) <strong>Spect</strong>roscopji<br />

1.8 Symbols and Abbreviations<br />

A<br />

All<br />

AL<br />

A 0<br />

p2<br />

D<br />

Dm<br />

DIRLD<br />

T<br />

t<br />

AA<br />

AAW<br />

AA<br />

AA<br />

AA’<br />

AA”<br />

0<br />

X<br />

Y<br />

CI<br />

P<br />

I<br />

&<br />

&<br />

V<br />

8<br />

T<br />

a)<br />

0<br />

Absorbance<br />

Parallel absorbance<br />

Perpendicular absorbance<br />

Structural absorbance =All + 2Al<br />

Orientation factor<br />

Dichroic ratio = Ai,/Al<br />

Ultimate dichroic ratio<br />

Dynamic infrared 1inea.r dichroism<br />

Observation period<br />

Time<br />

Dichroic difference 5 All - A1<br />

Ultimate dichroic difference<br />

Dynamic dichroic difference<br />

Amplitude of dynamic dichroic difference<br />

In-phase dynamic dichroic difference E AA cos P<br />

Quadrature dynamic dichroic difference = AA sin p<br />

Synchronous correlation intensity<br />

Two-dimensional cross-correlation function<br />

Asynchronous correlation intensity<br />

Transition dipole orientation angle<br />

Dynamic dichroism phase angle<br />

Dynamic strain<br />

Dynamic strain amplitude<br />

Wa ven urn ber<br />

<strong>Polymer</strong> chain segment orientation angle<br />

Correlation time<br />

Frequency of dynamic strain / dynamic dichroism<br />

Spatial average<br />

1.9 References<br />

[l] Noda, I. Bd. Am. Plzys. SOC. 1986, 31, 520.<br />

[2] Noda, I. .I. Am. Cliem. SOC. 1989, Ill, 8116.<br />

[3] Noda, I. Appl. <strong>Spect</strong>rox. 1990. 44, 550.<br />

[4] Noda, I. Clierrifracfs-l~fflcromof. Ed 1990, I , 89.<br />

151 Palmer, R. A.; Manning, C. J.; Chao. J. L.; Noda, 1.; Dowrey, A. E.; Marcott, C. Appl.<br />

SiJeCtrO,SC. 1991, 45, 12.<br />

161 Marcott, C.; Dowrey, A. E.; Noda, I. Aridj;t. C/zem. 1994, 66, 1065A.<br />

[7] Marcott, C.; Dowi-ey, A. E.: Noda. 1. A~JJ/. <strong>Spect</strong>rosc. 1993, 47, 1324.<br />

[8] Noda, I. Appl. <strong>Spect</strong>rosc,. 1993, 47, 1329.


1.9 References 31<br />

[9] Colthup, N. B.; Daly, L. H.; Wilberley, S. E. Introduction to Infrared urzd Ramon <strong>Spect</strong>rosc,oipv:<br />

Academic: New York, 1975.<br />

[ 101 Zbinden. R. Itfrored <strong>Spect</strong>roscopy of High <strong>Polymer</strong>s; Academic Press: New York, 1964.<br />

[I 11 Painter, P. C.; Coleman, M. M.: Koenig, J. L. The Theory uf Vibrationnl <strong>Spect</strong>roscopj~ rind Its<br />

AjJjdircitiori to Po/j]rireric Materials: Wiley: New York, 1982.<br />

[ 121 Koenig, 1. L. Sj~ei-tro~co/i.y of Po/yners; American Chemical Society: Washington, D.C., 1992.<br />

[I31 Aue, W. P.: Bartholdi, E.; Ernst, R. R.; J. Cheni. Pl~ys. 1976, 64, 2229.<br />

1141 Nagayama. K.; Kumar. A.; Wiithrich, K.; Ernst, R. R. J. Muyn. Rcson. 1980, 40, 321.<br />

1151 Ernst, R. R.: Bodenhausen, G.; Wokaun, A. Princ@les of Nriclecil- Muyi7rtic Resonunce in One<br />

and Two Dimensions; University Press: Oxford, 1987.<br />

[16] Kessler, H.; Gehrke. M.; Griesinger, C. Angew. Clzem. Int. Ed. Engl. 1988, 37, 490.<br />

[17] Bovey, F. A. Po/j777. Eny. Sci. 1986, 26, 1419.<br />

[18] Bliimich, B.; Spiess. H. W. Angew. Chem. Znt. Ed. Engl. 1988, 27, 1655.<br />

[19] Mirau, P.; Tanaka, H.: Bovey, F. Marromolecules 1988, 21, 2929.<br />

[20] Noda, I.; Dowrey. A. E.: Marcolt, C. J. Polym. Sci., Polym. Lett. Ed. 1983, 31. 99.<br />

[21] Palmer, R. A,; Manning, C. J.; Chao, J. L.; Noda, I.; Dowrey, A. E.: Marcott, C. Appl.<br />

<strong>Spect</strong>rosc. 1991, 45, 12.<br />

[22] Meier, R. J. Macroniol. Syinp. 1997, 119, 25.<br />

[23] Noda, I.; Dowrey, A. E.: Marcott, C. Appl. <strong>Spect</strong>rosc. 1988, 42, 203.<br />

[24] Noda, I.; Dowrey, A. E.; Marcott, C. J. Molec. Struct. 1990, 224, 265.<br />

[25] Noda, I.; Dowrey, A. E.: Marcott, C. Polyrn. Neivs 1993, 18, 167.<br />

[26] Ferry, J. D. I/iscoela.rtic Properties of'Polwiers; 3rd ed.: Wiley: New York, 1980.<br />

[37] McCrum, N. G.; Read, B. E.; Williams. G. Anelastic and Dielectric Effects in <strong>Polymer</strong> Solids;<br />

Wiley: New York, 1967.<br />

[28] Fraser, R. D. B. J. Chi. Plrys. 1953, 31, 1511.<br />

[29] Fraser, R. D. B. J. Chern. Phys. 1958, 28, 1113.<br />

[30] Stein, R. S. J. Appl. Phys. 1961, 32, 1280.<br />

[31] Gregoriou, V. G.; Chao, J. L.; Toriumi, H.; Marcott, C.; Noda, I.; Palmer, R. A. SPIE 1991,<br />

1575, 209.<br />

[32] Gregoriou. V. G.; Chao, J. L.; Toriumi, H.; Palmer, R. A. Chi. Pl7ys. Lett. 1991, 175, 491,<br />

[33] Nakano, T.; Yokoyama, T.; Toriumi, H. Appl. <strong>Spect</strong>rosc. 1993, 47, 1354.<br />

[34] Chazalviel, J.-N.: Dubin, V. M.; Mandal, K. C.; Ozaman, F. Appl. Sj~ertrosr. 1993, 47, 1411.<br />

[35] Budevska, B. 0.; Griffiths, P. R.; Analyt. Chem. 1993, 65, 2963.<br />

[36] Dittmar, R. M.; Chao, J. L.; Palmer, R. A,, in: Photoacoustic and Pliototherni~il Phenomena<br />

III: Bicanic, 0. (Ed.). Springer: Berlin, 1992; pp, 492-496.<br />

[37] Marcott, C.; Story, G. M.; Dowrey, A. E.; Reeder, R. C.; Noda, I., Mikrochirn. Acta [Sicppl.]<br />

1997, 14, 157.<br />

[38] Noda, I.; Dowrey, A. E.: Marcott, C. Mikvochim. Acta [Wien] 1988, I, 101.<br />

[39] Noda, I.; Dowrey, A. E.; Marcott, C.; Polym. Prep. 1992, 33(1), 70.<br />

[40] Noda, I.; Dowrey, A. E.; Marcott, C.; Makromol. Chem., Macromol. Synp 1993, 72, 121.<br />

[41] Noda, I.; Dowrey, A. E.; Marcott, C.; J. Appl. Polyni Sci.: Appl. Polynz. Svnip. 1993, 52, 55.<br />

[42] I. Noda, Polym. Prep Jclpan. Engl. Ed. 1990, 39(2), 1620.<br />

[43] Noda, I.: Dowrey, A. E.: Marcott, C. Po[vm. Pvepr,. 1990, 31(1), 576.<br />

[44] Noda, I.; Dowrey, A. E.; Marcott, C. in Time-Resoloed Vihrational <strong>Spect</strong>roscopy V; Takahashi,<br />

H. Ed.; Springer-Verlag: 1992: pp. 331-334.<br />

[45] Noda, I.; Dowrey. A. E.; Marcott, C. J. Molec. Stl-rict. 1990, 234, 265.<br />

[46] Stein, R. S.; Satkowski, M. M.; Noda, 1. in Polyri7er Blends. Solrifiorrs, and Interfirces, Noda, I.:<br />

D. N. Rubingh, Eds.; Elsevier, New York: 1992; pp. 109-131.<br />

[47] Gregoriou, V. G.; Noda, I.; Dowrey, A. E.; Marcott,C.; Chao, J. L.; Palmer, R. A.; J. Polym.<br />

Sci: Part B: Polym. Phys., 1993, 31, 1769.<br />

[48] Palmer, R. A,: Manning, C. J.; Rzepicla, J. A.; Widder, J. M.; Thomas. P. J.; Chao, J. L.;<br />

Marcott, C.; Noda, I. SPIE 1989, 1145, 277.<br />

[49] Marcott, C.; Noda, I.: Dowrey A. E. Analytica Chim. Acta 1991, 250, 131.<br />

[SO] Flory, P. J. Princip1c.s of <strong>Polymer</strong>. Chemistry; Cornell University Press: Ithnca. 1953.<br />

[51] Snyder, R. W.; Painter. P. C. <strong>Polymer</strong> 1981, 22, 1633.<br />

[52] Noda, I.; Dowrey, A. E.; Marcott, C. Oyo Buturi 1991, 60, 1039.


32 1 T~iio-DinieiisioiiiIInfrared (20 IR) <strong>Spect</strong>roscopy<br />

[53] Satkowski, M. M.; Grothaus, J. T.; Smith, S. D.; Ashraf, A,; Marcott, C.; Dowrey, A. E.;<br />

Noda, I. in <strong>Polymer</strong> Sohitions, Blends. and Interfaces, Noda, I.; Rubingh, D. N., Eds.; Elsevier:<br />

Amsterdam, 1992; pp. 89-108.<br />

[54] Garcia, D. J. Polynz. Sci., Polym. P11y.r. Ed 1984, 22, 107.<br />

[SS] Palmer, R. A.; Gregoriou, V. G.; Chao, J. L. Po/ym. Prep.. 1992, 33111, 1222.<br />

[56] Learry, D. F.: Williams, M. C.: J. Polvin. Sci.. Po/p. Phys. Ed 1974, 12, 265.<br />

1571 Hashimoto, T.; Fujimura, M.; Kawai, H. M~~cronio/rcn/es 1980, 13, 1660.<br />

[58] Noda, I.; Smith, S. D.; Dowrey, A. E.; Grothaus, A. E.; Marcott, C. Mur. Res. Soc. Sjv7p.<br />

Proc. 1990, 171, 117.<br />

[59] Smith, S. D.; Noda, I.; Marcott, C.; Dowrey, A. E. in <strong>Polymer</strong> Solutions, Blrnd~s, and hiterfrees,<br />

Noda, I.; Rubingh! D. N., Eds.; Elsevier: Amsterdam, 1992; pp. 43-64.<br />

[60] Dowrey, A. E.; Hillebrand, G. G.; Noda, I. Marcott, C. SPIE 1989, 1145, 156.<br />

[61] Noda, I.; Dowrey, A. E.; Marcott, C. Polym Prep. 1991, 32(3). 671.<br />

[62] Byler, D. M.; Susi, H. Biopolynzen 1986, 25, 469.<br />

[63] Ebihara, K.; Takahashi, H.; Noda, I. AppL <strong>Spect</strong>rosc. 1993, 47, 1343.<br />

[64] Noda, I.; Story, G. M.; Dowrey, A. E.; Reeder, R. C.; Marcott, C. Mncroniol. Symp. 1997,<br />

119, 1.<br />

[65] Nakano, T.: Shimada, S.; Saitoh, R.; Noda, I. Appl. Spccrrosc. 1993, 47, 1337.<br />

[66] Shoda, K.; Katagiri, G.; Ishida, H.; Noda, I. SPIE 1994, 2089, 254.<br />

[67] Ozaki, Y.; Liu, Y.; Noda, I. Macroiizolec~iles 1997, 30, 2391.


2 Segmental Mobility of Liquid Crystals and<br />

Liquid-Crystalline <strong>Polymer</strong>s under External<br />

Fields: Characterization by Fourier-tr ansform<br />

Infrared Polarization <strong>Spect</strong>roscopy<br />

H. W Siesler, I. Zebger, Ch. Kuliniia. S. Okretic, S. Shilov and U. Hoffnuinn<br />

2.1 Introduction<br />

Liquid crystals and maia-cliain/side-chain liquid-crystalline polymers have gained<br />

scientific and technical importance due to their applications as display materials,<br />

their exceptional mechanical and thermal properties, and their prospective applications<br />

for optical information storage and nonlinear optics, respectively [ 1-13].<br />

Among other aspects, the study of segmental mobility as a function of an external<br />

(electric, electromagnetic, or mechanical) perturbation is of basic interest for the<br />

understanding of the dynamics of molecular processes involved in technical applications<br />

of liquid-crystalline materials.<br />

Fourier-transform infrared (FTIR) polarization spectroscopy has proved an<br />

extremely powerful technique to characterize the segmental mobility of specific<br />

functionalities of a polymer or a liquid-crystal(1ine polymer) under the influence of<br />

an external perturbation [ 14-23]. In a side-chain liquid-crystalline polymer<br />

(SCLCP) aligned due to the influence of an external field, for example, the orientation<br />

of the polymer backbone relative to the rigid mesogeii is strongly influenced by<br />

the coupling via a more or less flexible spacer (Figure 2-1). Thus, the availability<br />

of characteristic absorption bands for these individual functionalities will allow<br />

a quantitative characterization of their relative alignment and contribute towards<br />

a better understanding of their structure-property correlation. Based on this<br />

knowledge, tailor-made systems can then be synthesized by a systematic variation<br />

of partial structures.<br />

In what follows, different phenomena will be discussed for selected classes of<br />

liquid-crystalline compounds:<br />

structure-dependent alignment of side-chain liquid-crystalline polyacrylates on<br />

anisotropic surfaces<br />

electric field-induced orientation and relaxation of a liquid crystal, a side-chain<br />

liquid-crystalline polymer and a liquid-crystalline guest-host system<br />

reorientation dynamics of a ferroelectric side-chain liquid-crystalline polymer in<br />

a polarity-switched electric field<br />

alignment of side-chain liquid-crystalline polyesters under laser irradiation<br />

shear-induced orientation in chrial smectic liquid-crystalline copolysiloxanes


igid inesugen<br />

mail1 chain<br />

Figure 2-1. Schematic structure of a side-chain liquid-crystalline<br />

polymer.<br />

. orientation mechanism of a liquid single crystal elastomer during cyclic deformation<br />

and recovery.<br />

Static and time-resolved spectroscopic data acquired both, by the conventional<br />

rapid-scan and by the novel step-scan technique will be presented and interpreted in<br />

terms of their individual applications.<br />

2.2 Measurement Techniques and Instrumentation<br />

Since the mid-l970s, FTIR spectroscopy has been widely used for polymer analytical<br />

applications and for studies of time-dependent phenomena in chemical and<br />

physical processes [ 141. With the current iinproveinent in electronics and data<br />

acquisition procedures, the time resolution for such investigations with an acceptable<br />

signal/noise-ratio is presently about 20 111s. Time resolutions far beyond this<br />

value are not accessible by the conventional rapid-scan technique with continuous<br />

mirror moveinelit without additional interferogram data manipulation. Thus, for<br />

time-resolved, rapid-scan measurements, generally, the retardation time of the<br />

moving mirror has to be considerably shorter than the time duration of the investigated<br />

event (Figure 2-2a). In search of new approaches, stroboscopic techniques<br />

have been proposed, where interferograms taken in the rapid-scan mode are subsequently<br />

subjected to a sampling of reordered interferograin segments and have<br />

thus led to an improvement in time resolution by a factor of about 500 [24, 251.<br />

A further improvement in time resolution to the submicrosecond level has been<br />

achieved with the step-scan technique 116, 26, 271. Here, the mirror is moved stepwise<br />

and the optical retardation is held constant during sampling of the individual<br />

interferogram elements. Thus, the time resolution is only limited by the response of


SRMPLE/DETECTOR<br />

e3<br />

‘retardation<br />


2.3 Theory<br />

Throughout this chapter the general form of the order parameter S will be used<br />

[3, 3 1 j to describe the long-range orientational order of the investigated liquidcrystalline<br />

materials. In contrast to the theory of Maiei- and Saupe 132, 331, which<br />

considers the uniaxial distribution of the mesogens in a nematic phase relative to the<br />

macroscopic preferential direction (director), this formalism also allows negative<br />

values. The relation of S to the inclination angle 0 of the inesogenic/structural-unit<br />

axis and the reference direction (if not otherwise stated, the external field direction)<br />

is outlined for the case of an SCLCP in Figure 2-3. Also shown are the values of<br />

S for parallel and perpendicular orientation of the mesogens and their isotropic<br />

distribution. The derivation of S, which corresponds to the orientation factor f in<br />

general polynier orientation terminology [14, 17-20], is given in Figure 2-3, where<br />

R is the dichroic ratio determined from the experimental absorbance values measured<br />

with light polarized parallel and perpendicular, respectively, to the chosen<br />

reference direction. Ro is the theoretical dichroic ratio for perfect aligiiinent of the<br />

mesogen, and p is the angle between the transition moment of the absorbing group<br />

and the local chain axis (in the present case the long axis of the mesogenic group)<br />

[14j. The structural absorbance Ao has been chosen as parameter which represents<br />

the intensity of an absorption band exclusive of contributions from the orientation<br />

of the investigated polymer. It is given by [14]:<br />

Both, order parameter S and structural absorbance A0 are based on the assumption<br />

of a uniaxial orientation syniinetry [14, 18-20].<br />

external field<br />

(referencc direction)<br />

s= I<br />

(e =or)<br />

parallel orientation<br />

a<br />

1<br />

S=-- (0=900)<br />

2<br />

perpendicular orientation<br />

s=o<br />

isotropic distribution<br />

Figure 2-3. Geometrical considerations for the<br />

derivation of the order parameter S of a side-chain<br />

liquid-crystalline polymer under the influence of an<br />

exterilal field (ff = inclination angle of the inesogen<br />

with the reference axis, v, = angle between the<br />

transition moment of the absorbing group and the<br />

local chain axis, Rn = theoretical dichroic ratio for<br />

a perfectly aligned structural unit).


2.4 Srvuc me-Dependenr Alignment 37<br />

In the case of the laser irradiation experiments (see Section 2.6), the polarization<br />

direction of the incident laser beam was chosen as reference direction. Contrary to<br />

the former publications we have changed the reference direction for the calculation<br />

of the photo-induced anisotropy due to a better understanding of the underlying<br />

orientation mechanism. We assume that the mesogens are aligned preferentially<br />

perpendicular to the polrivizniion iiectov [34, 351 and not to a polnvizntion plane [15,<br />

361 of the irradiating laser. Thus, we obtain negative S-values, whereas for the lastmentioned<br />

case the order parameter values would be positive, according to the<br />

Maier-Saupe theory.<br />

2.4 Structure-Dependent Alignment of Side-Chain<br />

Liquid-Crystalline Polyacrylates on Anisotropic<br />

Surfaces<br />

For the characterization of the anisotropic properties of liquid-crystalline (LC)<br />

materials it is necessary to induce a well-defined orientation of the mesogenic<br />

groups. Although spontaneous orientation phenomena can be regularly observed in<br />

limited domains, the necessary prerequisite for the application of LC-materials as<br />

optical construction elements, for example, is the preparation of uniform textures<br />

over larger areas. Such structures can be prepared by the alignment of LC-systems<br />

on anisotropic surfaces (e.g., glass, polymer films). The origin and theory of the<br />

interaction and the induced anisotropy has been treated in several books and<br />

articles [37-441. Here, the attention is primarily directed towards the influence of<br />

the spacer length in a homologous series of nematic side-chain liquid-crystalline<br />

polyacrylates on their alignment on an anisotropic polyimide surface.<br />

2.4.1 Materials and Experimental<br />

The polymers had a molecular weight M, (determined by gel permeation chromatography)<br />

of about 5000 [45] and their structural formula is shown alongside their<br />

nematic-isotropic transition temperatures (determined by polarization microscopy)<br />

[45] in Figure 2-4.<br />

For the preparation of the substrate surface, KBr-plates have been spin-coated<br />

with a thin layer of polyiinide (polyimide-kit ZLI 2650, Merck, Darinstadt, Germany).<br />

Upon drying and curing this coating to the final thickness of about 1 pm, the<br />

surface anisotropy was induced by unidirectional rubbing with a polymethacrylimide<br />

hardfoam-roll. The LC-polymers were sandwiched between the SUrfdCetreated<br />

KBr-plates above their clearing temperatures and finally annealed 3 "C<br />

below these temperatures for several hours. FTIR polarization spectra with a resolution<br />

of 4 cm-' were recorded by accumulating 20 scans.


main chain<br />

rt-<br />

YH2<br />

spacer rnesogeri<br />

+A 0<br />

H-C-C-O-CH2 ) -O-O-$-O-O-CN<br />

95 Figure 2-4. Structure and transition temperatures<br />

of the investigated side-chain liquid-crystalline<br />

polyacrylates.<br />

2.4.2 Results<br />

Despite identical sample pretreatment, gradually decreasing dichroic effects could<br />

be observed for the LC-polymers with decreasing spacer length. The FTIR polarization<br />

spectra of tlie P6CN- and P2CN-systen1, measured with radiation polarized<br />

parallel and perpendicular to the rubbing direction of the polyimide surface,<br />

respectively, are shown in Figure 2-5. From these data the coupling effect of the<br />

spacer between the main chain and the mesogen is clearly evident. Thus, the hesamethylene-spacer<br />

almost completely decouples the mesogen from the main chain<br />

and allows the induction of an extremely high orientation of the rigid mesogen<br />

(S(v(C3N)) = 0.62) (see also insert in Figure 2-5). The short diniethylene-spacer,<br />

on the other hand, strongly links the mesogen to the polymer backbone and no<br />

orientation of the mesogen on the anisotropic surface can be observed. The spectra<br />

of the system with the tetramethylene-spacer (not shown here) display intermediate<br />

effects (S(i(C-N) = 0.46). A detailed analysis of the dichroic effects in the polarization<br />

spectra of the whole series in terms of the relative orientation of the mesogen,<br />

spacer and main chain has been given in [36].<br />

2.5 Electric Field-Induced Orientation and Relaxation<br />

of Liquid-Crystalline Systems<br />

The reorientation of liquid crystals in an external electric field is the basis of their<br />

technical application in liquid-crystal displays (LCDs) and has been treated in<br />

numerous books and reviews [I-3, 12, 38,46-501. This section is intended to denionstrate<br />

tlie application of time-resolved FTIR spectroscopy for a better understanding<br />

of tlie electro-optical performance of liquid-crystalline systems. Due to tlie<br />

reversible character of the orientational and relaxational motions of liquid crystals<br />

under tlie perturbation of an electric field, the time-resolution of the FTIR measurements<br />

can be extended down to the microsecond range by the implementation<br />

of tlie step-scan technique (see Section 2.2). Thur, an insight on ;I molecular level


2.5 Electric Field-Innductid Orirntcitioii 39<br />

f 1<br />

P6CN<br />

0<br />

I<br />

Figure 2-5. FTIR polarization spectra of the P6CN- and P2CN-system ithe insert schematically<br />

shows the aligniiieiit of the mesogens on the anisotropic surface; parallel and perpendicular refers to<br />

the polarization of the IR radiation relative to the rubbing direction).<br />

can be provided into the relevant mechanisms. In what follows. the application of<br />

vibrational spectroscopy to the investigation of different phenomena will be trcated<br />

separately. On the one hand we will discuss the electric field-induced changes of<br />

nematic LC-systems (low-molecular weight compounds, side-chain licluici-crystallirie


40 2 Segnwital Mobility of Liquid Crystuls aid Liquidcrystalline Polymevs<br />

6CPe<br />

0<br />

C H 3 - ( C H z ) e C - O D C = N<br />

5<br />

44.5°C 48.5OC<br />

K - N - I<br />

NLCP<br />

- 0<br />

2, H -&-O-(CHZ)~$$C O a C - N<br />

-.<br />

CH2<br />

-n 3S°C 130°C<br />

G - N - I<br />

Figure 2-6. Sti ucture aiid transition temperatures<br />

of the investigated low- aiid highmolecular<br />

weight LC systems I6CPB and<br />

NLCP).<br />

polymers and liquid-crystalline guest-host system) in switch-on/switch-off experiments,<br />

and on the other hand we will describe the structural consequences incurred<br />

by a ferroelectric side-chain polymer in a poled electric field.<br />

Nematic liquid-crystalline molecules, which consist of a rigid inesogenic unit with<br />

an attached flexible part (sequence of methylene units, for example) (Figure 2-6),<br />

are among the most characteristic thermotropic systems. FTIR spectroscopy has<br />

been applied to study the mechanisiii of orientation of the liquid crystals in an<br />

electric field; however, at present the experimental data are not always unequivocal,<br />

and conclusions obtained are not in complete agreement with each other. T~us,<br />

some authors did not detect any differences in orientational rates of the rigid and<br />

the flexible part of the LC molecule [16, 51-54], which is in accordance with the<br />

commonly accepted mechanism of the orientation of an LC as a cooperative<br />

motion of a molecular ensemble [l, 2, 381. Other authors reported, that the motion<br />

of the flexible part precedes that of the inesogen in both orientation and relaxation<br />

[29, 55-57], and some investigators suggested that the motion of the mesogen precedes<br />

the motion of the flexible part [58, 591. With respect to a more detailed exploitation<br />

of the spectroscopic data in terms of the synchronization of the segmental<br />

motions of different functionalities of the investigated liquid-crystalline systems,<br />

2D-correlation analysis [ 16, 30, 60, 611 has proved an extremely valuable tool. In<br />

view of the detailed discussion of this mathematical technique in Chapter 1 of this<br />

book; however, this topic will not be treated here, although several results extracted<br />

from the presented spectroscopic data are based on the application of the 2Dcorrelation<br />

formalism.<br />

2.5.1 Materials<br />

To study the addressed phenomena, a commercial sample of p-cyanophenyl p-nhexylbenzoate<br />

(hereafter abbreviated as 6CPB) (Figure 2-6) was obtained from<br />

Roche (Basel, Switzerland), and used without further purification [53, 541. The<br />

transition temperatures of this LC are also presented in Figure 2-6. It should be<br />

mentioned, that the nematic phase will be retained during cooling down to 7 "C<br />

below the crystal-nematic transition point. To demonstrate the reduction of


2.5 Electric Field-nclticetl Orientntioi? 41<br />

mobility in an external electric field when the mesogen is attached to a polymeric<br />

backbone, the nematic liquid-crystalline side-chain polymer (NLCP), whose sidechain<br />

is largely identical to the structure of the 6CPB (Figure 2-6), was investigated<br />

under analogous experimental conditions.<br />

2.5.2 Experimental<br />

The schematic construction of the measurement cell and the principles of operation<br />

are outlined in Figure 2-7. For the construction of this cell, two Ge-plates (approximately<br />

10 x 10 x 2 mm3) are used as IR-transparent window material and electrodes.<br />

The distance between the two Ge-plates is ensured by thin stripes of<br />

poly(ethyleneterephtha1ate) film (7 pm) acting as spacer. To avoid the occurrence of<br />

interference fringes in the spectra, the precise parallel aligninelit of the Ge-windows<br />

was intentionally disturbed by two supplementary spacers of polycarbonate film (2<br />

pnij on one side of this assembly (Figure 2-7a). To fill the assembled cell, the LCsample<br />

was heated above the nematic-isotropic transition temperature and then<br />

introduced by capillary forces. For the preparation of a prealigiied nionodoniain-<br />

LC, the inner surface of each Ge-plate was covered with a thin layer of polyimide<br />

and then rubbed in one direction with polyamide cloth (see also Section 2.4.1). This<br />

pretreatment induced a homogeneous orientation of the investigated LC-molecules<br />

or mesogenic functionalities (in the case of the polymer) along the rubbing direction<br />

of the surface (Figure 2-7b). Upon application of an electrical voltage across the<br />

Ge-plates, the liquid crystals rotate with the long axes of their niesogens into the<br />

direction of the applied field (Figure 2-7c). Switching off the voltage leads to a<br />

relaxational motion back to the original state. The accompanying intensity changes<br />

were monitored with IR-radiation polarized parallel to the rubbing direction of the<br />

Ge-windows (Figure 2-7b). A home-built heating cell has been used to keep the<br />

investigated sample at a constant temperature within -1- 0.1 "C during the experiment<br />

(Figure 3-8).<br />

The principle of the electronic set-up for the step-scan measurements, and the<br />

method of data acquisition relative to the experimental time-scale, are shown in<br />

Figures 2.9 and 2.10, respectively. When the scanning mirror has stepped to a definite<br />

retardation point, a control electronic unit coupled with a sine-wave generator<br />

sends a 6 Vpp (peak-to-peak height) pulse with a frequency of 10 kHz and a definite<br />

length (25 nis for the experiments presented here) to the sample cell to induce the<br />

reorientation of the LC. After this period the voltage is switched off during 225 ms<br />

to allow relaxation to the original state. Simultaneously to the application of the<br />

electric field, the same electronic unit sends a trigger pulse to the spectrometer to<br />

start the data collection. The data are sampled by the internal analog-to-digital<br />

converter (ADC) of the instrument during 50 ms with 0.05 ms time resolution.<br />

During this time interval 1000 data points are acquired for one fixed retardation<br />

(the applied data-acquisition scheme is imposed by the capacity of the computer).<br />

This cycle is repeated several times (five times for the presented results on 6CPB)<br />

and the data points corresponding to the same time in the period of the event are<br />

averaged to improve the signal-to-noise ratio. When the data collection at a given


42 2 Segnien frrl Mobility of Liquid Cyystrils crrzd Liqziirl-Cr~~stallii~e Polyrnel-s<br />

GERMAHIUM WINDOU (ELECTRODE)<br />

I \<br />

A<br />

a<br />

I \<br />

I N I<br />

bIWFRARED BEAH<br />

DIRECTION<br />

'<br />

,-<br />

b<br />

POLARIZED INFRARED BEAM<br />

A<br />

C<br />

h<br />

I NFRRRED' BERM<br />

AND ELECTRIC FIELD DIRECTION<br />

ED<br />

Figure 2-7. Electric field-induced orientation<br />

of a liquid crystal (a) Construction<br />

of the meaauiement cell (b) Hornogeiieous<br />

al~gninent of the LC between the<br />

dnisotromc electrode surfaces before<br />

application of the electric field.<br />

(c) Hoineotropic alignment of the LC<br />

during application of the electric field.<br />

retardation is finished, the mirror is moved to the next retardation point (Figure<br />

2-2b).<br />

After a settling time of 40 ms, the experiment is repeated as described above. For<br />

the presented experiments all data were collected symmetrically around the centerburst<br />

of the interferogram. A total number of 2368 mirror positions were scanned,<br />

leading to a spectral resolution of 8 c1n-l. After completing the data collection at<br />

each retardation point, the data were resorted to interferograiiis on the time scale<br />

(Figure 2-2b) and transformed to the corresponding spectra. According to this


a<br />

-@<br />

COWER<br />

SRHPLE<br />

._<br />

2.5 Elec tric Fielrl-hiclirc.td Orieri tci tiori 43<br />

CORXIRL HERIIHG UIRE<br />

INPRRRED BERM<br />

STEEL CORE UITH SRMPLE COHPRRTMENI<br />

CELL /<br />

TEFLOH IHSULRTION<br />

'i<br />

GERMANIUM<br />

I ,<br />

INFRARED BEAM<br />

b -. , ,,__ TEFLON JACKET<br />

Figure 2-8. Variable-temperature cell utilized for the spectroscopic characterization of the electric<br />

field-induced alignment of liquid crystals. (a) Complete assembly of the accessory. (b) Expanded<br />

ciew of the sample cell.<br />

procedure, 1000 spectra with a time resolution of 0.05 ms were processed. A further<br />

improvement of the signal-to-noise-ratio was obtained by averaging 10 spectra<br />

recorded in 0.5-nis intervals, starting with the time corresponding to the onset of<br />

applying the electric field. Thus, finally 100 spectra with a time resolution of 0.5 ms<br />

were obtained. The data collection for the whole experiment required approximately<br />

50 minutes.<br />

Due to the increase of viscosity in the polymeric NLCP system, the reorientation<br />

times varied between minutes and seconds, depending on the experimental temperature.<br />

Thus, the conventional rapid-scan technique (Figure 2-2a) was applied to<br />

follow the orientation and relaxation dynamics of the NLCP. In order to synchronize<br />

the application of the electric field with the data collection. a common switch<br />

was used to control the voltage [53].


44 2 Srgnientcrl Mobility of Liquid Crystcr1.Y mid Lipid-Cryctallirzc <strong>Polymer</strong>.\<br />

TRIGGER PULSE<br />

SIHEWRUE<br />

GENERATOR<br />

SPEC TROMETER<br />

ELECTRONIC CONTROL ~ - - - l f l i , b<br />

___ SAMPLE CELL<br />

/I<br />

I<br />

Figure 2-9. Block diagram of the electronic measurement set-up.<br />

ORIENTATION<br />

PERIODICRL<br />

EVENT<br />

RELAXRTION<br />

TRIGGER<br />

PULSE (TTL)<br />

E 25 50 258 TIME/ms<br />

Figure 2-10. Time scale of the periodic event, control signal and data collection sequence for the<br />

step-scan measurements.<br />

2.5.3 Results and Discussion<br />

2.5.3.1 p-Cyanophenyl p-n-Hexylbenzoate (6CPB) in a Switching<br />

Experiment<br />

The FTIR polarization spectra of a prealigned 6CPB in the nematic phase are presented<br />

in Figure 2-11. The band assignment of the most prominent absorptions<br />

based on the spectral analysis of similar compounds [59, 551 is given in Table 2-1.<br />

The v(C-N) band (2230 cin-') and the two v(C=C)~~ stretching bands (1605 and<br />

1503 cm-') are polarized along the long axis of the mesogenic group and show<br />

parallel dichroisni with All > A1 (Ail and Al are the absorbance values measured


2.5 Electric Field-Itzduced Orientation 45<br />

Figure 2-11. FTIR polarization<br />

spectra of 6CPB (41 "C) taken<br />

with radiation polarized parallel<br />

(upper spectrum) and perpendicular<br />

(lower spectrum) to the<br />

rubbing direction of the polyimide<br />

surface.<br />

3300 3000 2750 2500 2250 2000 1750 1500 1250 1003 600<br />

Wavenuiber cm-'<br />

Table 2-1. Band assignment and waveiiuiiiber position of selected absorption bands for 6CPB (as =<br />

antisymmetric; s = symmetric; ar = aromatic; oop = out-of-plane).<br />

Band assignment Wavenumber (cm- ' )<br />

2929<br />

2558<br />

2230<br />

1742<br />

1605<br />

1503<br />

1415<br />

1464<br />

1264<br />

1207- 101 7<br />

843<br />

757<br />

with light polarized parallel and perpendicular to the rubbing direction of the polyimide<br />

layer, respectively). This means, that the long axis of the mesogenic unit is<br />

oriented predominantly parallel to the rubbing direction (Figure 2-7b). The 757<br />

cni-' band (ring CH out-of-plane defoimation) exhibits a perpendicular dichroism<br />

with Al > A,,, since the transition moment for this band is perpendicular to the<br />

ring plane and hence, perpendicular to the long axis of the mesogen. These four<br />

bands are characteristic of the mesogenic units. For the characterization of the<br />

flexible part of the LC-molecule the 2929 cm-' and 2858 c1n-l bands (antisymmetric<br />

and symmetric v( CH2)-stretching vibrations) were evaluated. All the selected<br />

bands had absorbance values < 1 for the given cell thickness.<br />

When a voltage higher than the threshold value is applied to the electrodes, the<br />

LC molecules undergo a transition from the homogeneous (parallel to the rubbing<br />

direction) to the honieotropic (parallel to the electric field) orientation. According<br />

to our data for 6CPB, the threshold voltage to induce such a reorientation is


Figure 2-12. Stack-plot of spectra measured during the electric field-induced orirnration and duriiip<br />

part of the relaxation of thc 6CPB molecules (41 -C'.<br />

Figure 2-13. Relative absorbance/time-plot during orientation and relaxation of 6CPB (41 C!<br />

for selected absorption bands: (0) 2929 cni-'. (7) 2858 cm-', (012230 cin-l. I + ) 1605 cm-'. I x<br />

1503 cm-'. (+) 757 cni.~'.<br />

approximately 1.5 Vpp. In the homeotropic type of orientation, the long axis of the<br />

LC molecule is normal to the electrode surface and parallel to the IR beam propagation.<br />

According to geometrical considerations [62] a reorientation to this honieotropic<br />

state thus leads to intensity changes of the Ail-component only. This component<br />

asymptotically approaches the Al intensity value at sufficiently high fields.<br />

For this reason, all time-resolved measurements have been performed with light<br />

polarized parallel to the polyimide-layer rubbing direction.<br />

A stack-plot of time-resolved spectra recorded during the honiogeneous-homeotropic<br />

transition and part of the relaxation process is presented in Figure 2-12. A<br />

good signal-to-noise-ratio is observed for all spectral regions (except when the<br />

absorbance exceeds a value of 2). Relative absorbance changes A, for selected<br />

absorption bands were calculated from:<br />

where A, is the absorbance at time t and A0 is the absorbance just before reorientation.<br />

These relative intensity changes of selected absorption bands are presented<br />

in Figure 2-13 and remarkable changes are observed only after a definite induction


2.5 Electric Field-IwrElrced Orientation 47<br />

period of several milliseconds after application of the electric field. This phenomenon<br />

is well described in the literature [46]. The intensities of the 2230, 1605 and<br />

1503 cni-' bands decrease, whereas that foi- the 757 cm-' band increases M,hen the<br />

field is switched on. This indicates that the mesogens 1-eorient with their long axes in<br />

the direction of the applied field (Figure 3-7c). The process of orientation at the<br />

given experimental conditions takes about 25 ms. After switching off the field, thz<br />

relaxation to the initial homogeneous state (Figure 2-7b) takes place. The intensity<br />

changes of the 2929 and 2858 cnir' bands reflect the orientation and relaxation<br />

process of the flexible part of the LC molecule. For an exact quantitative comparison<br />

of the orientational and relaxational rates of the rigid and flexible segments of<br />

the LC molecule, the order parameters of the different bands had to be calculated<br />

as a function of time. However, this is possible only for the case of a pure homogeneous<br />

orientation (it., the director of the LC monodomain and, hence) the axis of<br />

uniaxial orientation is perpendicular to the IR beam propagation) or for the case of<br />

a pure homeotropic orientation (where the axis of uniaxial orientation is parallel<br />

to the beam propagation). In intermediate states the orientation of the director is<br />

not constant. Thus, near the surface, due to strong interactions, the molecules are<br />

oriented perpendicular to the electric field whereas at the center of the cell they are<br />

oriented parallel to the electric-field direction [38]. For a qualitative compai-ison<br />

of the orientational rates of the rigid and the flexible parts of the investigated LC<br />

molecule a normalized intensity A,, was utilized which has been calculated from:<br />

where At is the band peak absorbance at time t, A0 is the value before the npplication<br />

of the electric field, and A,,, is the corresponding absorbance value at t = 25<br />

ms. The result for the 2929 and 2230 cm-* bands are presented for a temperature of<br />

41 "C in Figure 2-14a. According to these data there are no detectable differences in<br />

A, both, during the orientation and relaxation process, for the rigid and the flexible<br />

segments of the LC molecule. This means that under the given experimental conditions<br />

and with the evaluatioii by the usual one-dimensional technique, the different<br />

segments of the LC molecule orient and relax at comparable rates. The same<br />

conclusion can be derived from Figure 2-14b where the corresponding A,,-values<br />

are presented for 45°C. Due to the small temperature interval, the diRerence in<br />

orientational behavior is not significant. However, it is not possible to perform<br />

the experiment in a wider temperature range because of the narrow temperature<br />

interval of the investigated mesophase.<br />

The results of similar rates for the reorientation of the flexible and rigid segments<br />

of the 6CPB molecule in an electric field are in accordance with the coninioiily<br />

accepted mechaiiisni of orientation by a cooperative motion of the LC molecules [ l~<br />

2, 381. However, in recent investigations it has been reported, that the response of<br />

the mesogens and the flexible part on the applied field may be different [29, 55-59].<br />

If these observations are correct they could be the basis of a principally new<br />

description of LC orientation on a molecular level. Two explanations for thc Jisagreement<br />

of o~iresults with these views can be given:


48 2 Segliientril Mobility qf Liquid Crystals arid Liquid-Cr.~~stiillirie <strong>Polymer</strong>s<br />

ORIENTATION c+<br />

RELAXATION<br />

----------<br />

I0 20 30 40 5n<br />

TIME lms<br />

n 10 213 310 40 50<br />

TlMElms<br />

Figure 2-14. Normalized absorbaiice/time-plot during orientation and relaxation of 6CPB at two<br />

different temperatures for selected absoiptioii bands: (0) 2929 cm-', (0) 2230 cin-'.<br />

1. The effect of different response of the flexible and the rigid part of the LC<br />

molecule towards application of an electric field is not general. This is supported<br />

by results [56], where this phenomenon is detected under certain experimental<br />

conditions only. Moreover, up to now most experiments have been performed<br />

on one LC (4-n-pentyl-4'-cyanobiphenyl, 5CB) only, which has a different<br />

structure compared to the LC molecule investigated here.<br />

2. The effect is too small to be detected under the given experimental conditions<br />

with conventional data analysis. To detect this effect, Urano et al. [55, 561 have<br />

applied tiine-resolved spectroscopy by using a modified dispersive instrument.<br />

With this instrument, absorbance changes down to lop6 a.u. have been detected.<br />

However, the time requirement for the experiment in the step-scan inode<br />

depends on the period of the event (0.25 s in our case). To reach such a high<br />

signal-to-noise-ratio we had to average the data of many more events at the<br />

individual retardation points. For example, the data collection for 2368 retardation<br />

points over 100 events in each point (compared with our five events) would<br />

require approximately 17 h of experimental time. Due to uncontrolled changes<br />

in the environmental conditions and drifts of electronics during such a time<br />

period, the main requirement of the step-scan technique, i.e., the reproducibility<br />

of the periodical event, could be violated.<br />

To extract more information, the data were processed with the newly developed<br />

method of two-dimensional spectroscopy [60, 6 11. This technique is especially useful<br />

for the detection of small differences in the orientational rates of different molecular


2.5 Electric Field-Im/irced Orientutioii 49<br />

I<br />

- 1 ' I I I I<br />

32OU 3000 2750 2500 2250 2000 1750 1500 1250 I000 750 500<br />

WAVENUMBER / cm'<br />

Figure 2-15. FTIR polarization spectra of NLCP (25 "C) taken with radiation polarized parallel<br />

and perpendicular to the rubbing direction.<br />

groups. With the development of this formalism it is now possible to construct 2D<br />

correlation plots based on conventional time-resolved spectral data. The 3D analysis<br />

reveals [ 161, that there is a strong correlation in the synchronous spectrum and<br />

no appreciable asynchronicity in the asynchronous spectrum between the niesogen<br />

bands and the absorptions of the hydrocarbon chain (see Table 2-1). This means,<br />

that the mesogen and the flexible segment of the liquid crystal reorient synchronously.<br />

The strong correlation between the inesogen bands themselves indicates<br />

that the mesogen reorients as a rigid core. No differences in 2D correlation spectra<br />

between the orientation and relaxation processes were found.<br />

2.5.3.2 Nematic Side-Chain Liquid-Crystalline <strong>Polymer</strong> in a<br />

Switching Experiment<br />

The FTIR polarization spectra of the prealigned NLCP are presented in Figure<br />

2-15. Due to the almost identical structure of the side-chain compared with the lowmolecular<br />

weight liquid crystal (6CPB), most of the functionality-specific band<br />

assignments of Table 2-1 are also valid for the NLCP. The 1603 c1n-l and 1503<br />

cm-I (phenyl C=C-stretching vibrations) bands and the 2230 cm-' (v(C_N)) band<br />

are polarized along the long axis of the mesogenic functionality and show parallel<br />

dichroisni with All > AL. This means, that the long axis of the mesogenic unit is<br />

predominantly oriented parallel to the rubbing direction. The 757 cni- ' bund (ring<br />

CH out-of-plane deformation) exhibits a perpendicular dichroisni with Al > All<br />

because the transition moment for this band is perpendicular to the ring plane and<br />

hence, perpendicular to the long axis of the mesogen. Since the transition moment


50<br />

w<br />

CJ 1.50<br />

z<br />

3 1.00<br />

2<br />

s<br />

0. so<br />

w<br />

5 0.00<br />

4<br />

g -0.50<br />

ORIENTATION<br />

RELAXATION<br />

0 25 SO 75 100 0 100 200 300 400<br />

TIME / s<br />

TIME / s<br />

$ 1.50<br />

z<br />

4 2.00<br />

f 0. 50<br />

*<br />

W<br />

2 0.00<br />

k-<br />

4<br />

Fl -O. 50<br />

0.0 I. 5 3.0 4. 5 0<br />

TIME / s<br />

5 10 1s 20<br />

TIME / s<br />

Figure 2-16. Relative absorbance/time-plot during orientation and relaxation of NLCP at different<br />

temperatures for selected absorption bands: (0) 2926 cm-', (0) 2229 cm-', (*) 757 cm-I.<br />

for this band is also nearly prependicular to the traiis-polymethylene chain, this<br />

indicates that the spacer is oriented preferably parallel to the rubbing direction of<br />

the polyimide support and the mesogen. We used the 2130 and the 757 an-' bands<br />

for the characterization of the movement of the mesogen and the 2916 cm-' band<br />

for that of the spacer. Unfortunately, the bands which characterize the backbone<br />

are heavily overlapped with absorptions of the spacer and they are much less<br />

intense than the corresponding spacer bands (a monomeric unit contains one<br />

methylene group in the iiiaiii chain und six in the spacer). Therefore, in the case<br />

of this polymer, we cannot follow the inoveiiient of the backbone. Owing to the<br />

experimental geometry, the reorientation of NLCP led only to changes in intensity<br />

of light polarized parallel to the rubbing direction. For this reason, as before, we<br />

used only polarized radiation to monitor the molecular motion of the NLCP.<br />

The relative intensity changes during orientation and relaxation at different temperatures<br />

for selected absorptions bands are shown in Figure 2-16. As can be seen<br />

from this figure, the orientation and relaxation rates incrcased more than 20-fold<br />

when the temperature increased from 80 to 120 "C. At temperatures close to the<br />

glass transition, the orientation times increase to a few tens of minutes. This is<br />

apparently related to the decrease of viscosity with increasing temperature. At each<br />

temperature the relaxation takes more time than the reoreination. The A, values<br />

change less at higher temperatures than at lower temperatures (Figure 3-16). This is<br />

the consequense of the decrease in the order parameter of the NLCP with increasing<br />

temperature. The values of A, for different absorption bands depend on the angle


2.5 Electric Field-Induced Orientlition 51<br />

Figure 2-17. Normalized<br />

absorbaace/time-plot during<br />

orientation and relaxation of<br />

NLCP at dif'rerent temperatures<br />

for selected absorption<br />

bands: (x) 2926 ciii-l, i+)<br />

2229 cm-' .<br />

0 1 2 3 4 5 0 5 1 0<br />

m / r m / r<br />

between their transition moment and the molecular axis of the LC molecule. In<br />

order to coinpare the rates of molecular motion for the differents segments, the<br />

normalized absorbance A, was used. The A, values for bands characterizing the<br />

niesogen and the flexible part are shown in Figure 2-17. As is evident from this<br />

figure, no significant differences of their intensities can be detected during the orientation<br />

and relaxation process. Thus, it must be assumed, that the mesogen and<br />

the flexible part reorient as a rigid unit and the application of the electric field does<br />

not lead to conformational changes in the flexible spacer. This conclusion, based on<br />

the direct monitoring of the movement of different molecular segments, is in agreement<br />

with the results of [63] derived from the measurement of rotational viscosity of<br />

the same NLCP in a magnetic field and from theoretical modeling. However, it has<br />

been reported that the application of the electric field led to a decrease of trans<br />

conformers in a flexible spacer of an NLCP similar to that studied here. This conclusion<br />

was based on the analysis of the 1490-1410 cni-' region which is sensitive<br />

to the conformation of the polymethylene sequence. As has been shown recently<br />

for a set of side-chain liquid crystalline polymers with different length of the polymethylene<br />

spacer, this region also contains absorption bands of the aromatic ring<br />

[64]. In our opinion, the conformational analysis based on the absorbance in this<br />

region is thus very complex and ambiguous.<br />

If the main chain takes part in the reorientation, it is expected that its motion will<br />

be, at least partially, shifted in phase with respect to the side chain (spacer and<br />

rnesogen). A detailed 2D analysis [30] in the region of the in(CH2) bands (2936 and<br />

2863 cni-' ) does not reveal any significant asynchronicity. Hence, we may conclude<br />

that the main chain does not take part in the reorientation, or it moves only slightly.


Figure 2-18. Scheinatic representation<br />

of the segmental<br />

motions of the NLCP during<br />

the switching process.<br />

There is a strong correlation in the synchronous spectrum and no appreciable<br />

asynchronicity in the asynchronous spectrum between the mesogen band (2230<br />

cm-') and the v(CH2) bands (2926 and 2863 cm-I). The main part of the intensity<br />

changes in the region of 2850-2950 cm-' is due to the spacer; thus, on the basis of<br />

the 2D results we may draw the conclusion that the spacer and the mesogen reorient<br />

simultaneously.<br />

From a further detailed 2D analysis of selected absorption band intensities of the<br />

corresponding power spectrum (16, 60, 611 it was finally concluded that only part of<br />

the spacer of the NLCP takes part in the reorientation. This result is summarized<br />

graphically in Figure 2- 18, where the orientational behavior of a NLCP-mesogen<br />

during the switching process is symbolized schematically relative to the entire<br />

polymeric structure, including the spacer and the main chain.<br />

2.5.4 Nematic Liquid-Crystalline Guest-Host System in a<br />

Switching Experiment<br />

Generally, the phenomenon of dissolving and aligning a molecule or a group of<br />

molecules, such as dyes or probes, for example, by a liquid crystal can be called the<br />

guest-host phenomenon [65, 661. The host liquid crystal can be a single compound<br />

or a multicomponent mixture. Depending on its structural geometry (e.g., elongated<br />

or spherical) the guest molecule couples more or less significantly to the anisotropic<br />

intermolecular interaction field of the liquid-crystalline host. The liquid-crystalline<br />

solutions can be easily oriented by electric, magnetic, surface, or mechanical forces,<br />

resulting in highly oriented solvent and solute molecules. This phenomenon provides<br />

the basis for the application of liquid crystals as anisotropic solvents for<br />

spectroscopic investigations of anisotropic molecular properties (67, 681. Actually,<br />

in a broad sense. most of the commercially available liquid-crystal mixtures may be<br />

regarded as being related to the guest-host effect, as they incorporate some nonmesomorphic<br />

molecules to generate the desired electro-optical effects in the mixtures<br />

(69-711. In the present section we present an example of how a guest molecule<br />

is oriented anisotropically in a liquid-crystalline solution which is exposed to an


2.5 Electric Field-Induced Orientatioii 53<br />

GUEST-HOST SYSTEM<br />

ZLI - 1695<br />

Figure 2-19. Chemical structures and<br />

transition temperatures of the investigated<br />

nematic liquid-crystalline<br />

guest-host system.<br />

H<br />

I<br />

N<br />

2-NAPHTHALDEHYDE<br />

GCIEST<br />

external electric field in a switching experiment. Time-resolved step-scan FTIRspectroscopy<br />

has been applied to study how a solute follows the reorientation<br />

motion of the nematic solvent [72].<br />

2.5.4.1 Materials<br />

A 10% solution of 2-naphthaldehyde in the nematic liquid-crystal mixture 4-cyano-<br />

4'-alkylbicyclohexyls (ZLI-1695, Merck) (Figure 2- 19) was prepared for the measurements.<br />

This nematic phase is very convenient as an anisotropic solvent because<br />

of its weak IR spectrum, allowing one to obtain polarization spectra in practically<br />

the whole mid-IR region.<br />

2.5.4.2 Experimental<br />

For the switching experiment, the cell described in Figures 2-7 and 2-8 was utilized<br />

(here the path length was 10 pin). The spectral resolution was 16 cni-' and a function<br />

generator supplied the electrical signal (Figures 2-9 and 2-10) needed to orient<br />

the liquid-crystalline sample (16V,,, 1 kHz, 20 ms on, 260 ms off). The positive<br />

edge of this signal also synchronized the start of the data acqusition at each step<br />

during the complete step-scanning experiment. The time resolution of the experiment<br />

was 100 ps; each spectrum contains coadded data of two orientation/relaxation<br />

cycles.<br />

The reorientation behavior of the nematic solvent was determined by the v(C-N)<br />

and the v,(CH?) absorption bands at 2239 and 2857 cm-', respectively, whereas the<br />

v(C=O) absorption at 1698 cm-' was utilized to characterize the segmental motion<br />

of the solute (Figure 2-20).<br />

2.5.4.3 Results and Discussion<br />

The step-scan spectra of the studied solution, sampled at 0.0 and 19.5 ms, are pi-esented<br />

in Figure 2-20. The observed intensity changes correspond to the reorientation<br />

of the solute and solvent molecules from their preferential alignment in the<br />

rubbing direction (0.0 ms) into the applied electric field axis i 19.5 nisi. The detailed<br />

relative intensity changes of the three references bands versus time after the electric<br />

field was switched off are shown in Figure 2-21. Here, the lower values for the rela-


54 2 Segmental Mobility qf Liquid Cyvstrrls cmd Licliiid-Cvystalliiie Po1jwer.r<br />

I<br />

I<br />

3000 2400 is00 1200 600<br />

WAVENUMBER /[ cm-']<br />

Figure 2-20. Step-scan FTIR spectra of the investigated nematic liquid-crystalline guest-host<br />

solution before (0.0 ms) and after (19.5 ms) reorientation in the electric field.<br />

0.<br />

L<br />

2 0.<br />

4<br />

B<br />

I<br />

0 10 20 30 40 so<br />

Figure 2-21. Relative absorbance changes versus time of the v,(CH?) (0) and v(C=N) fV) nbsorption<br />

bands of the solvent and the solute-specific v(C=O) (0) band during orientation and relaxation<br />

of the individual segments.<br />

tive absorbance of the v(C-N) relative to the v(C=O) band are due to two effects:<br />

(1) better initial orienation in the rubbing direction; and (7) the more eficient<br />

reorientation of the C-N functionality into the electric field direction.<br />

The positive values for the relative absorbance of the vs(CH2) absorption reflect<br />

the preferential perpendicular alignment of the transition moiiieiit direction of this<br />

band relative to the long axis of the nematic solvent molecule. A recent study 1291


2.5 Electric Field-Induced Orienfation 55<br />

n 10 rn 30 40 50<br />

TIME / ms<br />

Figure 2-22. Normalized absorbance versus time during orientation and relaxation of the solventand<br />

solute-specific functionalities (see Figure 2-21).<br />

on the nematic liquid-crystal 4-cyano-4'-pentylbiphenyl (often denoted as 5CB)<br />

by means of time-resolved FTIR spectroscopy under the conditions of dynamic<br />

reorientation give evidence for different reorientation times of its non-ridgidly<br />

bonded fragments.<br />

Although Figure 2-2 1 shows that neither the orientation nor the relaxation of the<br />

solute and solvent molecules is complete under the experimental conditions, we<br />

expected that the delay phenomena between the solvent and solute molecules would<br />

be clearly expressed in a normalized absorbance versus time plot. This plot, however,<br />

as shown in Figure 2-22, does not provide any indications of such a behavior.<br />

In othei- words, the reorientation rates are equal for the solvent and the solute.<br />

Without drawing any general conclusions from a single example, the present results<br />

show that one may expect similar behavior in other nematic solutions.<br />

2.5.5 Reorientation Dynamics of a Ferroelectric Side-Chain<br />

Liquid-Crystalline <strong>Polymer</strong> in a Polarity-Switched<br />

Electric Field<br />

The surface-stabilized ferroelectric liquid crystals in the smectic C* (SmC*) phase<br />

are among the most interesting types of liquid-crystalline systems because of their<br />

potential applications in high-resolution flat panel displays and fast electro-optical<br />

devices [73-761. Within this class of compounds, ferroelectric liquid-crystalline polymers<br />

(FLCPs) have gained theoretical and practical interest as systems which combine<br />

the properties of polymers and ferroelectric liquid crystals. This combination is<br />

achieved by attaching the ferroelectric mesogen to a main chain via a flexible spacer


56 2 Segrnental Mohilitj~ of Liquid Crystals arid Lirjuid-C~.vstallirie <strong>Polymer</strong>s<br />

(mole ratio 0.95 : 0.05 : 2.70)<br />

C=O N<br />

I II<br />

? v<br />

o y 0 I 0<br />

8 "C 55 'C 73 'C<br />

G-S:e---,S,c---*I<br />

Figure 2-23. Chemical structure,<br />

coniposition and transition<br />

temperatures of the side-chain<br />

FLCP under investigation.<br />

[77, 781. Since the spacer is effectively decoupling the movement of the mesogen<br />

froni the polymer backbone, a different behavior of mesogen, spacer, and main<br />

chain can be expected in a reorientation process. So far, most of the research on the<br />

mobility of FLCPs has been limited to the determination of relaxation times by<br />

using dielectric spectroscopy [79] or response times by using electro-optical switching<br />

studies [80] without differentiation of the motions of individual functionalities<br />

in the macromolecule and only few time-resolved vibrational spectroscopic measurements<br />

are available on this subject [81-831. However, the movements of the<br />

individual segments control the reorientation process and hence, the study of their<br />

dynamics is a key to understand the behavior of an FLCP in an electric field. Based<br />

on the recent advances in data-acquisition techniques [ 161 we have attempted in<br />

the present work to follow the reorientation dynamics of the different segments of<br />

an FLCP from one stable state to another on the submillisecond time scale during<br />

poling of an electric field [83].<br />

2.5.5.1 Materials<br />

The chemical structure, composition and the transition temperatures of the investigated<br />

side-chain FLCP are shown in Figure 2-23. The molecular weight of this<br />

polymer is approximately 18800 and the synthesis and characterization have been<br />

described in detail elsewhere [84].<br />

2.5.5.2 Experimental<br />

The variable-temperature cell was the same as described in Section 2.5.2, the only<br />

difference being that a spacer with a uniform thickness of 5 pm was used between<br />

the electrodes. The empty cell and the FLCP were heated to 90 "C to introduce the


2.5 Elrc tric Field-Iiidircwl Orim tri tiori 57<br />

INFRARED BEAM<br />

Figure 2-24. Polai-ization geometry and<br />

mesogen alignment of the investigated<br />

FLCP in the polarity-switched electric field<br />

IA: polarity +. B: polarity -J.<br />

POLARIZATION<br />

polymer into the cell through the gap between the electrodes by capillary forces.<br />

After filling the cell, it was slowly cooled (O.S"C/min) and an AC-voltage of 30<br />

Vpp, IOHz, was applied during cooling with the aim to reduce defects in orientation.<br />

The FLCP was investigated in the SmC*-phase where two surface-stabilized<br />

orientational states exist, depending on whether the mesogens incline to the right or<br />

to the left relative to the rubbing direction of the polyimide support (Figure 2-24).<br />

The switching between these two states is possible by changing the polarity of the<br />

applied field. The propagation direction of the polarized 1R beam was perpendicular<br />

to the plane of the Ge-windows. The largest absorption intensity changes during<br />

the switching process were obtained by fixing the polarizer at 45" relative to the<br />

rubbing direction of the polyimide coating. All spectra were collected with a spectral<br />

resolution of 8 cm-' and a time resolution of 0.1 ms. For a more detailed<br />

characterization of the reorientation mechanism, the experiments were carried out<br />

at different temperatures (28 "C, 42 "C and 47 "C ( f 0.1 "C)). At the lower temperature<br />

only rapid-scan measurements were required during orientation at 20 V (DC),<br />

whereas at the two higher temperatures the step-scan technique had to be applied<br />

due to the much faster segmental movement. Here, for the switching between the<br />

two orientational states a bipolar rectangular signal with a voltage of 20 Vpp and a<br />

frequency of 20 Hz and 100 Hz was applied at 42 "C and 47 "C, respectively (Figure<br />

2-25). The spectral data collection was synchronized with the negative edge of the<br />

electric signal.<br />

2.5.5.3 Results and Discussion<br />

Polarization spectra of the preoriented FLCP are shown in Figure 2-26. For the<br />

characterization of the mesogen and the flexible spacer, the 1607 cm-' ( Y(C=C),,)<br />

and 2926 cm-' ( vas(CH?)) absorptions, respectively, were chosen. Unfortunately,<br />

no bands free of overlap can be attributed to the backbone. The band at 1020 cm-'<br />

(if( Si-0-Si)) is strongly overlapped by the i!,(C-O-C) absorption of the ester group<br />

and the band near 807 cm-' is also superimposed in a rather complex manner.<br />

For the characterization of the methyl substituents of the siloxane-backbone, the<br />

h'(Si-CH3) vibration at 1260 cm-' was chosen, although it contains a contribution<br />

from the I*~,,(C-O-C) vibration related to the mesogen. The 1607 and 1260 cm-'<br />

bands have parallel dichroism and the 2926 cm-' band has perpendicular dichro-


58 2 Segnientrrl Mobility of Liqriid Crystals aid LiCIuid-Ci.vJtulliiie Po1jwiei.s<br />

PERIODICRL EUEHT ..............<br />

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

TRIGGER PULSE .................<br />

DRTR RCOUISITION .1([1[([(/(1/1/ i-<br />

........-........ RESIILUTIoH z5us<br />

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

COHTRDL SIGHRL ..............m<br />

ZOU,~, ioonz<br />

, "------,<br />

I. 0<br />

1 ' 1 . 1 1 . 1 I<br />

><br />

-_---- __- -____ -.-<br />

Figure 2-25. Time scale of the<br />

periodical event. control signal and<br />

data collection sequence for the poling<br />

experiment of the FLCP (2OHz at<br />

0 10 28 I0 48 50 TIME/ns 42 "C, 100 Hz at 47°C).<br />

~<br />

- - ~<br />

POLARITY: @ t 47OC<br />

u<br />

3 1 2<br />

LL<br />

0<br />

!nn<br />

I<br />

n. 4<br />

1.6<br />

'3200 3flK 2800 2600 2400 2200 2000 1000 1600 1400 1200 1000 800 600<br />

WAVENUMBER / =in '<br />

Figure 2-26. FTIR-polarization spectra of the investigated FLCP prealigned at 47 "C and different<br />

polarity of the electric field (radiation polarized at 45" to the rubbing direction according to the<br />

geometry of Figure 2-74).<br />

ism. Due to the perpendicular transition moment of the 1260 cm-' band relative to<br />

the local chain axis it can be concluded that in the FLCP the mesogen and the<br />

spacer are aligned preferentially parallel to each other but are perpendicular to the<br />

backbone [16, 53, 831. The sinusoidal changes observed in the baseline of the spectra<br />

are a consequence of the interference of the IR-radiation between the Ge-windows,<br />

but do not disturb the evaluation of the absorption bands in the present case. A<br />

stack-plot of time-resolved spectra recorded during several cycles of polarity<br />

changes is shown in Figure 2-27. A good signal-to-noise-ratio is observed for all<br />

spectral regions, except where the intensity exceeds an absorbance of 1.7.<br />

Normalized absorbance/time-plots of the selected absorption bands are plotted<br />

for experiments at different temperatures in Figure 2-28. As can be seen from this<br />

figure, the reorientation at 47 'C takes only approximately 5 ms compared with


2500 2008 I s00<br />

WRVtNdMRtR CM-I<br />

Figure 2-27. <strong>Spect</strong>roscopic response of the FLCP upon changing the polarity of the applied electric<br />

field (1: 2926 cm-I, 2: 2854 cni-', 3: 1607 cm-'. 4: 1507 cm-I, 5: 1260 cm-I).<br />

I. 25<br />

1.00<br />

0. i5<br />

0. 50<br />

0. 25<br />

n. 00<br />

-0. 25'<br />

0. 0<br />

w u 1.2s<br />

z<br />

c 1.00<br />

m<br />

Q 0.75<br />

2 0.50<br />

0. 25<br />

1.0 2. 0 3.0 4.0 5. 0 6 0<br />

Tinds<br />

i<br />

0.00<br />

-0.2s<br />

a -0.50<br />

p ! o 5 10 15 20 25 30 35 40 45<br />

0 Time/ms<br />

z 1 . z<br />

!. 00<br />

0. i s<br />

0. 50<br />

0.25<br />

n. 00<br />

0. 25<br />

01<br />

"%,<br />

5 ID 15 20 iE<br />

Timelms<br />

Figure 2-28. Normalized absorbance/time-plots for poling experiments of Ihe FLCP at different<br />

temperatures. (a) 28 "C, rapid-scan measurement, 20 V (DC): (b) 42 "C, step-scan measurement,<br />

20 V,,, 20Hz; (c) 47"C, step-scan measurement, 20 V,,, 100 Ha, data acquisition for 2.5 periods.<br />

(1260 cm-' (o), 1607 cm-' (+), 2926 cni-' (0)).


about 1.5 s for 28 "C. The shape of the response curves are similar and can be<br />

described as exponential. The reorientation starts without time delay when the<br />

polarity of the applied field is changed. This means, that there is no retardation<br />

in the response of different structural segments. No distinct differences have been<br />

found in the orientation rates of the flexible and the rigid part of the molecules<br />

of ferroelectric nonpolynieric LCs [53, 581. However, it has been shown that the<br />

molecular motion cannot be described by a single exponential process but consists<br />

primarily of a fast and a subsequent slow process [52]. This motion also includes<br />

confornia tioiial changes in the flexible part of the molecule. In the polymeric ferroelectric<br />

system under investigation, such conforniational changes may take place<br />

in the spacer and the main chain. Due to the lack of conformationally sensitive<br />

absorptions in the FLCP, we were not able to detect or differentiate such changes.<br />

Taking into account the assignment of the absorption bands under consideration<br />

and their dichroic changes during poling the electric field, we can conclude from<br />

Figure 2-28 [83] that not only the mesogen but also the spacer and the main chain<br />

(at least some molecular groups which are in the neighborhood of the spacer attachment<br />

to the backbone) take part in the reorientation process during poling the<br />

electric field. This picture, however, is not supported by the 2D-correlation analysis,<br />

which shows, that there is a strong correlation between the ring-stretching bands<br />

(1506, 1607 cm-') and the two carbonyl bands around 1750 cni-' [16]. This is evidence<br />

that the mesogen reorients as a rigid unit. The strong correlation in the synchronous<br />

spectrum and no appreciable asynchronicity in the asynchronous spectrum<br />

between the mesogen bands and the v(CH2) bands (2926 and 2855 cin-'),<br />

which are mainly due to the spacer, indicates that the mesogen aiid the spacer<br />

reorient synchronously. Figure 2-29 compares the power spectrum of the FLCP to<br />

the reference spectrum (average spectrum of a reorientation process). One can easily<br />

notice, that the intensities of the spacer bands in the power spectrum are considerably<br />

reduced relative to the mesogen bands. Based on these data, we have to coiiclude<br />

that only a small part of the spacer, directly attached to the mesogen, takes<br />

part in the reorientation. The 2D analysis does not supply any evidence that the<br />

main chain of the FLCP participates in the reorientation. Based on these data, the<br />

picture shown in Figure 2-30 can be derived for the response of the FLCP to a<br />

poled electric field.<br />

2.6 Alignment of Side-Chain Liquid-Crystalline<br />

Polyesters Under Laser Irradiation<br />

It has been demonstrated that the photo-induced orientation of dye-containing<br />

SCLCPs can be used for reversible optical data storage [85-901. FTIR polarization<br />

spectroscopy has been applied to monitor the orientation of the main chain and the<br />

mesogenic side-chains which are of critical importance in these storage processes<br />

[15, 91-94]. Here, some results obtained by laser irradiation of a homologous series


-3.6 Aligmient OJ' Side- Clinin Liyuicl- Cv!~stnlline Poljwstevs 61<br />

161<br />

WAVENUMBER I cmd<br />

14 -<br />

12 -<br />

10 -<br />

c<br />

c<br />

zi<br />

(0 8-<br />

6-<br />

Figure 2-29. Comparison of the reference (top)<br />

and power (bottom) spectra intensities of the<br />

FLCP for the reorientation process during<br />

poling the external electric field.<br />

4r<br />

1500 2000 2500 3000<br />

WAVENUMBER /at-'<br />

Figure 2-30. Schematic representation of the<br />

segmental motions of the FLCP during poling<br />

the electric field.<br />

POLARITY SWITCH<br />

RESPONSE<br />

of novel SCLC-polyesters [ 151 are presented. In unoriented polyester films with a<br />

dodecamethylene spacing of the ester groups in the main chain, holograms with a<br />

resolution of 6000 linesjmm can be recorded by the interference of an Argon-ion<br />

laser beam [95]. The thermal erasure and rewriting of the laser-induced alignment in<br />

the SCLC polyesters is the basis for such storage processes. In order to investigate<br />

the mechanism of the involved processes in more detail, a variable-temperature,<br />

computer-controlled sampling system for on-line polarization FTIR nieasurements<br />

during laser irradiation as well as thermal erasure has been designed and<br />

constructed.


62 2 Segniental Mobility of Lipid Crystals and Liquicl-Cvystalliiie Po1jwiei.s<br />

SIDE-CHAIN<br />

p 2 ' m = main-chain spacer length<br />

n = side-chain spacer length<br />

P<br />

"=: ESTERLINKAGE<br />

R = azobenzene substituent<br />

Design parameter<br />

a) variation of the length of the methylene units in<br />

main-chain rn = 4, 10, 11, 12 ,I4<br />

and<br />

side-chain n = 6, 8, 10<br />

b) variation of the azobenzene substituent R<br />

R=CN(a), NO, (b), OCH, (c), H(d)<br />

Figure 2-31. General structure<br />

and nomenclature (PnRm) of the<br />

investigated SCLC polyesters used<br />

for reversible optical data storage<br />

(e.g., P6a12: polyester with 12<br />

CH? groups in the acid part of the<br />

main chain and six CH2-groups<br />

in the side-chain spacer and a<br />

CN-suhstitucnt on the mesogen).<br />

Table 2-2. Phase behavior of P8a12 [96].<br />

Structure 4 3 2 1<br />

Transition temperature range 30-40 "C 46-54 C 59-63 "C 65-71 "C<br />

2.6.1 Materials<br />

The variations in chemical structure of the investigated SCLC polyesters and their<br />

nomenclature is outlined in Figure 2-3 1. Differential-scaiining-calorimetry (DSC 1,<br />

polarization-optical-microscopy (POM) and X-ray investigations [95-971 reveal a<br />

rather complex phase behavior of this homologous polyester series, which strongly<br />

depends on the length (m) of the main-chain. For the polyadipates (m = 4), for<br />

example, an enantiotropic liquid-crystalline behavior is observed, involving disordered<br />

smectic mesophases in the range from T, at about 10°C to the sniectic tf<br />

isotropic transition (from about 55-70 "C, varying with the side chain spacer length<br />

n = 6,8, 10). The longer homologs, with a main chain length of 10-14 niethylene<br />

units could be classified as inonotropic LCPs, with polyinorphic character. Some<br />

of them exhibit up to four first-order transitions, which partially overlap, involving<br />

inany different organized phases. An exact phase assigninelit is very difficult. Candidates<br />

with the highest potential for optical data storage are the polytetradecanedioates<br />

(ni = 12) with 6 and 8 methylene groups in the spacer [95, 981. DSC and<br />

X-ray investigations reveal for both polyesters two low-order mesomorphic phases<br />

(structures 3 and 4) and two high-order mesomorphic or crystalline phases jstruclures<br />

I and 2). The glass-transition temperature of both systems is about 25°C.<br />

Representative of the discussed SCLC polyesters, the phase behavior of P8a12 is<br />

detailed in Table 2-2.<br />

The influence of azobenzene substituents with varying dipole moments has also


~ polyester<br />

2.6 Alignnient qf Side-Chaiii Liquid- Crystalline Polyrsters 63<br />

film<br />

polarization direction<br />

Figure 2-32. Simplified picture of the experimental<br />

set-up for laser-irradiation and symbolic<br />

representation of the motivation for the<br />

FTIR-spectroscopic detemiinatioii of the order argon.ion laser ,R beam<br />

parameters of the polyester functionalities.<br />

(488 nin)<br />

been studied, but no obvious iinproveinent of the data storage capability 1991 could<br />

be observed.<br />

Thin film of the polyesters were obtained by casting a chloroform solution<br />

(3mg/200 pl) of the polymer onto KBr plates with a 1.5 x 1Oinm' area. After<br />

evaporation of the solvent, the films were dried at room temperature under vacuum<br />

for 2 h. The film thickness obtained by this preparation procedure was approximately<br />

5 pni.<br />

2.6.2 Experimental<br />

Figure 2-32 provides a simplified picture of a laser irradiation/FTIR measurement<br />

experiment symbolizing the inherent problem to determine the laser-induced orientation<br />

of the different polymer segments. The detailed instrumental design of the<br />

apparatus used for these on-line ineasurements is given in Figure 2-33 and shows<br />

that the following experimental parameters can be varied:<br />

0 Based on an internally calibrated photo-diode, the laser intensity can be reduced<br />

from 800 mW/cin2 in a controlled fashion with a neutral-glass filter wheel.<br />

The irradiation time is adjusted by the shutter.<br />

The temperature of the sample is controlled by a Peltier element.<br />

The acquisition of the FTIR polarization spectra is performed by using the<br />

dedicated computer of the spectrometer, which is synchronized with the different<br />

devices. The laser beam passes through the shutter and the filter wheel and is<br />

directed via two mirrors onto the film sample, which is mounted on the Peltier<br />

heater/cooler arrangement. This set-up was used to follow the photo-induced<br />

anisotropy during and after irradiation as a function of time and temperature, as<br />

well as its thermal stability during erasing cycles. Time-resolved spectra with a<br />

resolution of 4 cin-' were taken with radiation polarized alternately parallel and


Figure 2-33. Instrumental details of the apparatus used for on-line FTIR measurements during<br />

laser irradiation.<br />

perpendicular to the argon-ion laser beam polarization direction, respectively. The<br />

irradiated sample area was about 3 mm’.<br />

2.6.3 Results and Discussion<br />

In the FTIR polarization spectra of the films prepared by casting from solution<br />

no dichroism could be detected before laser irradiation. Figure 2-34 shows the<br />

polarization spectra of a P8a12 polyester after irradiation with the 488 nm argonion<br />

laser line (laser power 800 mW/cm’, irradiation time 1000 s). A detailed assignment<br />

of the individual absorptions bands to specific structural units is given in<br />

Table 2-3. The observable dichroic effects clearly demonstrate the significant orientation<br />

induced in the different functionalities of the polymer. Thus, the mesogenic<br />

side groups are preferentially oriented perpendicular to the polarization direction<br />

of the laser beam. This picture is based on the 0-dichroism of the v(C-N) band and<br />

the aromatic ring modes (see also Table 2-3). Additionally, the 7c-dichroism of the<br />

v(CH2) absorptions indicates a perpendicular orientation of the side and main<br />

chains relative to the laser beam polarization direction, whereas the smaller K-<br />

dichroism of the ester linkage indicates a somewhat lower orientation. As the electromagnetic<br />

radiation influences only the orientation of the mesogenic groups with


I."<br />

0.9<br />

0.8<br />

Figure 2-34. FTlR 0.7<br />

polarization spectra of<br />

o, 0.6<br />

an irradiated P8a 12 0<br />

polyester (after 1000 s of 5 0.5<br />

irradiation with the 488<br />

0.4<br />

nm line of an at-gon-ion p<br />

laser at an intensity of a 0.3<br />

800 mW/cm' I. The IR<br />

0.2<br />

beam was polarized<br />

parallel and perpendicu- 0.1<br />

......... parallel<br />

- perpendicular<br />

lar to the polarization 0.0<br />

direction of the laser 3300 3000 2700 2400 2100 1800 1500 1200 900 600<br />

beam.<br />

Wavenumber I cm-1<br />

Table 2-3. Band assignment. wavenumber position and structural origin of selected absorptions of<br />

the polyester P8a12 (as = antisymmetric: s = symmetric; ar = aromatic: oop = out-of-plane).<br />

Band assignment Wavenumber Structural origin<br />

(cm-I)<br />

c+~(CH?)<br />

I's(CH2)<br />

v(C-N)<br />

I'(c=o)<br />

1*(C=C),,<br />

l'( C=C),,<br />

t ~ CH2) (<br />

is,,( C-0- Ar)<br />

Various modes of coupled<br />

COOR group vibrations<br />

I*( N-Ar)<br />

11, (C-0- Ar)<br />

jjooP (CH I,,<br />

;',(CH?)<br />

2920<br />

2851<br />

1379<br />

1733<br />

1603 / 1583<br />

1501<br />

1403- 1355<br />

1257<br />

1201 . . . 1 I05<br />

1141<br />

1022<br />

839<br />

723<br />

main chain / spacer<br />

main chain / spacer<br />

mesogen<br />

main chain / (ester linkage)<br />

mesogen<br />

mesogen<br />

main chain / spacer<br />

mesogen / spacer<br />

main chain / (ester linkage)<br />

mesogen<br />

mesogen / spacer<br />

mesogen<br />

main chain / spacer<br />

the azobenzene moieties. the alignment of the spacer and the main chain can only<br />

be induced by coupling effects. Investigations of polyesters with specifically deuterated<br />

aliphatic sites in the spacer and main chain demonstrate, that the perpendicular<br />

alignment of the acid-specific main chain relative to the laser polarization<br />

direction is even better than the alignment of the side-chain spacer units 1001. Thus,<br />

Figure 2-3% shows the region of the v(CH2) and v(CD2) polarization spectra of a<br />

P6a12 polyester with a completely deuterated main chain (except the CH-group


66<br />

i v(CHZ)<br />

fi<br />

0)<br />

m C<br />

*<br />

B<br />

P<br />

4<br />

a<br />

3000 2800 2600 2400 2200 2<br />

Wavenumbers (cm-1)<br />

' n<br />

v(C=N)<br />

I<br />

2300 2250 2200 2150 2100 2050 2000<br />

Wavenumbers (cm-q)<br />

30<br />

Figure 2-35. (a) FTIR polarization<br />

spectra of'the v(CH?), \s(C-N) and<br />

v(CD2) absorptions of a specifically<br />

denterated P6a13 (see text) after laser<br />

irradiation. (b) FTIR polarization spectra<br />

of the v(CD7) atid i(C=N) absorptions<br />

of a specifically deuterated<br />

P6a12 (see text) after laser ii-radiation.<br />

(IR radiation polarized parallel f. . .)<br />

and perpendicular (-i to the laser<br />

polarization).<br />

Laser polarization<br />

direction<br />

CN<br />

I<br />

Figure 2-36. Schematic model of the irradiation-induced segmental<br />

alignment in the polyesters relative to the polarization direction of the<br />

argon-ion laser beam.<br />

where the spacer merges into the main chain) and Figure 2-35b demonstrates the<br />

almost negligible dichroism of the v(CD*)-region of a polyester which has been<br />

deuterated in the alcoholic part of the main chain [loo]. On the basis of these<br />

experimental results, the model shown in Figure 2-36 can be derived for the laser-


induced alignment of the investigated polyester and the following scenario is suggested<br />

for optical storage: the film obtained by casting from solution consists of'<br />

macroscopically isotropic niicro-domains. The irradiation with linearly polarized<br />

laser light leads to a melting of these domains, caused by direct heating due to<br />

absorption or a deposition of heat through traris-cis-trms isonierization cycles of<br />

the azobenzeiie moieties [95]. The origin of the orientation effect is attributed to<br />

tivm,r-cis photoisomerization of the azo-group, followed by a c-is-trans thermal or<br />

photoisomerization. In the cis-form, this functionality is able to undergo some<br />

reorientation with respect to its original alignment. The absorption of the polarized<br />

'writing' beam will continue as long as a component of the electric dipole vector of<br />

the azo-group lies in the direction of the laser polarization. The trurzs-cis reorientation<br />

process will therefore continue until all the azo-groups have been oriented<br />

perpendicular to the polarization vector of the laser beam [86. 1011. In the polyesters<br />

investigated here, this orientation remains after the 'writing' beam is turned<br />

off and a long-tenn stable, macroscopic orientation is established.<br />

Apart from this long-term stability, the possibility to erase and reinduce this<br />

anisotropy in the polymer is a prerequisite for its application as a reversible optical<br />

data-storage medium. To demonstrate the loss of mesogen-alignment during heating<br />

and the perfect reversibility of a 'write-erase-write' cycle, Figure 2-37a shows a<br />

stack plot of the i(C-N)-band region in the polarization spectra of a P10a12 film<br />

during heating from 32 "C to 100 "C (isotropization temperature, 68 "C) and in<br />

Figure 2-37b the dichroism of the v(C-N)-band of the same polymer is shown<br />

(from left to right) after irradiation, upon heating the irradiated sample to 100°C<br />

and after reirradiation. The aliiiost perfect coincidence of repeated writing cycles<br />

is demonstrated in more detail in Figure 2-38a, which displays the simultaneous<br />

photo-induced alignment of the different segments during laser irradiation in an<br />

order parameter/tinie-plot. Figure 2-38b exhibits the intermediate thermal erasure<br />

of the first orientation as a function of temperature. The loss of orientation for the<br />

different polymeric segments is also coupled and takes place in a narrow teniperature<br />

interval. A comparison of Figure 2-38b with the transition temperature ranges<br />

of the individual phases in Table 2-2 shows that, during erasure, the anisotropy<br />

is mainly retained in the less ordered mesophases 3 and 4. Beyond 54 "C, only a<br />

small, residual anisotropy is observed due to recrystallization effects [96], which<br />

completely disappears at the clearing point of phase 2.<br />

2.7 Orientation of Liquid Crystals Under<br />

Mechanical Forces<br />

Apart from the orientation mechanisms discussed in the previous sections, a preferential<br />

alignment of liquid crystals or liquid-crystalline polymers may also bc<br />

induced by the application of mechanical forces. Depending on the investigated<br />

system, this can be either done by shearing between plates or in stress-strain niea-


100 “C<br />

2300 2200 2100<br />

walenumbers [cm-’]<br />

I00 “C<br />

2300 2200 2100<br />

wavenumbers [cm-11<br />

1<br />

e<br />

2<br />

0 0.20-<br />

z -<br />

4<br />

80<br />

wavenumbers [cm-l J<br />

Figure 2-37. Thermal erasure of a laser-induced alignment in a PlOal2 polyester demonstrated by<br />

the disappearance of the v(C=N) dichroism at the isotropization point (68°C) (a) and the effect of<br />

irradiation, thermal erasure and reirradiation on the v(C-N) dichroism of the same polymer (b).<br />

surements of bulk samples [36, 102-1041. <strong>Spect</strong>roscopic data on the orientation<br />

induced by these pretreatments may then be utilized for a better understanding of<br />

the different responses of specific functionalities to the external force field, or they<br />

can establish the basis for a correlation of the mechanical properties and the structure<br />

of the liquid-crystalline system.


2.7 Orientation oj Liquid Ci:i,stals Under Mechanicul Forces 69<br />

a<br />

b<br />

solid symbol 1st writing cycle, open symbol 6th wntlng cycle<br />

1 ' 1 9 I , I I I ' I<br />

0 200 400 600 800 1000 30 35 40 45 50 55 60 65 70 75 80<br />

Irradiation time Is Temperature I "C<br />

Figure 2-38. Reversibility of repeated write/erase-cycles demonstrated in time-resolved FTIR<br />

polarization measurements by the order parameter for the different segments of PSal?: (a) As a<br />

function of irradiation time (488 nm line of an argon-ion laser, at 30"C, SOOmW/cm') in the 1st<br />

and 6th cycle; and (b) Versus temperature during thermal erasure with a heating rate of 30 "C/niin.<br />

2.7.1 Shear-Induced Orientation in Side-Chain Liquid-Crystalline<br />

Copolysiloxanes<br />

As pointed out already in Section 2.5.5, low-molecular weight ferroelectric liquid<br />

crystals (FLCs) and FLCPs are attracting a lot of interest because of their potential<br />

for electro-optical applications. The polymers offer new possibilities, e.g., as elastomers<br />

for piezoelectric elements or by copolymerization [77, 78, 1051 due to the<br />

formation of intrinsic mixtures between SniCY' mesogenic units and other coinonomers.<br />

This leads to FLCPs combining several material properties which might be<br />

utilized for colored displays in the case of comonomers containing chi-omophores.<br />

For the differentiated evaluation of such copolymers with reference to the possible<br />

exploitation of nonlinear optical (NLO) properties, the interplay of the different orientation<br />

tendencies of the side-chain functionalities is of crucial importance [36, 1061.<br />

2.7.1.1 Materials<br />

The copolymers presented here are copolysiloxanes whose liquid-crystalline phase<br />

behavior is characterized by the formation of a SmC* phase. By the additional


70 2 Segrnentrrl Mobility of Liquid Crystals and Liquirl-Ci.vstalliiit1 Po1jwiei.r<br />

r-<br />

-45 % -47 70 -8 %<br />

- [ .I-[ 7- f - .}<br />

4H2 41<br />

C=O<br />

c=o<br />

I<br />

?.<br />

H 3C-FH<br />

C=O<br />

I<br />

9<br />

CH2 -CH3<br />

chiral mesogen<br />

phase transition<br />

main chain + mesogen<br />

copolymer<br />

98<br />

(- 8”h) Figure 2-39. Chemical structure, coiiiposition<br />

and phase-transition temperatures of the<br />

different investigated copolymers.<br />

+ ChreNphore n<br />

introduction of chromophores with extended n-conjugation it was attempted to<br />

combine the ferroelectric properties of the mesomorphic phase with the NLOproperties<br />

of the chromophore. The synthesis has been described in detail 11061, and<br />

Figure 2-39 shows the structure and composition of the investigated homopolymer<br />

(containing no chromophore) and the modified copolymers alongside their corresponding<br />

phase-transition temperatures.<br />

2.7.1.2 Experimental<br />

Oriented polymeric films were obtained in a mechanical force field by applying<br />

a shear stress. The NaCl plates used as sample supports were coated with a thin<br />

polyiinide layer to improve the surface properties. The shearing process along a<br />

defined direction was first performed in the isotropic state above the clearing temperature<br />

and then at the isotropic/SmC*-transition. Afterwards, the temperature<br />

was gradually decreased to room temperature. The mechanical process was simultaneously<br />

analyzed using polarization optical microscopy. The experimental set-up<br />

is shown in Figure 2-40. For the subsequent dichroic measurements the samples<br />

were mounted in the spectrometer so that the shearing direction and the polarization<br />

direction of the incoming radiation were initially parallel. By rotating the


2.7 Orieritation of Liquid Ciysttrls Under Mechanictrl Forces 71<br />

Figure 2-40. Experimental set-up for the shearing experiment<br />

combined with POM iiieasureinents (A: microscope,<br />

B: goniometer, C: fixed NaCl crystal, D: shearing NaCl<br />

crystal, E: temperature control).<br />

polarizer for 90" the spectrum with perpendicular polarization relative to the<br />

shearing direction was recorded.<br />

2.7.1.3 Results and Discussion<br />

The question to be clarified by the FTIR spectroscopic polarization studies was<br />

whether:<br />

the interaction between the polymer chain and the side-on linked mesogenic dye<br />

functionalities dominates and the chroniophores orient parallel to the polyi~ier<br />

chain [ 107, 1081 and thus perpendicular to the non-chroniophoric mesogens<br />

(Figure 2-41a); or<br />

the interaction between the different mesogens (non-chromophoric and chromophoric)<br />

is dominating and leads to their parallel alignment (perpendicular to the<br />

chain direction) (Figure 2-41b).<br />

In order to determine the orientational behavior of the main chain, the spacer and<br />

the mesogenic and chroniophoric side groups relative to the shearing direction,<br />

detailed band assignments of the monomers, homopolymer and copolymers preceded<br />

the dichroic measurements of the sheared polymers [36, 1061. Based on these<br />

investigations, the polarization spectra of the sheared copolymers I and I1 (Figure<br />

2-42) yielded the surprising qualitative result, that the orientation of the side-on<br />

fixed chroniophores is doiniiiated by the interaction with the mesogenic groups and<br />

leads to the structure represented in Figure 2-41b. Thus, the dichroisiii of specific<br />

absorptions indicates only low orientation effects for the main chain parallel to the<br />

shearing direction, whereas the mesogens and the spacers clearly orient preferentially<br />

perpendicular to the direction of shear. The o-dichroisni of the v,(NO?)-<br />

absorption (Figure 2-42) supports the perpendicular alignment of the chromophore<br />

long axis relative to the shearing direction, and thus the validity of the model shown<br />

in Figure 2-4 1 b [36]. These dye-containing copolymers luay therefore not be LIS~~LI~<br />

for nonlinear optics [ 1061.


. I<br />

polymer chain<br />

0 0<br />

“I<br />

chromophore with<br />

a donor I acceptor System<br />

(“sideon” - Ilnk)<br />

k<br />

cn,<br />

/\<br />

H,C<br />

chiral mesogen<br />

(“endon“ - Ilnk)<br />

chromophore<br />

3‘<br />

..-. I<br />

. .<br />

’.<br />

, polymer chain<br />

’. . ... .. .<br />

.. ’..<br />

spacer<br />

oriented<br />

mesogen<br />

Figure 2-41. Possible models of mesogen<br />

and chromophore alignment in the sheared<br />

FLC-copolymers. (a) The chromophore<br />

orients parallel to the local polymer chain<br />

axis and the shearing direction, whereas the<br />

mesogen aligns perpendicular to the chromophore.<br />

(b) The interaction between the<br />

chromophores and the mesogens dominates<br />

and induces a parallel alignment to each<br />

other and a perpendicular orientation to the<br />

shearing direction.<br />

2.7.2 Orientation Mechanism of Liquid Single Crystal Elastomers<br />

During Cyclic Deformation and Recovery<br />

The application of the previously discussed techniques to induce monodomain<br />

structures in side-chain liquid-crystalline polymers by the application of electric or<br />

electromagnetic fields, by shearing or on anisotropic surfaces, frequently leads to<br />

comparatively low, macroscopically uniform orientation. Additionally, the methods<br />

are limited to a sample thickness of about 100 pm. Liquid-crystalline side-chain<br />

elastomers do not have this restriction, because a high macroscopic orientation can<br />

be induced in polymeric networks by mechanical deformation up to a sample<br />

thickness of about a centimeter [103, 1091. The synthesis of such systems can be<br />

performed by crosslinking linear, side-chain liquid-crystalline polymers to networks<br />

[ 1 lo]. The inherent combination of rubber elasticity and liquid-crystalline phase<br />

behavior, may then be exploited for the induction of a macroscopic mesogen orientation<br />

by mechanical deformation.<br />

Liquid single crystal elastomers (LSCEs) are characterized by a high macroscopically<br />

uniform orientation in the liquid-crystalline phase without the require-


0 8<br />

18<br />

16<br />

1<br />

a<br />

W<br />

u<br />

z a<br />

m<br />

a<br />

0<br />

Ln<br />

m<br />

Q<br />

0 6<br />

0 4<br />

0 2<br />

0 01<br />

1 4<br />

1 2<br />

10' I<br />

0 0.<br />

a 6<br />

0 4<br />

0 2<br />

0 0<br />

,<br />

I<br />

WAVEPJUMBER / [ cm-'I<br />

b<br />

WAVENUMBER / I cm?I<br />

Figure 2-42. FTIR polarization spectra of the copolymers with chromophore I (a) and I1 (b) after<br />

shearing (IR-radiation polarized parallel (- - -) and perpendicular (-) to the shearing direction) (the<br />

arrow indicates the \),(NO?) absorption of the chromophore).<br />

ment of an external stabilizing (mechanical) field. A characteristic feature of this<br />

state of order is the optical transparency of the material. The optical behavior of<br />

such elastomers is comparable with organic or inorganic single crystals. Kiipfer and<br />

Finkelinam [ 1 1 11 have described the synthetic route for LSCEs with a permanent<br />

mesogen orientation.<br />

2.7.2.1 Materials<br />

The LSCE system investigated in the present work is based 017 a methylsiloxane<br />

elastomer which has been crosslinked in a two-step procedure (Figure 2-43 j. The<br />

mesogenic side groups have been attached via a vinyl end-group (compound 2) and<br />

the crosslinking agents are a hydrochinonedialkylether with vinyl end-groups


74 2 Seyrneiitiil Mobility qf' Liquid Cvj~stul~ iiizd Liquid-Cryrtulliiie Polyriievs<br />

L)FFORRl.AI~ION<br />

-%I {E".l<br />

CROSSLlhKING UNDER -<br />

STRESS<br />

.Y LSCE<br />

\<br />

CROSSLINKING WITHOUT<br />

DEFORMATION<br />

I>F.I'ORMATION<br />

- =<br />

1 REFERENCE ELASTOMER<br />

CH3 4<br />

g,n 3°C n,i 83OC LSCE<br />

g,n 2 OC n.i 81 'C REFERENCE ELASTOMER<br />

Figure 2-43. Cheinical structure, transition temperatures and schematic presentation of the crosslinking<br />

processes involved in the synthesis of the investigated LSCE and reference elastomer [l 1 11.<br />

(compound 3) and a vinyl- and niethacryloyl-substituted derivative (compound 4).<br />

which has also liquid-crystalline properties, The key point of the crosslinking reactions<br />

is the fact that, under the experimental conditions, the addition of the vinyl<br />

group is faster by approximately two orders of magnitude [ll 11. Thus, in the first<br />

step, a network with a low crosslink-density is synthesized because predominantly<br />

compound 3 becomes active. In the second reaction step, the mesogens of this<br />

network are macroscopically oriented by application of a stress at elevated temperature<br />

and the induced anisotropy is fixed in the system by crosslinking via the<br />

niethacryloyl groups. The crosslinking reactions were performed in the nematic<br />

phase of the system. Part of the methoxy functionalities in compound 2 have been<br />

replaced by nitrile groups to provide an IR-active probe for the Orientation measurements.<br />

Additionally to the LSCE system, a structurally identical reference<br />

elastomer was investigated, which had not been subjected to a mechanical deformation<br />

during the second crosslinking step (Figure 2-43). In the reference polymer<br />

the mesogens do not exhibit a macroscopic orientation and, contrary to the LSCE,<br />

this material is opaque.<br />

2.7.2.2 Rheo-Optical FTIR <strong>Spect</strong>roscopy<br />

The increased need for more detailed data and a better understanding of the mechanisms<br />

involved in polymer deformation and relaxation has led to the search


2.7 Orientation oj Liquid Crystals Uiider Meclimiicul Forces 75<br />

Figure 2-44. (a) Principle of rheo-optical FTIR spectroscopy of polymer films. (b) Variabletemperature<br />

stretching machine for rheo-optical FTIR and FT-Raman spectroscopy. (A) stretching<br />

machine; (B) sample transfer mechanism; (1) stress transducer; (71 strain transducer: (3) film sample;<br />

k4'1 sample clamps [17, 181.<br />

for new experimental techniques to characterize transient structural changes<br />

during mechanical processes. With the advent of rapid-scanning FTIR spectroscopy,<br />

rheo-optical measurements have been applied to obtain data on the orientation,<br />

conformation, and crystallization during mechanical treatment of a wide<br />

variety of polymers [14, 17-20, 1121. The experimental principle of rheo-optical<br />

FTIR spectroscopy is based on the simultaneous acquisition of polarization spectra<br />

and stress-strain diagrams during deformation, recovery or stress-relaxation of<br />

the polymer under investigation in a miniaturized, computer-controlled stretching<br />

niachine which fits into the sample compartment of the spectrometer (Figure 2-44).<br />

The impleinentatioii of a heating cell as a closed, nitrogen-purged system offers the<br />

additional possibility to study the above mechanical phenomena under controlled<br />

temperature conditions ( i 0.5 "C) up to 200 "C. The electromechanical apparatus,<br />

the experimental procedure and the evaluation of the relevant parameters from<br />

the polarization spectra series have been described in detail in several publications<br />

[17-20, 1121.<br />

2.7.2.3 Near-Infrared <strong>Spect</strong>roscopy<br />

For samples exceeding the thickness accessible for ti-ansmission measurements<br />

by mid-infrared spectroscopy, near-infrared (NIR) spectroscopy may be used as<br />

an alternative technique [ 1131. Adjacent to the conventional mid-infrared region<br />

14000-400 cni-l) the NIR region covers the interval from about 12 500-4000 cm ' .


76 2 Segnrental Mobilitv of Liquid Crystals cmd Liquid-Crystalline Poljvners<br />

The absorption bands observed in this region can be assigned to overtones and<br />

combinations of characteristic fundamental frequencies. Apart from analytical<br />

applications, NIR spectroscopy has been demonstrated to be of value also to rheooptical<br />

applications [17, 114, 11 51. Thus, although NIR spectra are generally less<br />

informative than their mid-infrared analogs due to overlap of absorption bands, the<br />

details of the changes in crystallinity, conformation, and orientation are also contained<br />

in the overtone and combination vibrations [I 7, 11 31.<br />

2.7.2.4 Experimental<br />

The rheo-optical investigations were performed in the stretching machine shown in<br />

Figure 2-44b. For this purpose, film samples of the LSCE and the reference polymer<br />

(sample thickness about 300-350 pni, original length 8 mm, width 5 mm) were<br />

subjected to an elongation-recovery cycle up to about 45% strain, with a strain<br />

rate of ll'l/o/min and 20-scan interferograms with a spectral resolution of 5 c1n-l<br />

were collected siinultaneously with radiation alternatingly polarized parallel and<br />

perpendicular to the drawing direction. After completion of the experiment the<br />

interferograms were transformed to the corresponding spectra and evaluated in<br />

terms of the order parameter (orientation function) [14, 17-20, 1121 of selected<br />

absorption bands. The experiments discussed here were performed at 27 "C. In<br />

order to study the reorientation mechanism of the preoriented LSCE [ 1 16- 12 I]<br />

and the isotropic reference elastomer during the mechanical treatment of the rheooptical<br />

experiment, the samples were mounted in the clamps of the stretching<br />

machine after a rotation of 90" from the deformation direction in the second<br />

crosslinking step (Figure 2-45). Thus, in the LSCE the mesogens are originally oriented<br />

perpendicular to the subsequent drawing direction. This is reflected in the<br />

direction of deforni;ition<br />

during crosslinking<br />

Q<br />

0%) deformation<br />

direction of mechanical<br />

force during rheo-optical<br />

experiment<br />

maximum strain 45%<br />

n<br />

change in director<br />

orientation during deformation<br />

8<br />

polarization direciian<br />

of NIR-radiation<br />

transparent opaque transparent<br />

Figure 2-45. Experimental sequence for the observation of the director reorientation during the<br />

rheo-optical elongation-recovery cycle.


2.7 Orientation of Liquid Crjvtals Under Mechuniccil Forces 77<br />

2 00<br />

f-<br />

I50<br />

0 50<br />

0 25<br />

7000 6500 6000 5500 5000 4500 4000 3500 3100<br />

wavenumlxrs [cin-l]<br />

I 00<br />

0 50<br />

0 00<br />

2300 2200 2100<br />

NIR MIR<br />

Figure 2-46. FTIR/FTNIR polarization spectra of the undrawn LSCE system (the radiation<br />

was polarized parallel and perpendicular to the subsequent drawing direction of the rheo-optical<br />

experiment).<br />

dichroic effects of the polarization spectra taken before the elongation procedure of<br />

(Figure 2-46). In contrast, no dichroisin can be observed in the undrawn reference<br />

polymer (see below).<br />

Due to the large thickness of the samples, the spectrometer had been modified to<br />

collect NIR spectra extending into the mid-infrared region ( 10 000-2000 cm-’ )<br />

(tungsten-halogen source, CaF1-beamsplitter, InSb-detector). Despite an extensive<br />

band-overlap in the NIR-region, several absorptions in the spectrum of Figure 2-<br />

46 could be assigned by spectral comparison with low-molecular weight reference<br />

compounds [36]. Here, however, only the data derived from the v(C=N) absorption<br />

will be discussed in some detail.<br />

2.7.2.5 Results and Discussion<br />

In Figure 2-47 the stress-strain diagram of an elongation-recovery cycle up<br />

to 43% strain at 27°C is shown for the reference elastomer. Further elongationrecovery<br />

cycles do not reflect significant differences. The change in order parameter<br />

of the v(C-N) absorption band during the mechanical treatment is shown in Figure<br />

2-48. Throughout the whole experiment the drawing direction was the polarization<br />

reference direction. Due to the crosslinking conditions, no orientation (S = 0) can<br />

be observed for the original sample and uniaxial elongation leads to a linear increase<br />

of the order parameter up to S = 0.45 at the maximum strain of 43%. During<br />

recovery, an almost linear decrease of S can be observed with a small residual<br />

orientation of the mesogens upon return to the original starting position. This is in<br />

agreement with the hysteresis effect and the permanent elongation obserL ed in<br />

the stress-strain diagram (Figure 2-47). When the structural absorbance A0 of<br />

the v(C-N) absorption is plotted as a function of elongation 1361, an almost linear<br />

decrease is observed which is in good agreement with the theoretically calculated


78 2 Segnientril Mobility of Lipid Crystals uizd Licliri~-Cr)istulliiie Polvrwrs<br />

0.07 -<br />

0.06 ..<br />

-<br />

N 005 .*<br />

Figure 2-47. Stress-strain diagram of<br />

0 10 20 30 40 an elongation-recovery cycle of' the<br />

STRAIN I"/]<br />

reference elastomer at 27 "C.<br />

0. 60<br />

S<br />

0. 40<br />

0. 20<br />

Figure 2-48. Order parameter/<br />

strain-plot for the v(C-N)<br />

I .<br />

I I I I I I 1 I absorption corresponding to the<br />

10 20 30 10 ' 40 30 20 10 elongation-recovery cycle shown<br />

STRAIN [Yo] in Figure 2-47.<br />

decrease in sample thickness based on the assumption of a constant sample volume<br />

during elongation [30, 361. A more complex behavior is observed during the elongatioii-recovery<br />

cycle of the LSCE system to a maximum strain of 45% at 27°C.<br />

The corresponding stress-strain curve is shown in Figure 2-49 and has been subdivided<br />

into different strain intervals. The order parameter of the v(C-N) absorption<br />

derived from the spectra acquired during this mechanical treatment is presented<br />

in Figure 2-50. Here too, the analogous strain intervals are indicated to assist<br />

the correlation and interpretation of the mechanical and spectroscopic changes.<br />

Due to the anisotropy of the starting material and the reorientation of the mesogens<br />

during the mechanical treatment, specific assumptions in terms of the reference<br />

direction were necessary. Thus, for the derivation of the order parameter during<br />

elongation between 0 and 32% strain (intervals I and I1 in Figure 2-50), the deformation<br />

direction during the second crosslinking step (which is perpendicular to the<br />

applied deformation in the rheo-optical experiment) has been chosen as polarization<br />

reference direction. Beyond the inversion point of S = 0, the machine direction of<br />

the rheo-optical experiment is applied as polarization reference (interval 111). During<br />

recovery this procedure is reversed. Therefore, the order parameter always has<br />

a positive value throughout the whole experiment. From Figure 2-50 the following<br />

picture can be derived for the orientational changes taking place during the<br />

elongatioii-recovery cycle:


2.7 Orientatioiz of Liquid Crystals Under Mechanical Forces 79<br />

Figure 2-49. Stress-strain diagram<br />

of an elongation-recovery cycle of<br />

the LSCE at 27 “C.<br />

0 10 20 30 40<br />

STRAIN [“/.I<br />

I). so-<br />

Figure 2-SO. Order parameter/<br />

strain-plot for the vjC_N)<br />

absorption corresponding to<br />

the elongation-recovery cycle<br />

shown in Figure 2-49 (see<br />

text).<br />

s 0.40-<br />

0.30.<br />

0. 20-<br />

0.10~<br />

S’I’RALN [%.I<br />

During interval I the stress increases linearly with strain and no significant<br />

changes in the mesogen orientation can be observed.<br />

When the external stress (a,) reaches the internal stress (0,) of the LSCE systein<br />

(induced by the second crosslinking procedure) [lll] at about 15% strain, the<br />

stress-strain curve levels off and siinultaneously the order parameter decreases<br />

rapidly towards zero (interval 11) (Figures 2-49 and 2-50).<br />

During interval 111, the reorientation of the niesogens in the direction of the<br />

external field takes place until maximum elongation. This procedure is accompanied<br />

by a slight decrease in stress due to the cooperativity of the reorientational<br />

motion after exceeding the internal stress of the LSCE.<br />

Depending on the experimental conditions (temperature, niaxiinuni strain), the<br />

described phenomena are completely reversible upon recovery (Figure 2-51).<br />

Due to the intrinsic anisotropy in the original LSCE network, the spectroscopically<br />

monitored inversion point of the order parameter (S = 0) could principally<br />

be interpreted by the three different mesogen alignments shown schematically in<br />

Figure 2-52 [36]. For the elucidation of this question the structural absorbance A,,


Figure 2-51. Reversible dichroic behavior of<br />

the v(C-N) absorption of the LSCE during an<br />

elongation-recovery cycle to 45'%1 strain at<br />

'7 'C.<br />

C<br />

H b<br />

es~ = merogenlongsxls<br />

e<br />

Figure 2-52. Possible mesogen orientations<br />

during the reorientation process at S=O (ul:<br />

direction of internal stress induced during the<br />

second crosslinking step. 0,: direction of external<br />

stress applied during the rheo-optical<br />

experimentj. (a I<br />

Homeotropic orientation<br />

perpendicular to the film plane: (b) isotropic<br />

orientation; (c) planar biaxial orientation.<br />

3.40 0. 80<br />

3. 20<br />

S<br />

I<br />

0. so<br />

0.40 s<br />

2.60<br />

2.40<br />

I I I I<br />

10 LO 30 40<br />

S'IKAIN IYuI<br />

n. w)<br />

0. 20<br />

0. 10<br />

Figure 2-53. Structural absorbance<br />

and order parameter/strainplots<br />

for the v(C3N) absorption<br />

corresponding to the elongation of<br />

the LSCE to 45% strain at 27°C.<br />

of the v(CsN) absorption has been plotted alongside the order parameter for the<br />

LSCE elongation step in Figure 2-53. The reference direction for the evaluation of<br />

the structural absorbance has been chosen in accordance with the assumption made<br />

for the order parameter (see Figure 2-50). As can be seen from Figure 2-53 the<br />

structural absorbance decreases linearly up to about 25% strain. This is in agreement<br />

with the reference polymer and can be attributed to the decrease in sample<br />

thickness. Between 25'%, and 35% strain. however, the structural absorbance shows<br />

an abrupt increase followed by only a slight decrease up to maximum elongational


2.8 Conclusions 81<br />

strain. As A0 is an experimentally determined parameter based on a uniaxial orientation<br />

model whose value should represent the intensity of an absorption band<br />

exclusive of orientation effects, the anomalous increase of the structural absorbance<br />

during elongation-where the sample thickness actually decreases-points towards an<br />

invalidity of the uniaxial syinmetry. T~us, the geometries a and b of Figure 2-52 can<br />

be excluded on the basis that both would lead to a decrease of A" during elongation.<br />

The occurrence of the structural model c at the inversion point of the order<br />

parameter has first been proposed on the basis of small-angle X-ray measurements<br />

by Finkelinann and Kiipfer [117]. Thus, the appearance of four diffraction maxima<br />

have been explained in terms of a biaxial, planar orientation of the mesogens in the<br />

film plane (Figure 2-52c). Depending on the inclination angle of the mesogens with<br />

the stretching direction, such a transition state would lead to an order parameter<br />

close to zero and could also explain the anomalous behavior of the structural<br />

absorbance [36]. In conclusion, it could be shown that, during elongation of the<br />

anisotropic LSCE network, the network structure changes from a uniaxial symmetry<br />

perpendicular to the applied mechanical field (parallel to the orientation induced<br />

in the second crosslinking step) via a planar-biaxial model (inversion of the order<br />

parameter) to a system aligned in the stretching direction of the rheo-optical experiment<br />

[36]. Simultaneously, the originally transparent polymer becomes opaque<br />

at the inversion point of the order parameter and turns transparent again upon<br />

reorientation of the mesogens into the machine direction.<br />

2.8 Conclusions<br />

The aim of this contribution was to demonstrate the potential of FTIR/FTNIR<br />

polarization spectroscopy at variable temperatures for the characterization of segmental<br />

mobility in liquid crystals and liquid-crystalline polymers under the influence<br />

of external fields. Selected examples have been discussed, where either timeresolved<br />

measurements or static experiments have been utilized to monitor the<br />

effects of electric, electroniagnetic, or mechanical perturbations on the liquidcrystalline<br />

system under investigation. For the extremely fast, reversible switching<br />

processes of low-molecular weight liquid crystals and side-chain liquid-crystalline<br />

polymers in electric fields, the novel step-scan FTIR technique has been applied.<br />

The selective character of vibrational spectroscopy generally allowed us to separate<br />

the orientational effects in different functionalities of the liquid-crystalline compound.<br />

Thus, side-chain liquid-crystalline polymers, for example, have been characterized<br />

under the influence of different external fields in terms of the relative<br />

alignment of their main chain, spacer, and mesogenic groups, respectively. Although<br />

the presented data are necessarily limited to selected examples, the variety of<br />

information and its detailed quantitative Character demonstrate, that FTIR/FTNIR<br />

polarization spectroscopy has certainly become an extremely valuable tool lo elucidate<br />

the dynamics of segmental motion in liquid-crystalline systems.


82 3 Segmental Mobility of Liquid CrVstLiIs und Liqziicl-Cr!).stalline Poljjmer-s<br />

2.9 Acknowledgments<br />

The authors would like to thank Prof. Dr. H. Finkelmann and Dr. J. Kiipfer<br />

(University of Freiburg, Germany), Prof. Dr. R. Zentel and Dr. T. Oge (University<br />

of Wuppertal and University of Mainz, Germany), Prof. Dr. B. Jordanov (University<br />

of Sofia, Bulgaria), Dr. N. C. R. Holm, Dr. S. Hvilsted, Dr. M. Pedersen and<br />

Dr. P. S. Rainanujam (Riso National Laboratory, Denmark), Dr. F. Andruzzi<br />

(CNR, Pisa, Italy), Dr. M. Paci, Dr. E. L. Tassi and Prof. Dr. P.-L. Magagnini<br />

(University of Pisa, Italy), and Dr. E. Happ (University of Ulm, Germany), Dr. C.<br />

Hendaiin (Deutsche Post, Darmstadt), Dr. M. Czarnecki (University of WroclaLy,<br />

Poland) for helpful discussions, individual experimental and theoretical data, and<br />

the supply of polymer samples. Financial and instrumental assistance by the Deutsche<br />

Forschungsgenieinschaft, Bonn, Germany, the Ministerium fur Wisseiischaft<br />

und Forschung, Dusseldorf, Germany, Fonds der Cheinischen Industrie, Frankfurt,<br />

Germany, and the European Community via contract BRE2.CT93.0449 is also<br />

gratefully acknowledged.<br />

2.10 References<br />

[I] Vertogen G., de Jeu, W. H., T/iennotro/iic Liquid Crystcils, F~rthmerzt~d~, Berlin: Springer<br />

Verlag, 1987.<br />

[2] Koswig, H. D., Fli e Kri:r.istalki~, Berlin: VEB Verlag, 1984.<br />

[3] Donald, A. M., Windle. A. H., Lipid Crystallirze Polytners, Cambridge: Cambridge University<br />

Press, 1992, pp. 38-41.<br />

[4] Finkelmann, H., in: Lipid Crj~~rallinirv<br />

Pol,yniers: Prirrcrjhs mid Atntkiriierird Pwprrirs:<br />

Ciferri, A. (Ed.), Weinheim: VCH Verlagsgesellschaft, 1991; pp. 315.<br />

[5] McArdle, C. B., in: Side Chain Liquid Crystal <strong>Polymer</strong>s: McArdle, C. B. (Ed.), Glasgow:<br />

Blackie & Son Ltd., 1989; pp. 357.<br />

[6] Wendorff, J. H., Eich, M.. Mol. Cryst. Liq. Cryst. 1989, 169, 133-166.<br />

[7] Ndkamura, T., Ueno, T., Tani, C., Mol. Cryst. Liq. Cryst. 1989, 169, 167-192.<br />

[8] Demus, D., Goodby, J., Gray, G. W., Spiess, H.-W., Vil, V. (Eds.), H~~rirlDooX o/ Liqid<br />

Crystals, Weinheim: Wiley-VCH, 1998.<br />

[9] Carfagna, C. (Ed.), Lipid Crystalline <strong>Polymer</strong>s, Oxford: Pergamon, 1994.<br />

[lo] Chilton, J. A., Gossey, M. T. (Eds.), Special Po/j3rners ,/ijr Electronic cititl O/~roe/~,c~rro~iic.s,<br />

London: Chapman & Hall, 1995.<br />

[ll] Kelly, S. M., Liq. Cryst. Today 1996, 6(4), 1.<br />

[I?] Schadt, M. in: Lipid Crystal.y: Baunigartel, H., Frank, E. U., Grunsbein, W. (Eds.), Stegenieyer,<br />

H. (Guest Ed), New York: Springer, 1994, Chapter 8, p. 195.<br />

[ 131 Chilton, J. A,, Gossey, M. T. (Eds.), Special Pol~~mers ,for Elecfronics crritl Opto-electrotiics,<br />

London: Chapman & Hall, 1995.<br />

[ 141 Siesler, H. W., Holland-Moritz., K., Zr?/iar.ed crnd Rtrriinrz S{)t~c.fr.u,scopjJ of Poljwiets, New<br />

York: Marcel Dekker, 1980.<br />

[15] Hvilsted, S., Andruzzi, F., Kulinna, C., Siesler H. W., Ramanujam, P. S.? i~~icroriiole~~~rlc..r.<br />

1995,-?8, 2172.<br />

1161 Shilov, S.. Okretic, S., Siesler. H. W., Czarnecki, M. A., Appl. <strong>Spect</strong>rosc. Reus., 1996,<br />

31 (1&2), 125-165.


3. I0 Refcreiice.r 83<br />

[I 71 Hoffinann, LJ., Pfeifer, F., Olirelic, S., Vdlld, N.. Zahedi, M., Siesler. H. W., Appl. Spci~tro.~~~.<br />

1993: 47, 1531-1539.<br />

[l8] Siesler. H. W., in: Oriented Po(iwirr Muterinl.7: Fakirov F. (Ed.), Zug: Hiithig & Wepf, 996.<br />

pp. 138-166.<br />

[ 191 Siesler, H. W., in: Adi1aizce.s in Applied Fourier Transform In/kured <strong>Spect</strong>roscopy: McKenzie,<br />

M. W. (Ed.j, Chichester: Wiley 8: Sons Ltd., 1988; pp. 189-246.<br />

1201 Siesler, H. W.. in: Atloances in Pollvner Science. Berlin: Springer Verlag, 1984; 64. pp. 1- 77.<br />

1211 BrBuchler, M., Boeffel, C., Spiess, H. W., Makroniol. Clic7n. 1991, 192, 1153.<br />

[22] Menzel, H., Hallensleben, M. L., Schmidt, A., Knoll, W., Fischer, T., Stumpe, J. ilfiicrum~leculrs,<br />

1993, 26, 3644-3349.<br />

[231 Buffeteau, T., Natansohn, A,, Rochon, P., Pezolet, M., i~~ucroi7zolecitls~ 1996, 29, 8783.<br />

1241 Graham. J. A.. Hammaker. R. M., Fateley, W. G., in: Cl~errzical. Biological urrd Inchstrial<br />

Appli~~ufims nf’ h~firrr.c~r/ S~~CC/~O.SUJ~,I~: Dnrig, J. R. (Ed.), Chichester: Wiley Interscience,<br />

1985; pp. 301-334.<br />

[25] Souvignier, G., Gerwert, K.. Biopl7~s. J. 1992, 63, 1393.<br />

[26j Uhmann, W., Becker, A,, Taran, C., Siebert, F., Ajyl. Speetro.~c~. 1991, 45, 390.<br />

[27] Palmer. R. A., <strong>Spect</strong>roscopy 1993. 8, 26.<br />

[28] Palmer, R. A,, Chao, J. L., Dittmar R. M., Gregoriou, V. G.? Plunkett, S. E., &p/. <strong>Spect</strong>rosc.<br />

1993, 47, 1297-1310.<br />

[29] Nakano, T., Yokoyama, T., Toriunii, H.. Appl. <strong>Spect</strong>rosc. 1993, 47, 1354-1366.<br />

[301 Gregoriou, V. G., Noda, I., Dowrey, A. E., Marcott, C., Chao, J. J., Palmer, R. A., J. Polym.<br />

Sci. B, 1993, 31, 1769-1777.<br />

[31] Lafrance, C. P.: Nabet, A,, Prud’homnie R. E., Pezolet, M., Can. J. Clwni. 1995, 73, 1497.<br />

[32] Maier, W., Saupe, A.? Z. Nr/turfor.tc/ztiny 1959, Ida, 882.<br />

[33] Maier, W., Saupe, A.. Z. Nriturforschung 1961, 160, 816.<br />

[34] Anderle, K., Birenheide, R., Eich, M., WendorK, J. H., Ilfulcroniol. C//em. Rqid CO~ZIIILII/.<br />

1989, 10, 471.<br />

[35] Natansohn, A., Rochon. P., Gosselin, J., Xie, S., Mclcrori~olec,ri/e,~ 1992, 25, 2268.<br />

[36] Kulinna, C., Dissertotion, Universitat Essen, 1994.<br />

[37] Chatelain, P., Bull. Soc. Fr. Miner. 1943, 66, 105.<br />

[38] Blinov, L. M., Electro-Optical mid ~1~iyrieto-O~)ticcd Proprrtics of Liquid Crysfrtl.r, Chichester:<br />

Wiley-Interscience, 1983.<br />

[391 Frank, F. C., Disctrss. Faradq) Soc. 1958, 25, 19.<br />

[40] Geary, J. M., Goodby, J. W., J. Appl. Phys. 1987, 62, 4100.<br />

[41] Castellano, J. A., Mol. Cryst. Liq. Cryst. 1983, 94, 33.<br />

1421 Jerdme, B. in: Hcindbook of Liquid Crystuls: Demus, D., Goodby, J., Gray, G. W., Spiess.<br />

H.-W., Vil, V. (Eds.), Weinheim: Wiley-VCH, 1998, Vol. 1, p. 535.<br />

[43J Geary, J. M., Goodby, J. M.: Kmetz, A. P., Patel, J. S., J. Appl. Phjis. 1987, 62, 4100.<br />

[44] Toney, M. F., Russel, T. P., Logan, J. A,, Kikuchi, H., Sands, J. M., Kumar, S. K. Nature<br />

1995, 374, 709.<br />

[45] Happ, E., University of Ulm, Sektion <strong>Polymer</strong>e, personal communication, 1994.<br />

[46] Kaneko, E., Principles and Applications of Liquid CrJtstul Displaj~s, Tokio: KTK Scientific<br />

Publ., 1987.<br />

[47] Finkenzellner, U., Spektruni deer. PL’issensclzuft, 1990, 8. 54.<br />

[48] Schadt, M., Helfrich, J., App/, Plzys. Lett. 1971, 18, 127.<br />

[49] Gelhaar, T., Pauluth, D., Ndir. Chern. Tech. Lnh. 1997, 4S( l), 9.<br />

[50] Schadt. M., Phys. Bl. 1996, 7/8, 695.<br />

[51] Kaito, A,, Wang. J. I(.. Hsu, S. L., Anal. Cheni. Acta 1986, 189, 27.<br />

[52] Masutani, K., Yokota, Y., Furukawa, Y., Tasumi, M., Yoshizawa, A., App/. Siiectrosc. 1993,<br />

47, 1370-1375.<br />

[53] Okretic, S., Dissertutiorz, Universitat Essen, 1995.<br />

[54] Shilov, S. V., Okretic. S., Siesler, H. W.. Vib. <strong>Spect</strong>rosc. 1995, 9, 57.<br />

[55] Urano, T., Hamaguchi, H., Clwm Phys. Lett. 1992, 195, 287.<br />

[S6] Urano, T., Hamaguchi, H., Appl. <strong>Spect</strong>rosc. 1993, 47, 2108.<br />

[57] Gregoriou, V. G., Chao, J. L., Toriumi, H., Palmer, R. A.. Chenz. Phys. Lett. 1991, 179,<br />

491.


84 2 Segniental Mobility oj Liquid Crystals and Liqui~~-Cr),st~illine Pol?,nzeus<br />

[58J Czarnecki, M. A,, Katayanm, N., Ozaki, Y., Satoh, M., Yoshio, K., Watanabe, T., Yanagl,<br />

T., Apyl. <strong>Spect</strong>rosc. 1993, 47, 1382-1385.<br />

(591 Powell, J. R.. Krishnan. K., Yokoyamn, T., Nakano, T., in: Proc. 9th 177t. Conference on<br />

Foiwier Trunsjbnn <strong>Spect</strong>roscopy: Bertie, J. E., Wieser, H. (Eds.), Proc. SPIE 2089. 1993; 428-<br />

429.<br />

[60J Noda, I., Appl. <strong>Spect</strong>rosc. 1993, 47, 1329-1336.<br />

[61] Noda, I., -4ypl. <strong>Spect</strong>rosc. 1990, 44, 550-561.<br />

[62] Michl, J., Thulstrup, E. W., <strong>Spect</strong>roscoyj, with Polu d Light. Solute Alignment by Photoselection<br />

in Liquid Crystals, <strong>Polymer</strong>s and Membranes, Weinheim: VCH, 1986.<br />

1631 Gotz, S., Stille, W., Strobl, G., Scheuermann, H., Macronzoleczrles, 1993, 26, 1520.<br />

[64] Hendann, C.. Dissertation, Universitat Essen, Germany, 1997.<br />

[65] Heilnieier, G. H., Zanoni, L. A., Appl. Phys. Lett., 1968, 13, 91.<br />

[66] Bahadur, B., Mol. Cryst. Liq. Cryst. 1983, 99, 345.<br />

1671 Sackmann, E. in Applicntions of' Liqziid Cryytnls, Meier, G.. Sackniann, E., Grabmaier, G.<br />

(eds.j, New York: Springer-Verlag, 1975.<br />

[68] Schumann, C., in Harzdbook of Liquid Crystals, Kelker, H.. Hatz, R. (eds.), Weinheim:<br />

Verlag Chemie, 1980.<br />

[69] Scheffer, T. J., Nehring, J., in P/?.vsics and Chernisfry ofLiquid Crystal Devices, Sprokel, J. G.,<br />

(ed.), New York: Plenum Press, 1979, pp. 173.<br />

[70] Uchida, T., Wada, M., Mol. Cryst. Liq. Cr,yst. 1981, 63, 19.<br />

[71] Bahadur, B. in Liquid Crystuls-Applicutioris and Uses, Bahadur B. (ed.), Singapore: World<br />

Scientific, 1992.<br />

[72] Jordanov, B., Okretic, S. Siesler, H. W., Appl. <strong>Spect</strong>rosc. 1997, 51(3), 447-449.<br />

[73] de Barny, P.. Dubois, J. C., in: Side CI7airt Liquid Crystal <strong>Polymer</strong>s: McArdle, C. B. (Ed.),<br />

Glasgow: Blackie & Son Ltd., 1989; chap. 5.<br />

[74] Lagerwall, S. T. in Handbook of' Liquid Crystals, Demus, D., Goodby, J., Gray, G. W..<br />

Spiess, H.-W., Vill, V. ieds.), Weinheim: Wiley-VCH, Vol. 2B, 1998, 515-664.<br />

[75] Dijon, J., in Liquid Crystals-Applications and Uses, Bahadur B. (ed.), Singapore: World Scientific,<br />

1992.<br />

[76] Goodby, J. W., Blinc, R., Clark, N. A,, Lagerwall, S. T., Osipov, M. A.; Pikiii, S. A,,<br />

Sakurai, T., Yoshino, K., Zeks, B., Ferroelectric Liquid Crj'stuls: Priizc@le,r; Properties und<br />

Applicutions, Philadelphia: Gordon and Breach, 1991.<br />

[77] Kapitza, H.. Poths, H., Zentel, R., Mulironiol. Cheni. Mucron7o1. Syinp., 1991, 44, 117.<br />

[78] Brehmer, M., Zentel, R., Wagenblast, G., Sieniensmeyer, K., Mukroinol. Cheni., 1994, 195,<br />

1891.<br />

[79] Schonfeld, A,, Kremer, F., Hofniann, A,, Kiihnpast, K., Springer, J., Scherowsky, G.,<br />

Makronzol. Clwii. 1993, 194, 1149.<br />

[80] Poths, H., Zentel, R., Liq. Cryst. 1994, 16, 749.<br />

[Sl] Shilov, S. V., Skupin, H., Kremer, F., Gebhard, E., Zentel, R., Liq. Cryt. 1997,22(2), 203-210.<br />

[82] Verma, A. L., Zhao, B., Jiang, S. M., Sheng, J. C., Ozaki, Y., Phys. Rev. E, 1997, 56(3), 342.<br />

(831 Shilov. S. V., Okretic, S.. Siesler, H. W., Zentel, R., Oge, T., Mucron7ol. Rapid Connnun.<br />

1995, 16.<br />

[84] Og, T., Zentel, R., Mabornol. Chem. Phys. 1996, 197, 1805-1813.<br />

[85] Eich, M.! Wendorff, J. H., Reck, B., Ringsdorf, H., Makrornol. Chem. Rupid. Cornnzun. 1987,<br />

8, 59.<br />

[86] Fischer, Th., Lasker, L., Stumpe. J., Kostromin, J. Photocheni. Photobiol. A: Chem. 1994, 80,<br />

453-459.<br />

[87] Wiesner, U., Antonietti, M., Boeffel, C., Spiess, H. W., Macromolecules 1990, 8, 2133.<br />

[88] Stumpe, J., Miiller, L., Kreysig, D., Hauck, G., Koswig, H. D., Ruhmann, R., Rubner, J.,<br />

hfuiirornol. Chew. Rapid. Cornnzun. 1991, 12. 81.<br />

[89] Haitjema, H. J., von Morgen. G. L., Tan, Y. Y.. Challa, G., Macromolecules 1994,27, 6201-<br />

6206.<br />

[90] Fischer, T., Lasker, L.. Czapla, S., Rubner, J., Stumpe, J., ilfol. Cryst. Liq. Cryst. 1997, 298,<br />

2 13-220.<br />

[9 I] Wiesner, U., Reynolds, N., Boeffel, C., Spiess, H. W., hfakrornol. Chern. Rupid. Cornnwn.<br />

1991, 12, 457.


2.10 References 85<br />

[92] Wiesner, U., Reynolds, N., Boefel, C., Spiess, H. W., Liq. Crysr. 1992, 11, 251.<br />

[93] Natansohn, A., Rochon, P., Pezolet, M.. Audet, P., Brown, D., To, S., il/lacromo/ecu/es 1994,<br />

27, 2580.<br />

[94) Buffeteau, T. and PCzolet, M., ,4pp/. S~wctrosc. 1996, 50. 948.<br />

1951 Ramanujam, P. S., Holme, C., Hvilsted, S., Pedersen, M., Andruzzi, F., Paci, M., Tassi, E.<br />

L., Magagnini, P., Hoffmann. U., Zebger, I., and Siesler, H. W., Polyni. Ah. Teclmol. 1996,<br />

7, 768.<br />

[96j Tassi E. L., Paci, M., Magagnini, P. L.. Mol. Cryst. L;q. CFJJS~., 1995, ,736, 135.<br />

[97J Tassi E. L., Paci, M., Magagnini, P. L., Yang, B., Francescangeli, O., Rustichelli, F., Liq.<br />

Cryst. 1998, 24(3), 457-465.<br />

[SS] Holnie, N. C. R., Ramanujam, P. S., Hvilsted, S., Appl. Opt., 1996, 35, 4622.<br />

[99] Zebger, I., Kulinna, C., Siesler, H. W., Andruzzi, F., Pedersen, M., Ramanujam, P. S.,<br />

Hvilsted, S.; Mrrlironiol. Cliern. Murrorid. Symp., 1995, 194, 159.<br />

[IOO] Kulinna, C., Hvilsted, S., Hendann, C.. Siesler, H. W., Ramanujam, P. S., Macr.orno/ecu/es<br />

1998,31,2141-2151.<br />

[I011 Rochon, P., Gosselin, J., Natausohn, A,, Xie, S., App/. Php. Lett. 1992, 60, 4.<br />

[lo?} Kneppe, H., Schnejder F. in Hunrllmok ofLiqitid Crysrds, Demus, D., Goodby, J., Gray, G.<br />

W., Spiess, H.-W., Vill, V. (eds.), Weinheim: Wiley-VCH, Vol. 2A, 1998, 142-169.<br />

[I031 Brand, H. R., Finkelmann, H., in Handbook oj. Liquid Crysluls, Demus, D., Goodby, J.,<br />

Gray, G. W., Spies, H.-W., Vill, V. (eds.), Weinlieim: Wiley-VCH, Vol. 3: 1998, 277-302.<br />

[I041 Onogi, S., Asada, T., in YIZIhlerniitionnl Congwss of Rlieolog.y, Astarita, G., Marrucci, G.,<br />

Nicolais, L. (Eds.), New York: Plenum Press, 1980, Vol. I, 127.<br />

[I051 Schmidt, H. W.? Angeir. Cfiem. Int. Ed. Adti Mateu., 1989, 28, 940.<br />

[lo61 Dunion, M., Zentel, R., Kulinna, Ch., Siesler H. W.; Liquid Crystals, 1995, 18(6j, 903.<br />

[lo71 Hessel, F., Finkelmann, H., Po(vrw7 Bztlletin, 1985, 14, 375.<br />

[IOS] Keller, P., Hardouin, F., Mauzac, M., Achard, M., Mol. Cryst. Lig. Cryst., IZS, 171.<br />

[I091 Finkelmann, H., Kock, H. J., Gleim, W., Rehage, H., Mukromol. Clleni. Rapid Commun.<br />

1984, 5, 287.<br />

[I 101 Finkelniann. H., Kock, H. J., Rehage, H., hfCikIw170/. C/wn. Rapid Cornmiin. 1981, 2, 317.<br />

[I 111 Kupfer, J., Finkelmann, H., Mdmmol. Cliem. Rupid Coninurn. 1991, I?? 717-726.<br />

[I 121 Siesler, H. W., Makroniol. Cllern. Macromol. Svmp. 1992, 53, 89-103.<br />

I1131 Siesler, H. W., n/Iakromot. Chern. Maruomol. Syrr2y. 1991, 52, 113.<br />

[114] Ameri, A., Siesler, H. W., J. Appl. Poljvn. Sci., 1998 (in press).<br />

[115] Wu, P., Siesler, H. W., Dal Maso, F., Zanier, N., Anahr.sis hfagazZi7c~. 1998, 26, 45-48.<br />

[I 161 Knndler, I., Finkelmann, H., Il/lrrcuonzo/. Cheni. Phys., 1998, 199, 677-656.<br />

[1l7] Finkelmann, H., Kupfer, J., Macromol. Chem. Phys. 1994, 1Y5, 1353-1367.<br />

[I 181 Kundler, I., Nishikawa, E., Finkelmann, H., Mucvoniul. S~.nzy., 1997, 117, 11.<br />

[I 191 Bladon, P., Warner, M., Terentjev, E. M., MmronmIearles, 1994, 27, 7067.<br />

[I201 Weilepp, J., Brand, H. R., Etirophvs. Lett., 1996, 34(7), 495.<br />

[121] Kiipfer, J., Finkelmann, H., Po/jwt Adu. Techtzol., 1994, 5, 110-115.


This Page Intentionally Left Blank


t<br />

Order/Disorder in Chain Molecules and<br />

<strong>Polymer</strong>s<br />

G Zerhi mid M. Del Zoppo<br />

3.1 Introduction<br />

At present the theoretical calculation of vibrational properties of polyatomic molecules<br />

can be carried out with reasonable success with the help of large, fast computers<br />

using filly automatic computing programs. Normal mode calculations have<br />

developed along three main lines of thought. On the one hand, a major effort has<br />

been made by the pioneers of vibro-rotational spectroscopy with the aim of determining<br />

accurate ro-vibrational constants to determine harmonic and/or anharinonic<br />

vibrational force fields and the derived physical properties [ 11. Once norinal<br />

modes are known accurately (both fundamental frequencies and relative vibrational<br />

displacements [2, 3]), they become the essential ingredients for the treatment of<br />

vibrational intensities in infrared and scattering cross-sections in Raman [4-61 or<br />

neutron-scattering experiments [7]. A huge effort along these lines has been made<br />

by many authors in the past 50 years, with the pioneering works froin the group of<br />

B.L.C. Crawford at Minnesota and of T. Shimanouchi at Tokyo having been followed<br />

by many authors. The basic mathematics and the computing programs first<br />

derived by Curtiss [S] and then by Schachtschneider and Snyder [9] and freely distributed<br />

worldwide [ 101 have been reproduced with different names and are used<br />

ubiquitously and on most occasions without due mention and thanks to the original<br />

authors for their great efforts.<br />

A second school of thought has actively pursued the idea of calculating vibrational<br />

properties froin ‘ah initio’ quantum chemical calculations [ 1 I]. The underlying<br />

concept is that if optimal atomic wavefunctions (and if large, fast computers)<br />

are available all physical observables can be calculated crb initio, thus making<br />

nonnal-mode calculations from empirical force constants obsolete. A major effort is<br />

currently being made to assess critically the reliability of Lib initio calculations in the<br />

prediction of spectroscopic observables [ 11-14].<br />

A third group of workers has introduced the idea that vibrational force constants<br />

derived from critically determined semiempirical potentials provide an easy way<br />

to calculate vibrational properties, as well as many other structural and dynaniical<br />

molecular properties. These semiempirical potentials are part of fully automated<br />

coinputer packages offered commercially and used widely. For expressing judgment<br />

on the reliability of these semiempirical potentials related to spectroscopic properties,<br />

one matter is to provide a qualitative description of molecular dynamics


88 3 Vibvationul <strong>Spect</strong>ru us a Pvobe of Structurcil 0rdt.vlDisordt.r<br />

(molecules can always be made to ‘wiggle’ on a screen), another matter is to account<br />

quantitatively for frequencies, amplitudes, infrared (dipole derivatives), and Ranian<br />

or hyper-Raman intensities (polarizability derivatives at various orders).<br />

The latter attempt to obtain vibrational force constants is a combination of<br />

the two last methods, i.e., ab initio methods are used to obtain the parameters of<br />

two-body or many-body interactions generally adopted in semiempirical molecular<br />

mechanics calculations [15].<br />

The size of the molecules that can be treated using the above-mentioned methods<br />

is quite large (20-30 atoms) and is not a problem if periodicity [ 16- 181 or symmetry<br />

[3] are intrinsic properties of the molecule. However, the problem becomes more<br />

complex when very large and irregular molecular objects need to be studied.<br />

In this chapter we do not wish to dwell on the problem of the usefulness and<br />

limitations of the various methods for calculating vibrational force constants.<br />

Rather, we will deal with the problems of carrying out normal-mode calculations on<br />

very large molecular systems which have no symmetry are of irregular shape, and<br />

chemically are highly disordered [19]. It appears obvious that none of the abovementioned<br />

methods of calculation can be applied to such huge and structurally<br />

complex systems.<br />

In nature, chemical systems with chemical and/or structural irregularities or disorder<br />

are very common and their relevance in natural phenomena is of fundamental<br />

importance. Biological as well as synthetic oligoiners and polymers form a large<br />

class of compounds whose structure and properties need to be characterized qualitatively<br />

and quantitatively. Vibrational infrared and Rarnan spectroscopy is one of<br />

the few physical tools available to study the microstructure of these systems. Many<br />

empirical assignment and correlative and analytical studies have been carried out,<br />

but the possibility of a theoretical understanding of the vibrational properties (and<br />

the derived infrared and Raman spectra) have not been yet thoroughly explored.<br />

In this chapter we present and discuss the dynamics and vibrational spectra of<br />

long-chain molecules with disordered structure, and discuss the current possibilities<br />

of calculating the vibrational infrared, Raman and neutron-scattering spectra of<br />

these systems.<br />

3.2 The Dynamical Case of Small and Symmetric<br />

Molecules<br />

Let us briefly define the elementary parameters and review the basic equations for<br />

the treatment of molecular and lattice dynamics. The concepts presented here form<br />

the background for a further discussion later in the chapter for the case of large<br />

molecular systems.<br />

In classical molecular spectroscopy, a vibrating molecule is always represented by<br />

a set of balls connected by weightless springs. The springs represent the chemical


onds which hold atoms together through directional forces generated by the interactions<br />

of valence electrons. Since the vibrational molecular potential is unknown it<br />

is approximated by a Taylor expansion about the equilibrium geometry in terms of<br />

a set of suitably chosen coordinates. Let x,, y, and z, be the instantaneous Cartesian<br />

coordinates of the a-th atom of the vibrating molecule and let x,", y: and z," be the<br />

corresponding equilibrium coordinates when the molecule is at rest. For simplicity<br />

let x, (i = 1-3N) label the vibrational displacements of the type xu - x:, yr - y: and<br />

z, - z," for the N atoms. The generally unknown vibrational potential can be<br />

approximated by the following Taylor expansion about the equilibrium structure<br />

2V = 2Vo + 2 C (CJV/~X~)~X~<br />

1 'J<br />

+ C (8sV/8~,d~J)o~,~,<br />

+ . . . (3-1)<br />

The zero-th order term is removed by a suitable shift of the origin, the first-order<br />

term is zero because the forces (dV/ax,),, vanish at equilibrium, the second-order<br />

term defines a quadratic harmonic potential. In most cases of chemical interest the<br />

expansion is truncated at the second order. The analysis of the higher-order terms is<br />

outside the scope of our discussion. The truncation at the second order implies that<br />

the restoring forces are assumed to be linear with the infinitessimal displacements<br />

from the equilibrium position.<br />

The quadratic potential can be written in matrix notation as<br />

2V = x'Fxx (3-2)<br />

where x is the vector (and x' its transpose) of the 3N Cartesian displacement coordinates<br />

and F, is the matrix of the quadratic force constants:<br />

Correspondingly, the kinetic energy can be written as:<br />

2T = X'MX (3-4)<br />

where M is the diagonal matrix of the atomic masses. Only 3N - 6 of the xi are<br />

independent because the Cartesian coordinates are related by the Eckart-Sayvetz<br />

conditions [3].<br />

With such a model, atoms are allowed to perform very small harmonic oscillations<br />

about their equilibrium positions and the dynamical treatment describes the<br />

normal modes Q, of a set of coupled harmonic oscillators.<br />

When the kinetic and potential energies are introduced into the Lagrange equation<br />

the solutions are of the form:<br />

where A, are called frequency parameters and define the frequency of oscillation of


90 3 Vihrntimzal <strong>Spect</strong>ra as n Probe qf Strzrcturol Or-cler<br />

the atoms during the j-tli normal mode Q,. During each Q, all atoms move in phase<br />

(i.e., they go through their equilibrium positions at the same time) with frequency v,<br />

(in cm-') = (Eb,/4n-c-)<br />

7 7 112 and with amplitudes L,.<br />

Frequency parameters and vibrational amplitudes can be calculated by solving<br />

the secular equation<br />

M-'F,L, = L,A (3-6)<br />

Six of the 2, vanish because of the Eckart-Sayvetz conditions and describe three<br />

rigid translations and three rigid rotations of the molecule; each of the remaining<br />

3N - 6 nonvanishing A, is associated to the normal mode Q,. Depending on the<br />

shape (symmetry) of the molecule, degenerate symmetry species may occur and<br />

some of the nonvanishing may turn out to be equal (doublets, triplets, and even<br />

multiplets with very high symmetrical molecules, e.g., fullerene). Accidental degeneracy<br />

may occur for large and asymmetric molecules when intramolecular coupling<br />

is small or zero. When all 3N solutions are considered together, the relation between<br />

Cartesian displacements and normal modes is as follows:<br />

x = L,Q (3-7)<br />

The matrix L, is nonorthogonal because M-'F, is not symmetric. The normalization<br />

conditions are<br />

L,L', = M-' (3-5)<br />

which ensures that<br />

2T = Q'Q (3-921,)<br />

2V = Q'AQ<br />

(3-9b)<br />

i.e., normal modes are mutually independent and can be described as isolated<br />

harmonic oscillators. The shape of the vibrational motions sought in any work of<br />

vibrational assignment is described precisely by the solution of Eq. (3-6), once the<br />

intramolecular potential and the geometry of the molecule is suitably chosen.<br />

A few observations on the treatment of molecular and lattice dynamics in cartesian<br />

coordinates are necessary.<br />

1. When molecular dynamics is applied to the understanding of chemical intramolecular<br />

phenomena the use of chemical internal displacement coordinates as<br />

initially defined by Wilson et al. [3] are much more useful and may have a direct<br />

chemical meaning (see below).<br />

2. When molecular vibrations are studied by quantum chemical or molecular mechanics<br />

methods, the problem is first treated in Cartesian coordinates and possibly<br />

later transformed into internal coordinates. These programs generally<br />

provide the numerical values of each element of the F, and L, matrices; the


3.2 The Dyriamical Case qf Simll arid Symnietric Molecziles 91<br />

elements of Lx are simply the numbers needed to directly pro-iect on a screen the<br />

atomic vibrations for each Q,,.<br />

3. The treatment of vibrational infrared and Raman intensities provides quantities<br />

such as (7p/dx, and d((a))/ax; (,where p is the total molecular dipole electric<br />

dipole moment and ((a)) is the total niokcular polarizability tensor). These<br />

quantities have a direct physical meaning and are directly calculated ah iizitio or<br />

sometimes by extended programs of molecular mechanics calculations [4-61.<br />

4. When lattice dynamic calculations for one- or three-dimensional systems need<br />

to be carried out, the Cartesian space is more suitable in order to describe both<br />

inter- and intra-molecular vibrations [20].<br />

When the chemical reality of an isolated molecule needs to be considered through<br />

its vibrational spectrum, new sets of ‘chemical’ coordinates can be defined [3]. Let<br />

R = BX (3-10)<br />

define a set of internal displacement coordinates where B is the matrix of geometrical<br />

coefficients and<br />

where<br />

xvib 1 AR (3-1 1)<br />

For a discussion of the A matrix, see [21]. For simplicity in this discussion<br />

we assume that a nonredundant set of 3N - 6 Rs has been defined [22]. The Rs<br />

describe bond stretching, valence angle bendings, and torsions. The vibrational<br />

potential can be rewritten in terms of Rs as<br />

2V = R’FRR (3-13)<br />

and the kinetic energy matrix as<br />

where<br />

2T = R’(GR)-’R (3-14)<br />

GR = BM-~B’ (3-15)<br />

G contains all information on equilibrium geometry and masses. The potential<br />

energy matrix is written as<br />

FR = A‘F,A (3-16)<br />

Equation (3-16) is the one to be used when F, is calculated by (16 iizitio methods and


92 3 J,’ibrationrrl <strong>Spect</strong>ra as a Probe of Structural Order<br />

the problem requires a treatment in terms of internal displacement coordinates R.<br />

Care must be taken in the construction of A [21].<br />

The eigenvalue equation in internal coordinates becomes<br />

and the relation to normal coordinates is given by<br />

(3-17)<br />

R = LRQ<br />

(3-18)<br />

with the iiornialization condition<br />

LL’ = GR (3-19)<br />

As discussed in Section 3.4, many norinal modes are characterized by the fact<br />

that only a few atoms belonging to a specific chemical functional group are moving,<br />

while the other do not feel the intramolecular geometrical and electronic changes<br />

which occur at a specific site of the molecule. Such group uihrutiorzs are then characteristic<br />

of a given functional group and allow its identification in a chemically<br />

unknown material. The existence of ‘localized’ group vibrations showing specific<br />

and characteristic group frequencies have fueled the development of spectroscopic<br />

correlations which have made infrared (and recently also Raman) spectra a valuable<br />

physical tool for chemical and structural diagnosis (23-261. These correlations<br />

form the basis of many automated computing programs for the chemical identification<br />

of unknown compounds.<br />

For the above reason, researchers felt the need to describe normal modes in terms<br />

of ‘group coordinates’ (or local symmetry coordinates) which can be defined by an<br />

orthogonal transformation<br />

X=CR (3-20)<br />

The expressions for Gx and Fx and for the corresponding secular equations are<br />

the results of a straightforward similarity transformation such as in Eqs. (3-13) and<br />

(3-14). The Japanese School of Shimanouchi has provided many force fields and<br />

vibrational assignment in terms of group coordinates [27].<br />

When all redundancies between internal coordinates are removed, the size of the<br />

eigenvalue equation to be solved is equal to the number of normal modes (3N- 6)<br />

expected for the n~olecule under study. If the molecule has some symmetry it<br />

belongs to a given symmetry point group g; group theory provides the structure<br />

of the irreducible representation of g, i.e., the number of normal modes in each<br />

symmetry species r,. By a suitable linear and orthogonal transformation<br />

S=UR 01 (S=UC‘) (3-21)<br />

(with U’ = U-’) it is possible to define a new set of symmetry coordinates S which<br />

foiin the basis of an irreducible representation of the point group g 131.


3.2 The Dyrinniical Case of Sinnll and Sjwimetric Molecules 93<br />

The scalar quantities V and T in Eqs. (3-13) and (3-14) can be re-expressed in<br />

terms of symmetry coordinates<br />

2V = S’UFRU’S = S’FsS (3-22)<br />

2T = S’U(G,)p’U’S = S’(Gs)p’S (3-23)<br />

The new matrices Fs and Gs are factored into blocks with dimensions identical<br />

to the structure of the irreducible representation [3]. The corresponding secular<br />

equation becomes separated into blocks of smaller size. Numerical calculations<br />

carried out separately for each symmetry block enable to identify the normal modes<br />

which belong to a given symmetry species. This is essential when the vibrational<br />

assignment needs to be carried out [2]. Band shape analysis (in IR and Raman)<br />

of gaseous samples, Raman depolarization ratios (for gaseous or liquid/solution<br />

samples) and dichroic ratios in IR or Raman experiments in polarized light on<br />

single crystals or stretch-oriented solid samples [28] provide the experimental data<br />

supporting group theoretical predictions.<br />

The numerical solution of the eigenvalue Eq. (3-17) is obviously greatly simplified<br />

when symmetry factoring can be carried out for molecules which possess some<br />

symmetry elements. Symmetry factoring was essential in the past (even for small<br />

molecules) when numerical processes were necessarily cumbersome and time consuming<br />

because of primitive computing technologies. In spite of the explosive development<br />

of computational tools this problem cannot yet be fully overlooked. The<br />

molecular systems to be treated have become large and the size of the secular determinant<br />

can still pose some problems.<br />

A typical example for a modern textbook is the case of the molecule of fullerene<br />

(C~O) which shows a very high symmetry (point group Ih): its 174 normal vibrations<br />

are separated into smaller symmetry blocks. Since icosahedral symmetry gives rise<br />

to a large number of degenerate modes, only 46 distinct mode frequencies are<br />

expected:<br />

The application of group theory for the determination of the structure of the<br />

irreducible representation (for the prediction of optical selection rules for infrared<br />

and Ranian spectra) is generally easy and straightforward [2, 31. On the contrary<br />

the construction of symmetry coordinates (i.e., the construction of the U matrix,<br />

Eq. (3-21)) is sometimes diflicult and cumbersome for large systems especially when<br />

degenerate species occur (e.g., adaniantane and fdlerene). An automatic method<br />

for the construction of symmetry coordinates using computers based on the diagonalization<br />

of the GR matrix has been proposed [29]. The method is based on the<br />

fact that GR contains in itself all the information on the symmetry of the problem<br />

and on the redundancies, if some or all have not been previously removed. The<br />

diagonalization of GR provides directly the elements ofthe U matrix (Eq. 13-21)) as


94 3 Vibrational <strong>Spect</strong>ro us a Probe oj Strzrctural Order<br />

follows. Let GR be diagonalized by the unitary transformation D<br />

GRD = Dl- (3-23)<br />

where r is the diagonal matrix of the eigenvalues. The new set of coordinates<br />

C = D’R (3-25)<br />

forms an irreducible representation of the symmetry point group of the molecule,<br />

thus the C, form a set of symmetry coordinates and D’ can replace U in Eq. (3-71).<br />

The method is particularly useful for the automatic removal of all redundancies<br />

in structurally complex systems. Indeed if in > 3N - 6 is the size of the eigenvalue<br />

equation<br />

r=m-(3N-6) (3-26)<br />

is the number of zero eigenvalues in Eq. (3-34) which provide the required r<br />

redundancy relations<br />

D’~R = o (3-27)<br />

between the coordinates used in setting up the starting B matrix (Eq. (3-10)).<br />

A fully automated computer method for the handling of group theory has been<br />

proposed and is recommended for extremely large and highly symmetrical systems<br />

with many local and cyclic redundancies [30]. The numerical methods in treating<br />

group theory and symmetry coordinates have been generally neglected, as few large<br />

and highly syininetrical molecules have required the attention of spectroscopists. At<br />

present, the need to understand the vibrations of fullerene, fulleroid, tubular molecules,<br />

dendrites, etc., may revive the use of such numerical methods.<br />

3.3 How to Describe the Vibrations of a Molecule<br />

The main purpose of a dynaniical analysis of a molecule is to assign the vibrational<br />

transitions observed in infrared and Rainan in terms of molecular motions. As<br />

mentioned above, chemical group frequency correlations have been the justification<br />

of most of the vibrational assignment. The description of the normal modes has<br />

posed some problems throughout the years.<br />

It would seem obvious that the only way to describe the atomic motion would be<br />

that of plotting the Cartesian displacements of each atom during normal mode Q,<br />

with frequency vJ. At present, with the availability of computer techniques and fully<br />

automated computing programs, the Cartesian atomic displacements are a norinal<br />

part of the output, while some programs even provide animated descriptions of the<br />

molecular vibrations which can be viewed in all directions by appropriate use of


3.4 S11or.f- and Loizg-Rcrizye Vibrational Coupliiig in Molmrles 95<br />

the software. Using Cartesian displacement coordinates, the identification of group<br />

frequencies is not always straightforward and unambiguous. Moreover, the publication<br />

of many drawings or of many tables of numbers becomes often graphical and<br />

editorial problems.<br />

In the past, in order to describe the normal modes much use has been made of the<br />

so-called potential energy distribution (PEDI. Torkington [3 I] has first proposed<br />

and Moriiio aiid Kuchitsu [32J have later generalized the concept of PED defined<br />

as:<br />

PED = APJF (3-28)<br />

where J,,,, = Li, and J,,ml = 2LmlLll.<br />

The advantage of PED is that it can be expressed either in internal or group<br />

coordinates, thus allowing a more direct description of the main characteristic of<br />

the nomal modes.<br />

In simple words, PED provides quantitatively (generally PED is expressed as 'XI<br />

contributions) the extent of involvement of one or several diagonal and off-diagonal<br />

force constants in a given normal mode. The classical textbook case is that of the<br />

well-known group frequency modes of a CH2 group. Using group coordinates (Eq.<br />

3-20)), the modes commonly described as CH2 antisymmetric stretch (d-), CH2<br />

symmetric stretch (d+), CH2 scissoring (a), wagging (w), twisting (t), and rocking<br />

(P) can be easily identified with reference to the corresponding Fx group force<br />

constants [27]. If the same motions were expressed in tenns of classical internal<br />

coordinates of CCH bending the PED would indicate unselectively that in all the<br />

four motions CCH bendings are involved without allowing any distinction of the<br />

characteristic group modes. The same situation is found in the case of the characteristic<br />

group modes of the -CH3 group which are recognized by their characteristic<br />

frequencies of bending, umbrella (U), in-plane, and out-of-plane rocking.<br />

A typical case, famous in organic vibrational spectroscopy, is represented by the<br />

work on trans-planar n-alkanes by Schachtschneier and Snyder [33] aiid by Snyder<br />

et a/. [34]; the latter has extended his study to branched paraffins [35], to a-alkanes<br />

in the liquid phase [36] and to oligo and poly-ethers [37]. The use of PED has<br />

allowed to distinguish clearly between band progressions and the evolution of the<br />

normal modes by changing the phase-coupling (see later in this chapter).<br />

3.4 Short- and Long-Range Vibrational Coupling in<br />

Molecules<br />

As in this chapter the aim is to understand tlie vibrations of large and highly disordered<br />

and structurally irregular molecules, tlie basic problem to be faced is<br />

whether and to what extent normal vibrations are able to probe the molecular<br />

intrainolecular environment, i.e., whether normal vibrations are mostly localized or<br />

extended over a large portion of the molecular system.


96 3 Vihratioiial <strong>Spect</strong>ru as a Probe qf Structural Order<br />

This issue has rarely been faced by molecular spectroscopists for several reasons.<br />

All those who searched for chemically useful characteristic frequencies aimed at<br />

modes strongly localized within the functional group [23-261. In contrast, those<br />

who dealt with large systems generally considered oligomeric or polymeric molecules<br />

as structurally perfect systems, often with translational periodicity, and<br />

studied 'phonons' which, by definition, are collective phenomena [ 17-1 91. The<br />

problem cannot be overlooked when the vibrations of large and structurally complex<br />

molecules must be understood.<br />

The first theoretical approach seeded by the explosive development of chemical<br />

spectroscopic group frequency correlations has been presented by King and Crawford<br />

who gave the dynamical conditions under which a normal mode can be defined<br />

as 'localized' [38].<br />

111 correlative vibrational spectroscopy, a group of chemically similar molecules<br />

shows characteristic group frequencies [23-261 vi (say, the stretching of the carbonyl<br />

group >C=O, very strong in infrared) which exhibit systematic changes that chemical<br />

spectroscopists would like to ascribe to inductive and/or mesomeric effects by<br />

the various substituents placed in the molecules either at short or large distances<br />

from the functional group of interest. On the other hand, the observed wavenumber<br />

shifts may also originate from changes of mass and/or geometry.<br />

Use is made here [38] of the dynamical quantities presented in Chapter 2. Let<br />

Go, Fo and La-' be the kinetic, potential, and normal coordinate transformation<br />

matrices of one molecule of the series taken as reference [39]. In going from one<br />

molecule to a chemically similar one within the same class, one may expect small<br />

changes in the geometry (AG) or in the force constants (AF), or in both, which may<br />

be the cause of the observed changes of vi. The corresponding matrices of the<br />

molecule so modified can be written:<br />

G=Go+AG<br />

F = Fo + AF<br />

L-I = L~-' + A(L-')<br />

(3-29)<br />

(3-30)<br />

(3-31)<br />

If the perturbations of G and F matrices are small it can be assumed that the<br />

eigenvectors do not change and the zeroth-order values in Eq. (3-31) are used.<br />

When the eigenvalue Eq. (3-17) is expanded in terms of the quantities in equations<br />

(3.29) through (3-31) in terms of the first-order changes AG and AF and the zerothorder<br />

eigenvectors, the final expression for the (group) frequency parameter turns<br />

out to be:<br />

in which (lo), is the frequency parameter of the i-th group-frequency mode of the<br />

reference unperturbed molecule, AG,1 and AFml are the changes of those elements<br />

of the kinetic and potential energy matrix whose contributions to the changes of A,<br />

are scaled by the related elements of the Lo and L,' matrices.


3.4 Slzort- and Loiig-Range Vibrational Couyliiig in Molecules 97<br />

According to King and Crawford [38], if li has to be insensitive to changes of the<br />

molecular framework (ix., if the motion is to be localized) the sufficient conditions<br />

are that the coupling terms in Eq. (3-32) be small, namely either<br />

or<br />

(3-33)<br />

(3-34)<br />

AGml z 0<br />

AFm, E 0<br />

(3-35)<br />

(3-36)<br />

It is interesting to note that the problem of dynamical localization versus collectivity<br />

has never been a key issue in molecular spectroscopy and dynamics in the past<br />

40 years as the molecules considered were of limited size, thus obviously implying<br />

the existence of collective motions throughout the whole molecule. Moreover, most<br />

of the molecules considered contained either only covalent CJ bonds or, if n bonds<br />

existed, they were either isolated or with limited conjugation (e.g., aromatic molecules).<br />

The recent explosive interest in highly conjugated polyene and polyaromatic<br />

systems [40] has asked molecular dynamics to reconsider the problem of the extent<br />

of conjugation (delocalization) of n: bonds aloiig the whole molecular chain 141, 421.<br />

At the present stage of the spectroscopic studies for organic systems, a few<br />

observations on Eqs. (3-32) to (3-36) must be made.<br />

1. From the definition of the GR matrix (Eq. (3-15)) or, more directly, from the<br />

analytical expressions of the GR matrix elements provided by Wilson et al. [3], it<br />

transpires that the element (GR)~, connecting internal coordinates i and j may be<br />

different from zero when they possess at least one atom in common. It follows<br />

that atoms away from the functional group under study certainly will not affect<br />

the group frequency; Eq. (3-35) holds, i.e., the G matrix favours localization.<br />

2. For molecules made up exclusively by CJ bonds or containing also isolated and<br />

noncoiijugated K bonds inductive effects are generally considered not to extend<br />

at large distances, thus limiting the distances of the possible electronic effects<br />

described by AFllli to a few bonds around the ftinctional group (hence to a few<br />

internal coordinates). Equation (3-36) holds and again localization seems to be<br />

favored when strong electronic delocalization over large distances is not likely to<br />

occur within the molecule.<br />

3. When intramolecular mechanical coupling is very large (i.e,, Eqs. (3-33) and (3-<br />

34) do not hold) mechanical delocalization over a certain molecular domain<br />

takes place and the concept of group frequency is lost. This generally happens<br />

in the energy range between z 1500-400 cm-' where skeletal stretching and<br />

bending motions are strongly coupled. In this case small mass, geometry. or<br />

electronic perturbations may induce large changes in the spectrum, as indeed<br />

observed experimentally.


98 3 I i’brational <strong>Spect</strong>ra as a Probe of Structural Ordrv<br />

4. The situation represented by Eq. (3-36) is at present of particular interest when<br />

polyconjugated molecules (polyenes and polyaromatic systems) are considered.<br />

The role played by vibrational infrared and Ranian spectroscopy in this new<br />

field of material science has been essential for the understanding of the new<br />

physical and chemical phenomena associated with the existence of extended n<br />

electron delocalization. The reader is referred to specialized references for a<br />

thorough discussion on the spectroscopy of these systems [4143]. For these<br />

systems recent studies have pointed out the occurrence of characteristic normal<br />

modes which show ‘frequency dispersion’ with conjugation length, i.e., the<br />

characteristic fi-equencies lower by adding repeating units which tale part in the<br />

conjugation [41-431. Dealing with a chain molecule, the elements (Lo),m, and<br />

entering Eqs. (3-33) and (3-34) are necessarily large. The physically<br />

relevant problem is to discover whether, to what extent, and at which distance<br />

the elements of AF change by adding conjugated units. The solution of this<br />

problem would shine some light on the distance and the extent of electronic<br />

interactions in polyconjugated systems 1441. This is a challenge to spectroscopists,<br />

theoretical quantum chemists, and computational chemists.<br />

5. It is obvious that the addition of strong electron donors or acceptors at a site<br />

away from the functional group of interest will modify the electronic environment<br />

and possibly the geometry at the site of the substitution. Certainly for<br />

particular internal coordinates, large AF and/or AG must occur, but their contribution<br />

to the frequency shift of the group frequency of interest is reduced to<br />

zero since the corresponding elements of L or L-‘ which relate the motion of<br />

the functional group and the motions at the site of the substituent are all zero<br />

(Eq. (3-33)).<br />

3.5 Towards Larger Molecules: From Oligomers to<br />

<strong>Polymer</strong>s<br />

Molecular spectroscopy and lattice dynamics of oligonieric and polymeric materials<br />

have been treated thoroughly in several textbooks or articles [16-181. We wish to<br />

point out here a few fundamental concepts which form the basis of the understanding<br />

of the spectra of disordered materials which will follow in this discussion.<br />

Let LIS first consider the vibrations of polymers considered as one-dimensional<br />

perfect and infinite lattices. The usual basic assumptions are the following:<br />

1. The polymer is obtained by forming a one-dimensional chain of chemical units<br />

linked with a pre-assigned cherwicnl seqzirnce. For sake of example, we consider<br />

propylene (CH~-CH=CHZ) as the prototype of a monomer unit capable of<br />

producing a polypropylene chain. The first step is to make a chain of polypropylene<br />

in which all monomers are chemically linked head-to-tail (cheiizicul


eyularity), e.g.,<br />

2. During the polymerization process the catalysts is chosen in such a way that<br />

.stereo-rrgzilur.itJi is assured [45], e.g.,<br />

isotactic: CH3 CH3<br />

I<br />

-C-CH?-C-CH2<br />

I<br />

H<br />

I<br />

I<br />

H<br />

3. Upon solidification from solution (or from the melt), inacromolecules coil in an<br />

energetically preferred minimum energy conformation and form long chains<br />

along which roriforniatioiznl periodicity is assured [46-48). Chains are considered<br />

long enough to be considered ‘infinite’. For isotactic polypropylene, the conforniational<br />

sequence is TGTGTG. The overall shape of the chain can be described<br />

as a regular helix with a threefold screw axis [45].<br />

4. The process of solification proceeds with the formation (when possible) of crystalline<br />

material in which chains pack in a lattice with tri-diinensional periodicity.<br />

The conditions for perfect packing are that chains possess chemical regularity,<br />

configurational periodicity, and conforinational regularity.<br />

Since interniolecular Van der Waals-type forces which hold molecules in a crystalline<br />

lattice are very weak, when compared with intramolecular covalent forces,<br />

lattice dynamics of polymers is made easy by considering firstly the vibrations of<br />

infinite isolated polymer chains [ 16- 181. Three-dimensional order is considered<br />

later as a sinall perturbation of the normal modes of the one-dimensional chain<br />

by the weak intermolecular forces [49]. In other words, we are considering the<br />

dynamics of one chain ‘in ziacuo’.<br />

The structural characteristics of the polymer chain constructed as described<br />

above are that translational or roto-translational symmetry can be found along the<br />

chain in addition to the traditional point group symmetry operations (rotation axes,<br />

symmetry planes, inversion center, and identity) [18].<br />

In a dynamical treatment the labeling of the internal coordinates starts froin the<br />

reference repeat unit and proceeds along the chain by applying the one-dimensional<br />

translational operator which identifies the translationally equivalent internal coordinates.<br />

The structural situation is clarified if one considers that a polymer chain<br />

can be generated by application to a chemical repeat unit of a certain rototranslational<br />

operator $(Q, 1) which corresponds to a rotation of an angle 8 about the<br />

chain axis and a translation of a fraction 1 = d/n of the crystallographic unit length<br />

(repeat distance) d. In general, the crystallographic repeat unit can contain L<br />

chemical units coiled in a helix with t turns [47].


100 3 T%xitionul <strong>Spect</strong>in as a Pvohe of Straictuval Order.<br />

Let p be the number of atoms in the chemical repeat units and let the indices n<br />

and n’ label the sites of the chemical repeat units along the polymer chain and i and<br />

1 label the 3p internal co-ordinates within the repeat unit. The potential energy can<br />

be formally written in a way similar to Eq. (3-13) [SO].<br />

11; n‘<br />

i. 1<br />

(3-37)<br />

where<br />

(FR)llll’i, = (FR)ni”li (3-39)<br />

let s : In - 11‘1 define the distance of interaction, which (see Section 3.6’1 is very<br />

important in determining the vibrations of the chain; it follows that because of<br />

translational symmetry<br />

Substituting Eq. (3-40) into Eq. (3-37), one obtains<br />

In Eq. (3-41), the first term gives the contribution of the intramolecular potential<br />

by the reference unit at the n-th site, and the second and third terms collect the<br />

contribution of intramolecular coupling of the n-th unit with the neighboring units<br />

at distances s and -s respectively.<br />

A similar equation can be written for the kinetic energy in a way similar to what<br />

has been done for the finite molecule (Eq. (3-14)). Since the number of oscillators<br />

is taken to be infinite (the chain is considered infinite) the dynamical problem results<br />

in the writing of an infinite number of second-order differential equations, the<br />

solutions of which are of the type<br />

R”+S.<br />

, - A, exp 1-i (At + sv,)] (3-42)<br />

It should be noted that A, is independent of n, v, is the phase shift between two<br />

adjacent roto-translationally equivalent internal co-ordinates, and A refers to the<br />

vibrational frequency as in the case of small molecules (see Eq. (3-5)). Physically,<br />

Eq. (3-42) tells us that the j-th coordinate at the n + s site oscillates with frequency /?<br />

with an amplitude equal to the amplitude of the j-tli coordinate in the reference cell<br />

at site n multiplied by a phase factor exp (-isq). Attention should be paid to the


3.5 Towards Lnrger Molecules: From Oligoiiiers to <strong>Polymer</strong>s 101<br />

fact that y is the phase shift of two equivalent oscillators belonging to two adjacent<br />

chemical units within the crystallographic repeat unit (see above).<br />

Following the same algebraic procedures as for the finite molecule the final<br />

secular equation to be solved is [50]<br />

where<br />

GR(~) = (GR)’ + c [(GR)S eisq + (GR’)S ePisV‘] (3-44)<br />

S<br />

FR(~) = (FR)’ + c [(FR)~ eisv + (FR’)~ eCiSv]<br />

S<br />

(3-45)<br />

Equation (3-43) is of 3pth degree in A. There are 3p characteristic roots (i = 1 . . .3p)<br />

for each value of the phase y. The r = 3p functions i~,(v,) can be interpreted as<br />

branches of a multiple valued function which is the dispersion relation for a onedimensional<br />

polymer chain. The solution reached in Eq. (3-43) is analogous to that<br />

previously attained by many authors in solid-state physics for simple tridimensional<br />

lattices [51]. The important digerelice is that Eq. (3-43) treats the problem in internal<br />

chemical coordinates, while, generally, simple lattices are treated in Cartesian<br />

coordinates (Eq. (3-6)). The fhction is periodic with period ~ Tand C calculations are<br />

limited to values of q~ within the first Brillouin zone [51, 521 (BZ) (-TC I y i TC). In<br />

most cases, the results for only half of the BZ are reported, the second half being<br />

symmetrical through y z 0.<br />

In the spectroscopy of polymers as one-dimensional ID lattices it is generally<br />

easier to deal with the phase shift y at odds with solid-state physics, dealing with<br />

three-dimensional 3D crystal, which treats the whole dynamics in terms of the<br />

vector k. For ID lattices we can describe the wave-motion (phonon) propagating<br />

along the 1D chain either by the phase shift v, or by the vector Ik/ = y/d where k<br />

has only one component along the chain axis k,. In contrast, in the spectroscopy<br />

and dynamics of 3D lattices (where intermolecular forces are active in all directions)<br />

phonons are labeled with k vectors with three components (kk, k,, k,) [51]. When k<br />

is used in polymer dynamics Eq. (3-43) can be easily rewritten as<br />

It should be remembered that in a single and isolated polymer, chain atoms<br />

are allowed to move in the tridimensional space even if phonons are considered<br />

to propagate in one direction along the chain axis. This means that no neighboring<br />

intermolecular interactions are taken into account, i.e., the dynamics is that of a<br />

chain ‘in uacuo’.


102 3 l,i'brationcil <strong>Spect</strong>ra as n Probe of Structural Order<br />

300G i<br />

-<br />

2Got<br />

3000<br />

r<br />

7<br />

250{ -<br />

1500<br />

l 5 O 0 L =<br />

loook<br />

Figure 3-1. Dispersion curves of single-chain trans planar polyethylene. (a) v/p; (b) i$/k.<br />

As an example, in Figure 3-1 we report the dispersion curves of a single chain of<br />

polyethylene in v, and k space. The chain of trims-planar polyethylene -(CW?),,- is<br />

generated by application to a single CH2 group of the operator $(n. d/3). It follows<br />

that the crystallographic repeat unit contains two CH? groups. The physics and the<br />

numbers are not changed. The dispersion curves in Figure 3-lb are obtained by just<br />

halving Figure 3-la at .n/2 [53].<br />

As an example of a more complex polymer chain we quote the case of polytetrafluoi-oethylene<br />

-(CFZ)n- for which one can identify the rototranslational<br />

operator $(2n/15, d/15) which describes a polymer chain made up by z = 15 CF2<br />

units coiled in a helix with t = 7 turns. In cases such as polytetrafluoroetliylene the


3.5 TOIIYW~ Laiyer Molertiles: Fvoiii Oligomevs to Polimers 103<br />

-<br />

Figure 3-2. Dispersion curves \./a, for single chain helical<br />

(z = IS, t = 7) polytetrafluoroethylene.<br />

0' 60' 120" lBi<br />

plot of v(q) becomes usable while v(k) becomes unreadable (117 frequency phonon<br />

branches in k space) (Figure 3-2) [54].<br />

Dealing with 1D chains iii vuczio carries an important consequence on the shape<br />

of the dispersion curves if compared with the same systems considered as 3D<br />

lattices. From lattice dynamics in 3D it is known that if q is the number of atoms<br />

in the crystallographic repeat unit the dispersion relation provides 3q phonon<br />

branches. three of which are labeled as 'acoustical' and 3q - 3 as 'optical' branches.<br />

The former three have zero frequencies at the so-called r symmetry point in k space<br />

(i.e., with k, = k, = k, = 0; see Figure 3-3) and correspond to the three nongenuine<br />

vibrations with zero frequencies which describe the three rigid translations of the<br />

whole lattice along the three Cartesian coordinates [49. 551. The 3q - 3 optical<br />

branches may interact with the electromagnetic waves according to specific selection<br />

rules and provide the system with particular optical properties (absorption and/<br />

or Raman scattering), as discussed below.<br />

In ID systems iii u~~cuo the rigid rotation of the whole chain about its axis is not<br />

hindered. It follows that an additional zero eigenvalue is expected from Eq. (3.43)<br />

when applied to 1D systems at k = 0. The lack of intermolecular forces for the<br />

polymer chain in uact[o carries an important consequence on the shape of the two<br />

acoustic branches near k = 0. While the three acoustic branches for 3D crystals<br />

have a discontinuity at k = 0, for ID crystals two of the acoustical branches may<br />

approach ~(0) almost asymptotically [56]. It follows that the dynamics of 1D systems<br />

based on the model of an isolated chain at small k values and at very low<br />

frequencies is totally wrong and cannot be used for the interpretation of related<br />

physical properties of real systems (neutron-scattering experiments, thermodynamic<br />

properties, Debye Waller factors, mean square amplitudes, and other physical<br />

properties where thermal population is involved which is strongly affected by lowenergy<br />

vibrations).<br />

For the benefit of the discussion which follows we give here a calculated schematic<br />

example of the procedures followed in setting up the matrices F~(rp) and


104 3 ViLmrtional <strong>Spect</strong>ra as u Probe oj'StrLrctzrra1 Order<br />

cm<br />

12oc<br />

1100<br />

1000<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

"<br />

n<br />

0<br />

I<br />

Figure 3-3. Dispersion curves of crystalhe orthorhoiiibic polyethylene with two molecules per unit<br />

cell (from [49]). Comparison with Figure 3-1 shows the splitting of the frequency branches and the<br />

shape of the acoustical branches at q~ - 0 (see text). Attention should be paid to the fact that the<br />

frequencies and the shape of the dispersion curves shown in Figure 3-1 and in this figure may differ<br />

because they have been calculated with two different force fields.


3.5 Towmu" Larger Molecules: F~om Oligoi?iers to Polyriers 105<br />

a<br />

-1 0 +I<br />

Figure 3-4. (a) Chemical<br />

structure of trarzs-polyacetylene<br />

and ib) model<br />

assumed in the calculation<br />

reported in the text. The<br />

model gives the numbering<br />

of the one-dimensional<br />

repeat unit cell and the<br />

labeling of the internal<br />

coordinates.<br />

GR (y) for a simplified chain of planar trans-polyacetylene, (-CH=CH-), where<br />

each CH group is taken as a point mass. Only in-plane motions are considered.<br />

From the viewpoint of chemistry and materials science, polyacetylene is a prototypical<br />

relevant material in molecular electronics and shows a low (HOMO-<br />

LUMO) band gap (= 1.4 eV) because pz orbitals of the carbon atoms in the sp2<br />

hybridization are strongly delocalized along the molecular chain (large conjugation).<br />

The distance of interaction between 7c bonds is yet unknown and is the subject<br />

of active research [40-421.<br />

Figure 3-4 shows the drawing of the actual molecule as well as of the simplified<br />

model with the labeling (1) of the sites of the translational repeating unit and the<br />

labeling of the internal coordinates of the 1 = 0 unit and the equivalent ones at sites<br />

+/- s. The unit cell is made up by two inasses (p = 2) and the in-plane internal<br />

coordinates are 2 x 2 = 4 which is also the size of the secular equation. We expect<br />

to calculate four branches in the dispersion relation, two of which are optical and<br />

two acoustical.<br />

We calculate explicitly the algebraic expressions of the elements of the GR(~)<br />

matrix which enters Eq. (3-43). The following simplifications have been accepted in<br />

such a simple calculation: (i) the groups C-H are taken as point masses of in = 13<br />

daltons; (ii) C=C and C-C bond lengths are equal; and (iii) bond angles are taken<br />

equal to 120".


106 3 Vibrutioricil <strong>Spect</strong>ra as a Probe of Structural Order<br />

(3-48)<br />

where p is the reciprocal of the C-C distance and p is the reciprocal of the mass of<br />

the CH unit taken as point mass.<br />

It is immediately seen that nonzero elements occur in the diagonal block and only<br />

in the two neighboring blocks which describe the interactions of the central unit<br />

with the nearest neighbors in position +1 and -1, i.e., the distance of interaction is<br />

s = 1, very short indeed. The construction of the GR(w) matrix through Eq. (3-44)<br />

yields a 4 x 4 matrix where the generic element has the following form:<br />

(3-49)<br />

The opposite is the case of the potential energy matrix whose structure is synibolically<br />

given in Table 16, p. 294 of [41].<br />

The most relevant fact is that while the Gp, matrix is certainly truncated at the +I<br />

and -1 units in the FR matrix the distance of interaction of the zero-th unit with its<br />

s-th neighbour is yet unknown since the so-called 'delocalization length' needs to be<br />

determined [41-431. The truncation at the site s of the long ribbon of interaction<br />

submatrices is then arbitrary. Many ab iiiitio calculations are at present dealing with<br />

this essential problem which is common to all poly-conjugated chains presently<br />

available.<br />

The truncation of the sum in Eq. (3-45) is then arbitrary, thus affecting the<br />

validity of the physical conclusions which may be derived. It follows that any<br />

addition or variation of terms or extension of the interactions (increase of delocalization<br />

length) will affect the numerical values of the elements of the FR(~)<br />

matrix,<br />

thus changing the shape of the dispersion curves [41, 421. It is apparent that the<br />

shape of the dispersion curves is a direct consequence of the electronic properties of<br />

the poly-conjugated chains [58] (Figure 3-5).<br />

An opposite case is found for trans-planar polyethylene taken as a prototype of<br />

all exclusively o bonded chains for which conjugation is not predicted (the existence<br />

of o-conjugated chains like polysilanes needs further studies). In the case of polyethylene,<br />

experiments or calculations indicate that the electronic interactions between<br />

o C-C bonds fall OR very quickly along the chain with s = 2 [59]. Indeed, as<br />

shown by Schachtschneider and Snyder, the normal modes of finite n-alkanes lie<br />

almost perfectly on the dispersion curves calculated for polyethylene [33-351. No<br />

chain length effect in the FK(~) matrix is revealed for these classes of systems.<br />

Analogously for o bonded tridimensional lattices like cubic diamond and silicon, a<br />

valence force field limited to a few first-neighbor interactions nicely accounts for the<br />

optical and neutron-scattering data available to date [60] (Figure 3-6).


1700<br />

- CIO<br />

--- ca<br />

c 22<br />

_- c 12<br />

l l l l l l l l l<br />

Figure 3-5. Effect of the long-range interactions (s in equation 3.45) on the dispersion curves of the<br />

frequency branch v3 (associated with the in-phase skeletal stretching) of trms polyacetylene calculated<br />

for the unsimplified and realistic equilibrium geometry. The force field has been derived by ‘LI~J<br />

initio’ calculations on polyene oligoinei-s with n = 2,4,5,6, and 11 double bonds [%I. Attention<br />

should be paid to the fact that the ‘softening’ of the ~ ( = 90) mode iiiodulated by the distance of<br />

iriteraction generates a ‘dispersion’ with chain length of a very strong and very characteristic<br />

Ranian line ([41, 421).<br />

These observations are essential when the dynamics of disordered chains will be<br />

discussed later in this chapter.<br />

3.6 From Dynamics to Vibrational <strong>Spect</strong>ra of<br />

One-Dimensional Lattices<br />

The experiments which provide direct information on the vibrations of one- and<br />

tridimensional lattices are infrared absorption (reflection) spectra, Ranian scattering<br />

and neutron scattering. For a general discussion on the theoretical principles and<br />

experimental methods the reader is referred to the abundant literature on these


108 3 Vibrational <strong>Spect</strong>ra as a Probe of Structurtil Order<br />

K<br />

DIAMOND r-x<br />

Figure 3-6. Example of lattice dyiiamical calculations on 0-bonded tridimenional crystals with<br />

short range interactions. Dispersion curves for cubic diamond along the r + (K) i X symmetry<br />

direction. Experimental points from neutron-scattering experiments; dispersion curves from least<br />

squares frequency fitting of a six parameters short range valence force field (from [60]). The Raman<br />

active phonon is the triply- degenerate state indicated with r;, near 1300 cm-'. Notice that at k # 0<br />

the degeneracy at is removed because of the lowering of the symmetry throughout the whole BZ.<br />

Notice also the three acoustic branches for which 1' - 0 at k i<br />

r.<br />

techniques (e.g., refs. [16-19]). Our aim is to present here methods for extracting<br />

from the experimental data the maximum amount of information on structure and<br />

properties of these relevant materials.<br />

The next step in our analysis is to define the experimental observables predicted<br />

by lattice dynamics. The spectroscopic selection rules are very restrictive and, of the<br />

very many phonon frequencies calculated, very few can be directly observed experimentally<br />

[ 16- 191 in the optical vibrational spectra.<br />

First, only phonons with k = 0 can interact with the electromagnetic wave; thus,<br />

only the fundamental frequencies at the centre of the Brillouin zone in k space are<br />

potentially infrared or Raman active. Second, further restrictions on the spectroscopic<br />

activity are introduced by the symmetry of the polymer chain. Once the


3.6 From Dynamics to Vibrational <strong>Spect</strong>ra of One-Diniensiorzal Lattices 109<br />

translational (or rototranslational) symmetry element is properly taken into account<br />

(in combination with point-group symmetry operations) a factor group isomorphous<br />

to a symmetry point group can be defined and k = 0 phonons can be separated<br />

into symmetry species, just as in the case of finite molecules. Traditional<br />

infrared and/or Raman activities can be predicted as well as the optical transitions<br />

associated to specific dipole allowed transitions with their directional properties for<br />

oriented samples.<br />

The physical meaning of the selection rule which states that only k = 0 phonons<br />

can be observed spectroscopically is that the atomic displacements in translationally<br />

equivalent crystallographic units must be all in-phase. When the geometry of the<br />

polymer chain (i.e. $(8, 1) and t = no. of turns of the helix within the rototranslational<br />

unit cell) is taken into account, phonons are infrared active for y = 0 and<br />

y z 0 (doubly degenerate); Raman activity is achieved for phonons with y = 0<br />

(nondegenerate), 0 and 20 (both doubly degenerate) [15, 61, 621. Degeneracy is<br />

clearly understood when folding of the dispersion curves is carried out at the zone<br />

boundary in going from y to k space. If p is the number of atoms in the chemical<br />

repeat unit for a helical polymer, one expects in the infrared 3p - 2 totally symmetric<br />

11 modes with q = 0 and 3p - 1 doubly degenerated I modes with y = 0.<br />

Coincidences with the infrared spectrum are expected in the Raman with the modes<br />

at y = 0 and y = 8, while extra modes are expected at y = 28 (practically never<br />

observed because their Rainan intensity is very weak).<br />

Let us further consider the direction of the transition dipole moments in infrared<br />

for a stretch-oriented polymer sample in which it is assumed that polymer chains<br />

are aligned along the stretching direction. Phonons with 9 = 0 have their transition<br />

dipole moment parallel to the chain axis which coincides with the direction of<br />

stretching (11 modes); modes with y = 8 have transitions moments perpendicular to<br />

the chain axis (1 modes). They can easily be identified in an infrared experiment on<br />

stretch-oriented samples carried out in polarized light. Analogously, the elements of<br />

the polarizability tensor can be selectively excited by Ranian scattering experiments<br />

on oriented polymer samples examined in polarized light in diflerent scattering<br />

geometries. For a study of the Raman spectrum of a stretch-oriented polyethylene<br />

rod in various geometries see [28] and Figure 3-7 [63].<br />

A few comments of general relevance need to be made.<br />

The shape of the dispersion curves and the corresponding eigenvectors which<br />

describe the vibrational displacements (phonons) depend necessarily on the vibrational<br />

force field adopted in the calculations. The fitting of the experiniental frequencies<br />

with the calculated ones, although being a good indication that the force<br />

field is reasonably acceptable, can by no means be taken as proof of its general<br />

reliability throughout the whole BZ. Indeed, fitting has been obtained only at two<br />

values of the phase coupling y. The intramolecular interactions may be different at<br />

y # 0 and y # 8 or 20, thus in principle changing the shape of the theoretically<br />

predicted dispersion curves (see. for example, [64]). Other independent experiments<br />

are needed before claiming the success and generality of the intramolecular force<br />

field.<br />

So far, we have discussed k = 0 phonons of the one-dimensional lattice and have


1 10 3 Vibrritionnl SIJectrcr as CI Pvohe of Stvuctrirnl Order<br />

<<br />

7<br />

T-<br />

I<br />

0<br />

0 Ln<br />

P-<br />

N<br />

0<br />

0<br />

co<br />

N<br />

0<br />

m<br />

co<br />

N<br />

0<br />

0<br />

m<br />

N<br />

0<br />

m<br />

Di<br />

0<br />

0<br />

0<br />

C’?<br />

I<br />

__-<br />

0 Lo<br />

0<br />

c3


3.6 Froin Dynaiiiics to T’ihratioiial Sjlectra of Onr-DiriimsioriLII Lattices 1 1 1<br />

neglected the possible existence of a tridimensional arrangement of the polymer<br />

chains in the solid. Experimentally, it transpires that for most of the polymers the<br />

entire observed spectrum can be accounted for in terms of the k = 0 modes of the<br />

single infinite chain [65]. When the material is melted, oi- is dissolved in some suitable<br />

solvents, k = 0 modes disappear and the spectrum becomes typical of a conformationally<br />

disordered ‘liquid-like’ molecule with the classical ‘group frequency’<br />

modes which can be interpreted using the classical spectroscopic correlations (see<br />

Section 3.2).<br />

The fact that, in going from the solid to the liquid state, maiiy bands disappear<br />

has been taken by authors as an indication that these bands must be associated with<br />

the material in the Crystalline state. In spite of the existence of clear and easy<br />

dynamical theories on polymer vibrations, a whole body of literature has accepted<br />

the direct correlation between k = 0 modes of the single chain and content of<br />

material in the crystalline (3D) state.<br />

This view is wrong both in principle and in practice. In spite of strong warning<br />

and extensive discussions (theoretical and experimental [65, 661) the general definition<br />

of ‘crystallinity bands’ has been widely and uncritically accepted by the chemical<br />

polymer community and even analytical determinations on the concentration<br />

of crystalline material have been carried out on polymers which show no direct<br />

spectroscopic indications of crystallinity [65, 661. The disappearance of such bands<br />

upon melting is simply due to the fact that the conformational regularity of the<br />

chain collapses and no 1 D-translational periodicity can be found anymore over a<br />

reasonable chain length. Chemical and stereo regularities are not modified in the<br />

melt or in solution, but all optical selection rules are removed because of the lack<br />

of phase coupling between adjacent units. From the above discussion it becomes<br />

apparent that the k = 0 bands previously discussed (generally and reasonably<br />

called ’regularity bands’ [65, 661) arise from a polymer molecule organized as a onedimensional<br />

crystal iiz uacuo, as if there were no intermolecular lattice forces. It<br />

becomes apparent that their labeling as crystallinity bands is a conceptual error.<br />

Certainly, the dynamical treatment can be and has been carried out for a few<br />

cases by taking into account the tridimensional arrangement of polymer chains [49].<br />

A model of suitable intermolecular nonbonded atom-atom potential has to be<br />

chosen critically [49, 671 and calculations can be carried out using the same principles<br />

discussed previously in this chapter. The number (q) of atoms per tridiniensional<br />

unit cell increases, the complexity of the BZ increases, many more phonon<br />

branches are calculated for different directions of the wave-vectors, and special<br />

symmetry directions and symmetry point in the BZ can be found depending on the<br />

syiiinietry of the whole lattice [60, 641. Typical examples are given in Figures 3-3<br />

and 3-6.<br />

Since intermolecular forces in polymers are weak, their effects on the phonons<br />

of the whole lattice are relatively small, thus originating small splitting of few of<br />

the regularity bands [68]. Rigorously speaking, the limited splitting of the regularity<br />

bands observed for a few polymers originate from phonons at the point ik, =<br />

k, = k, = 0) of the tridimensional BZ. Such splitting can indeed be considered as<br />

crystallinity band and certainly originate from material organized in a tridimensional<br />

lattice. This occurs when intermolecular forces are strong enough, the nuin-


112 3 Vibrutionnl <strong>Spect</strong>ru as ci Probe of Strirctzirul Order<br />

do0 2600 2600 2300 zboo 1foo 1400 1ioo Boo 500<br />

WAVENUMBERS<br />

Figure 3-8. Infrared absorption spectrum of highly crystalline polyethylene where crystal field<br />

splitting is shown in the 1450 and 720 cm-' range due to crystal field splitting (crystallinity bands).<br />

ber of chains in the unit cell is appropriate, and space-group selection rules are<br />

favorable [67-691. Among the true crystallinity bands we have also to consider the<br />

low-frequency lattice modes associated with the translation and librations of the<br />

molecules which behave almost as rigid objects moving within the 3D lattice.<br />

The classical case of crystalline polyethylene (Figure 3-8) may be mentioned in<br />

reference also to the vibrational spectra of its model compounds consisting of a<br />

series of n-alkanes. Relatively short alkanes with an odd number of C atoms crystallize<br />

in the orthorhombic lattice with two molecules in the unit cell. Relatively<br />

short n-alkanes with an even number of C atoms crystallize in monoclinic or triclinic<br />

unit cells with one molecule per unit cell. Very few of the normal modes of<br />

orthorhombic n-alkanes show in infrared and/or Raman doublets due to phonons<br />

at r (i.e., true doublets due to weak intermolecular forces; such splitting is coinmonly<br />

known and studied as correlation field splitting 1701); the spectra of monoclinic<br />

(or triclinic) n-alkane do not show any splitting as if the molecules, in a first<br />

approximation, were trans-planar, but in vuczio. Indeed, there exists only one molecule<br />

per unit cell and no correlation field splitting can occur. Analogously, when<br />

orthorhombic n-alkanes molecules are heated just before melting they transform<br />

in a probably monoclinic unit cell (there may be different interpretation of the<br />

so-called monoclinic modification). The classical doublets (phonons at r) observed<br />

in infrared and Raman disappear and singlets (phonons at k = 0) are observed as<br />

if the molecules were iiz ouc~io. This problem will be matter of extensive discussion<br />

later in this chapter.


3.7 The Case of Isotactic Polypropylene - A Tesbook Cme 113<br />

The fact which needs to be justified is that in some analytical determinations of<br />

the concentration of crystalline material in a polymer bulk, correlations are found<br />

between the intensity of the regularity bands and the data from X-ray diffraction<br />

experiments (which measures scattering from 3D periodicity). The measure of the<br />

intensity in infrared or Raman provides the amount of material organized as 1 D<br />

straight chains. If the ‘rIinylzrtti’ model is accepted, the measure of the intensity<br />

of the regultrr-ity bands [65, 661 gives an estimate of the concentration of straight<br />

spaghetti in an otherwise disordered environment of ‘boiled’ spaghetti (conformationally<br />

disordered chains). Whether the chains are packed in a crystal (3D order)<br />

or exist in a sort of ‘liquid crystalline’ arrangement (1D order) cannot be revealed<br />

from k = 0 regularity bands.<br />

True crystallinity bands have been observed for only a few polymers and their<br />

origin experimentally verified by isotopic dilution studies [70]. We may quote the<br />

classical prototype case of orthorhombic polyethylene [68, 691, orthorhombic polyoxymethylene<br />

[7 I], aiid-with caution-possibly a few others [72, 731.<br />

3.7 The Case of Isotactic Polypropylene -<br />

A Texbook Case<br />

It is beneficial here to discuss the case of isotactic polypropylene (IPP) as a worked<br />

example of the analysis of the vibrational spectrum of a polymer molecule. Let us<br />

assume that polymerization of propene with Ziegler-Natta catalysts has produced a<br />

fully head-to-tail isotactic polymer.<br />

Let us first consider the vibrational spectra (infrared) of IPP in the melt (Figure<br />

3-9a) [74]. In this physical state, polymer chains possess a conformationally irregular<br />

structure like any liquid branched n-alkane. We expect the infrared spectrum to<br />

consist of relatively few bands easily identified as the group frequencies of CH3 and<br />

CH2 and CH groups. Other modes may be identified with caution, but their location<br />

is irrelevant to the present discussion. The experimental infrared spectrum of<br />

Figure 3-9a is in full agreement with the expectations.<br />

It must be pointed out that the infrared spectrum of liquid IPP shows a few bands<br />

(especially the band at 973 cmr’) which are neither observed in the spectrum of<br />

atactic PP nor in the spectrum of liquid methyl-branched hydrocarbons. Moreover,<br />

in the Rainan spectra of molten IPP, three characteristic tacticity bands have been<br />

located at 1002, 973 and 398 cnir’ which allow IPP to be distinguished from the<br />

syndiotactic stereosionier (993, 966 and 310 cnr’). It follows that the ‘isotactic’<br />

chiral units generate vibrations mostly localized on the asymmetric carbon atom,<br />

but with some coupling with the neighboring units. This allows the chirality of the<br />

neighbors to be probed. These bands become characteristic of the tacticity and<br />

allow us to distinguish between isotactic and syndiotactic sequences [75, 761.<br />

Next, let the sample of molten IPP solidify. <strong>Polymer</strong> chains are allowed through<br />

their conformational flexibility to reach the minimum energy conformational struc-


114 3 Vil,rcitionciI S)ectra as ci Probe of Strzictitrnl Order<br />

li<br />

I)


3.7 The Cuse of Isotactic Polypvop.vleiie - A Tevbook Case 1 15<br />

ture which is known from calculations and experiments [77] to form a 1D periodical<br />

lattice with the coiiformational sequence -GTGTGT- which defines a threefold<br />

helix with $ = (3rr/3. d/3) and one turn (t = 1). The isolated chain belongs to the<br />

point group C3. Each chemical unit contains nine atoms and three chemical units<br />

(z = 3) are contained in the crystallographic repeat unit [77]. Thus, we expect<br />

3 x 9 = 27 phonon branches in q space or 3 x 9 x 3 = 81 in k space.<br />

Following the previous discussion we expect:<br />

0 27-2 totally symmetric q = 0 nonzero frequency modes of A species, infrared<br />

active with 1) dichroisni and Raman active in zz polarization.<br />

0 54-2 doubly degenerate modes (thus 27-1 frequencies) at y = B = 2n/3 of E<br />

species infrared and Ranian active (1). The phonons for q~ = 28 coincide with<br />

the phonons at q = 0 because of the periodicity of the functions ~ (y) (see Figure<br />

3-10a).<br />

The observed infrared spectrum of an isotropic sample of IPP is shown in Figure<br />

3-9a, and clearly identifies the regularity bands associated with 1 D periodicity. The<br />

assignment of the observed infrared bands to vibrations of species A, B = 0" (11) and<br />

E, B = 120" (1) is carried out on a stretch-oriented sample (Figure 3-9b).<br />

Dispersion curves have been calculated (Figure 3-10a) [56] from which spectroscopically<br />

active phonons have been derived for q = 0 and 27c/3. One-phonon density<br />

of states have been calculated and compared with the few data experimentally<br />

available from neuron-scattering spectroscopy (Figure 3-lob) (see Section 3.8). In<br />

Figure 3- 1 Oa, the dispersion curves of isotactic polypropylene are reported in terms<br />

of the phase coupling y; in the same figure the six high-energy branches (associated<br />

with the C-H-stretching modes in the 3000 cnir' range are omitted. The C-H<br />

branches are flat (i.e., C-H-stretching phonons practically do not show dispersion<br />

with p).<br />

Next, let IPP chains crystallize in a cell with space group isomorphous with the<br />

point group C, with four molecules per unit cell in three dimensions. In principle,<br />

one should expect (9 x 3 x 4 x 3) - 3 = 321 vibrations in the spectrum for the<br />

spectroscopically active phonons at the symmetry point r (i.e. k, = k, = k, = 0).<br />

This would mean that, in principle, each of the 25A + (2 x 26)E k = 0 bands of the<br />

single chain should split into four lines if intermolecular Van der Waals-type forces<br />

between atoms were strong enough to generate an observable splitting. Generally,<br />

such forces are very weak, especially since interchain distances are relatively large.<br />

If such splitting could be observed, true infrared crystallinity bands of crystalline<br />

IPP would be identified. The search for such splitting has not been easy and we<br />

made careful experiments at very low temperature aiming at shrinking the crystal,<br />

thus increasing the interchain interactions which should increase the correlation<br />

field splitting (i.e. k, = 0 line-group modes in 1D should show splitting into niultiplets<br />

due to phonons with k, = k, = k, = 0 in 3D (phonons at the syinnietry point<br />

r in the BZ). Moreover, bands should become sharper since 'multiphonon states'<br />

are depopulated. After much work the experiments were finally successful and the<br />

Ranian spectrum of IPP at liquid nitrogen temperature shows indeed multiplet<br />

splittings due to crystallinity (Figure 3-1 1) [78].


116 3 Vihtionul Syectru us n Probe oj Stritcturul Order<br />

Q


3.8 Density of Vibrational States and Neutuopi Scattering 117<br />

1360<br />

I/‘\<br />

i r<br />

/ A<br />

1040<br />

‘.<br />

81 0<br />

I<br />

/<br />

I<br />

I<br />

Figure 3-11. Raman spectrum of isotactic polypropylene from ‘smectic’ (- - -) to crystalline (-)<br />

(tridimensional order) obtained after long annealing processes at low temperatures. Raman regularity<br />

bands split into doublets (crystallinity bands) due to phonons at r (k, = k, = k, = 0) (from<br />

[781).<br />

The case of isotactic polypropylene becomes an interesting texbook case which<br />

shows clearly and separately the many steps in the formation of a polymeric<br />

material.<br />

3.8 Density of Vibrational States and Neutron Scattering<br />

The next step in the development of vibrational spectroscopy of large organic<br />

molecules is to realize that other physical measurements, such as neutron-scattering<br />

spectroscopy, can provide essential information on the normal vibrations of organic<br />

molecules. This information is complementary to that provided by infrared and<br />

Raman spectroscopy which is intrinsically limited by the optical selection rules.<br />

Indeed, the limitations to the spectroscopic activity imposed by symmetry selection<br />

rules are very strict and select only a few phonons. Moreover, when some sort of<br />

molecular disordering is introduced into the system (and this is always the case),<br />

symmetry is removed, selection rules are relaxed, and the spectra show sharp<br />

absorptions (or Raman scattering) originating only from characteristic localized<br />

vibrations floating on top of a broad background absorption whose origin needs to<br />

be explored and understood.


11 8 3 Vihrationul <strong>Spect</strong>ra ns a Probe of Structural 0rdek.v<br />

a b C


3.8 Dtvz.sit,v of L’ihrntioizal Strrtes aid Nezrtroiz Scutteriny 1 19<br />

We will introduce the basic principle and use of neutron-scattering spectroscopy<br />

by first introducing an important dynamical quantity, namely the density of vibrational<br />

states g(i1).<br />

Let g(v) be the fraction of normal modes with frequencies in the interval v and<br />

v + dv with dv 0. g(v) can be calculated analytically only for very simple atomic<br />

chains which are never encountered when realistic polymeric materials must be<br />

studied.<br />

Let us take a very large molecule and let 3N z C be the number of vibrations of<br />

the whole system; if c is the number of frequencies in the interval AV = v + dv then<br />

go)) = c(Av)/C (3-50)<br />

The total density of states is evaluated as the sum over all the r branches of the<br />

dispersion relation (Eq. (3-43)). Numerical methods such as the ‘root sampling<br />

method’ have been proposed for simple one- or tridimensional lattices [511. When<br />

the systems become more complex, as in the cases discussed in this chapter, the<br />

easiest numerical method is that proposed by Dean based on the application of the<br />

Negative Eigenvalues Theorem (NET)[79]. The method is extremely useful and has<br />

been extensively used in the laboratory of the writer for many polymeric cases. The<br />

details of the method are discussed in the next section.<br />

NET is a way to calculate g(v), i.e., it provides the number (or the fraction) of<br />

vibrational energy states or normal modes in a chosen y (or k) interval. g(v) can be<br />

plotted as an histogram whose accuracy depends on the frequency interval Ait chosen<br />

in the numerical calculation (numerical resolution of the computer experiment).<br />

This will be used later in this chapter in the study of disordered systems.<br />

Let us consider g(v) of an ideal infinite polymer and examine the case of singlechain<br />

polyvinylchloride in its three possible structures (threefold helix isotactic,<br />

trans-planar syndiotactic, folded syndiotactic) [go]. Foi- easier understanding, Figure<br />

3-12 shows both dispersion curves and g(v) for this polymer in its three possible<br />

structures. It is apparent that the shape of g(v) is determined by the shape of the<br />

dispersion curves which in turn depend on the vibrational force field used in the<br />

calculation of ~(y). Flat sections of v(q) give rise to strong ‘singularities’ in g(v).<br />

The shape of the dispersion curves and of electronic and/or vibrational g(v) for<br />

both electronic bands and vibrational bands have been the subjects of extensive<br />

theoretical treatments [51]. For infinite and perfect systems, the shape of g(i1) is<br />

characterized by different types of singularities which are related to the dimensionality<br />

of the systems and to many other physical factors. We restrict our analysis to<br />

the case of polymeric materials which can be considered as one-dimensional lattices.<br />

Our discussion can only be sketchy and qualitative, but it is aimed at pointing out<br />

the features most relevant to the problems of disorder treated in this chapter.<br />

T<br />

Figure 3-12. Phonon dispersion curves v(p) (a, c). i(k) (b) and density of vibrational states gii,) from<br />

0 to 1500 cn-’ of three possible models of a chain of polyvinylchloride showing the lR-Rainan<br />

active and the inactive ‘singularities’ corresponding to flat sections of the disperion CUI-vcs (from<br />

[801).


120 3 Vibratiorial <strong>Spect</strong>ra us a Probe of Structural Order<br />

The following observations should be kept in mind.<br />

1. Because of the shape of the calculated v(q), which may show many maxima and<br />

minima (due to the vibrational coupling intrinsic to the system and to the force<br />

field used), g(v) will also show corresponding strong or weak singularities. Some<br />

of these singularities necessarily coincide with k = 0 spectroscopically active<br />

phonons, while other will not be seen in the optical spectrum of a perfect system,<br />

but will (in principle) be observed in the neutron-scattering spectra and possibly<br />

in the optical spectra of a partially disordered polymer. In a disordered material,<br />

no symmetry selection rules are active and all modes may gain some activity.<br />

The vibrational spectra of a disordered material can thus be considered the<br />

mapping of g(v) dipole (or polarizability) weighted. The g(v) calculated with<br />

NET or any other numerical or analytical method is, instead, dipole (polarizability)<br />

unweighted. Any comparison with the experimental g(v) must be made<br />

with great caution.<br />

2. As discussed in Section 3.5 for tridimensional lattices, v i 0 for k 40 for the<br />

three acoustical branches with a positive slope and with a discontinuity at k = 0<br />

(r). The shape of g(v) i 0 has been matter of strong interest in physics for the<br />

calculation of specific heat and related thermodynamic quantities. As already<br />

mentioned for one-dimensional lattices in LIUCUO, some of the acoustical branches<br />

tend asymptotically to zero. The derived calculated g(v) shows a very strong<br />

singularity at v = 0. Such a singularity is meaningless and only due to the limitation<br />

of the molecular modes adopted in the calculations (1D lattice) which is<br />

unable to account for intermolecular forces. Also, for real polymer samples the<br />

experimental g(v) i 0 for k i 0 as expected for classical crystals since they do<br />

interact with the neighboring chains with very weak intermolecular forces.<br />

3. Additional singularities both in calculated and experimental g(v) are expected at<br />

very low energies (~0-30 cm-’) originating from very low-frequency 3D-lattice<br />

phonons. These singularities may coincide with the optical phonons in infrared<br />

and/or Raman spectra at very low frequencies.<br />

As already anticipated, a complementary experimental technique for deriving<br />

information on the dynamics (frequencies and vibrational amplitudes) of polymers<br />

or of materials in general is the use of inelastic neutron-scattering techniques (INS).<br />

After a long development time, during which experiments were difficult and provided<br />

limited information, the instruments in a few specialized centers recently<br />

began to provide detailed data covering the whole spectrum. Thus, we predict a<br />

‘renaissance’ of INS techniques for the studies of molecular and lattice dynamics.<br />

For a thorough discussion of INS spectroscopy we refer the reader to specialized<br />

publications [8 13 (see also other references quoted below); here, we restrict ourselves<br />

to the basic principles of the technique connected especially with the dynamic<br />

quantities which are of interest in optical spectroscopy.<br />

Let a beam of cold neutrons made mono-energetic with wave-vector kll be shone<br />

onto a sample, and let the beam be scattered inelastically and incoherently. The<br />

outgoing neutrons have wave-vector kf and are collected and detected by suitable<br />

devices. Since the energy of the thermal neutrons is of the same order as that of


3.8 Driisity of Vibratioiid StLitrs arid Neutron Scuttering 121<br />

lattice and molecular vibrations, neutrons impinging on the sample may gain or lose<br />

energy in a way proportional to g(v).<br />

It has been shown [S2] that the scattering cross-section of a polycrystalline material<br />

is approximately the same as that of cubic crystals. We are, at present, concerned<br />

with one-phonon processes and neglect the possibility that multiphonon<br />

processes may occur (also in vibrational spectra we have neglected anharmonic<br />

effect with the consequent appearance of overtones, combinations, and hot bands).<br />

The one-phonon scattering cross-section for a cubic lattice can be written as [83,<br />

S4]:<br />

d2a,,,,h/d0 do = A1 kf 1 /I ko lep2w (hK2/2M)[g(co)/oi]<br />

x [h'(of - wo + o)/(eP" - I)] (3-51)<br />

In Eq. (3-51), and in a few equations which will follow the vibrational frequencies<br />

are not expressed as v in wavenumbers, but as w in Hz. In Eq. (3-51), A is the<br />

scattering cross-section of the scattering nucleus (since the H atom has a very large<br />

cross section, namely AH = 82.5 barn, Ac = 5.5 barn, molecules containing hydrogen<br />

are very suitable for neutron experiments), lkfl and lkol are the moduli of the<br />

wave-vectors of the scattered and incident neutron respectively, e-2w the Debye-<br />

Waller factor or temperature factor, K = kf - ko, M the mass of the unit cell, g(w)<br />

the desired density of vibrational states, and p = h/?kBT (where kB and T are the<br />

Boltzman factor and the temperature respectively).<br />

It is clear that thermal population is strongly contributing to the scattering crosssection<br />

with the Debye-Waller factor and the term p. Measurements must then be<br />

carried out at very low temperature.<br />

The experimental g(co) (or g(v)), while showing qualitatively all the features of<br />

the calculated g(w), cannot yet be quantitatively compared because the vibrational<br />

amplitudes (generally called polarization vectors in neutron-scattering circles) of the<br />

various normal modes (phonons) are not yet taken into account. The situation is<br />

similar to the features of the vibrational spectra of an isotropic sample compared<br />

with the features of the spectra in polarized light of an oriented anisotropic sample.<br />

Let us take a stretch-oriented sample of a polymer subject to a neutron beam with<br />

incident wave-vector ko. A more complete expression of the scattering cross-section<br />

is the following [S2, 841:<br />

Some of the terms appearing in Eq. (3-52) are already defined for Eq. (3-51). The<br />

new terms are the following: A, is the scattering cross-section of the n-th nucleus of<br />

the unit cell, e-2W11 the Debye-Waller factor for the n-th nucleus of mass MI,, coi the<br />

frequency of the j-tli normal mode of the r-th branch in the dispersion relation. The<br />

most important quantity is ek. C:' which gives the projection along ko of the atomic


122 3 Vibrational <strong>Spect</strong>ra as N Probe of Structural Order<br />

displacement C of atom n for the j-th normal mode of the r branch with frequency<br />

wi. The sum is extended to all the atoms n in the unit cell and to all branches r in<br />

the dispersion relation. The Debye-Waller factor W, of each atom n has an explicit<br />

expression [83, 841 which again contains quantities defined in Eqs. (3-51) and (3-52).<br />

In spite of the seemingly complex expressions given in Eqs. (3-51) and (3-52) for<br />

the scattering cross-sections, all terms appearing in these equations are known from<br />

dynamics and from experiments? and have been defined in the previous sections of<br />

this chapter. Automated computer programs have been added to the classical programs<br />

for normal coordinate calculations of polymers for the calculations of the<br />

INS spectra.<br />

The reason of requiring the experiments to be carried out on stretch-oriented<br />

samples mounted with known geometry with respect to the incoming neutron beam<br />

is the existence of the ek . Cy,r term which indicates that the scattering cross-section<br />

is determined by the directional interaction of the neutrons with wave vector ko and<br />

the direction of the atomic displacements. This situation is very similar in the experiments<br />

of infrared spectroscopy in polarized light with a stretch-oriented material.<br />

In dichroism experiments the absorption intensity is proportional to /pi. El’, i.e.,<br />

to the square of the projection of the molecular dipole moment change p during the<br />

i-th normal mode onto the electric field of the light beam.<br />

The g(w) measured with INS experiments on stretch-oriented materials is then<br />

‘amplitude-weighted’ and is ‘directional’. The directionality of the displacements<br />

associated with a given normal mode can be determined by placing ko 11 and I to<br />

the chain direction.<br />

The history of polymer neutron spectroscopy can be divided in two different time<br />

domains. The enthusiasm for INS experiments in polymers reached its highest peak<br />

in the 1970s, but the experiments have been restricted to a few polymers with a<br />

limited number of data [84, 851 which could hardly solve particular problems in<br />

polymer dynamics. The availability of modern instruments with higher resolution<br />

and with the possibility of removing most of multiphonon scattering has presently<br />

revived the field and vibrational spectroscopists await with great interest for the<br />

revival of this important field of research.<br />

In Figure 3-13 we show the most recent INS results reported by Parker [86] on<br />

oriented polyethylene at 30 K obtained with the high-resolution, broad-band spectrometer<br />

(TFXA) at ISIS pulsed spallation neutron source at the Rutherford Appleton<br />

Laboratory, Chilton, UK. Notice that the whole spectral range 0-4000 cnc‘<br />

range has been explored with considerable resolution, especially in the lower-energy<br />

region.<br />

3.9 Moving Towards Reality: From Order to Disorder<br />

3.9.1 Basic Concepts<br />

In reality, polymeric materials never possess the perfect chemical stereochemical<br />

and conforinational regularity which were assumed at the beginning of Section 3.5.


3 9 Mouiiiq ToIvmds Rerrlitv: From Order to Disordrr 123<br />

I " ~ ' " " l " " l ' ' ' '<br />

I 1<br />

Figure 3-13. Inelastic neutron scat- 1 , , , , I , , , , I , , , I , , , ,<br />

teriiig (INS) spectrum of polyethyl- 0 1000 2000 3000 4000<br />

ene at 30 K (from [86]).<br />

wavenumbedcm -'<br />

The ideal infinite polymer chain iiz ziucuo has allowed to consti-uct dynamical<br />

theories which describe ideal spectroscopic properties of the polymer determined<br />

through strict optical selection rules and obtained by ideally clean experiments. In<br />

reality, polymerization reactions never yield perfect chemical linking and perfect<br />

stereo-regularity; these chemical and stereochemical defects necessarily introduce<br />

conformational distortions as to avoid steric encumbrance. Moreover, the complex<br />

processes of crystallization from melt or solution of a very long and flexible object<br />

implies the generation of various kinds of morphologies which reflect particular<br />

packing of 'conformationally' ordered chains. New concepts such as chain folding,<br />

looping, tie molecules, fringed micelles, etc., are necessarily introduced for the description<br />

of the building up of the supermolecular structure of a polymer chain. It<br />

follows that any polymer material which has been allowed to crystallize consists of<br />

a fraction of matter crystallized in a 3D-lattice and a fraction of 'conformationally<br />

irregular' material [46, 871.<br />

In polymer material science it is essential to know the detailed independent<br />

structural characteristics (and the relative concentrations) of the crystalline and<br />

'amorphous' (or irregular) fractions of a polymer sample. All mechanical, viscoelastic<br />

or, in general, the physical properties of real polymeric materials depend on<br />

the relative fraction of crystalline versus non crystalline material.<br />

Since the spectroscopic manifestations of molecular disorder are many and very<br />

characteristic, the vibrational spectrum has become a usefyl probe for molecular<br />

order-disorder which can be described at the level of a few Angstroms.<br />

<strong>Spect</strong>roscopists are asked to find ways to account for the new spectral features<br />

associated to disorder, thus making vibrational spectra useful for the practical<br />

structural characterization of a real polymer material. Obviously, the theories presented<br />

in the previous sections of this chapter must be adapted or re-worked in<br />

order to understand the spectra of disordered organic polymers.<br />

The prototypical case of an experimental observation to be accounted for is the<br />

spectrum of a sample of orthorhombic polyethylene [88] (Figure 3-14). Let us con-


124 3 Vibvcrtinntrl <strong>Spect</strong>vci as ci Prohe qf Stvzicturrrl Order<br />

100, II I I I<br />

Figure 3-14. Comparison of the calculated<br />

g(v) of single-chain polyethylene with the<br />

infrared spectrum of an e\rtended chain of<br />

polyethylene.<br />

sider the experimental infrared spectrum of high-density crystalline polyethylene<br />

(Figure 3-14). The infrared active phonons at r for an orthorhonibic lattice with<br />

two chains per unit cell show the predicted crystallinity bands as doublets originating<br />

from the k = 0 modes of the former regularity bands for a chain with a ID<br />

symmetry operator $(Q = n, d/2). In Figure 3-14, the infrared spectrum is superposed<br />

to the singularities of the calculated g(v) while the Ranian active peaks are<br />

indicated in the figure. It is quite clear that not all the peaks in the observed vibrational<br />

infrared or Ranian spectrum find a one-to-one correspondence with the<br />

singularities in the calculated g(v). It follows that, in addition to the ‘perfect’ part of<br />

the all-trans polyethylene chain, extra structural features must exist whose existence<br />

is indicated by the additional peaks in the infrared or Raman spectra. It will be<br />

shown that the extra absorption or scattering are associated with various kinds of<br />

conformational defects.<br />

In this section we proceed as follows.<br />

1. First, we take the perfect and infinite polymer chain and truncate the chain in<br />

increasingly shorter segments and predict theoretically and watch experimentally<br />

the effects of such truncation. The structural perfection is retained.<br />

2. Any kind of chemical and/or structural disorder can be introduced into the<br />

chain. We chose the various plausible kinds of disorder, their population and<br />

distribution, and try to predict the infrared, Raman and neutron-scattering<br />

spectra by suitable theories and computational techniques.<br />

3. Calculated theoretical and experimental cases will be presented.<br />

3.9.2. Finite Chains<br />

The theoretical modeling of the vibrations of molecular chains with finite length has<br />

been nicely treated by Zbinden [18] and by Snyder and Schachschneider [9, 341.<br />

While the approach by Zbinden is mathematically complete, but it applies to simplified<br />

monoatomic chains, quite away from chemical reality, the model by Snyder<br />

and Schachschneider is more directly applicable to real molecules and will be


3.9 Moving Towards Reality: From Order to Disorder 125<br />

quickly mentioned here. We restrict our discussion to the most connnon case of 0-<br />

bonded chains, which, as discussed in Section 3.5, do not permit long-range electronic<br />

coupling.<br />

Let y~ be the number of identical chemical units in a chain segment (different<br />

chemical groups at either ends are neglected) and let each chemical unit consist of n<br />

atoms; we treat here the case of a model molecule with free ends [18]. Let us take<br />

one chemical unit and its 3 x n oscillators (which, for sake of clarity, we describe<br />

here as internal (R) or group coordinates II). Because of elastic coupling with its<br />

neighbor at either side, each oscillator couples with its identical neighbors and<br />

generates 11 normal modes which can be described as waves characterized by a<br />

phase coupling y = jn/ll (where j = 0,1,2,. . . y - 1). Their 11 frequencies y(y) lie at<br />

finite y values along the dispersion curve of the corresponding infinite polymer with<br />

identical chemical, stereochemical, and conformational structure. Each vibration<br />

corresponds to a given phase coupling defined by the number of chemical repeat<br />

unit which make up the finite chain. The introduction of phase coupling implies the<br />

fact that within such short chains, nomal modes describe quasi regular waves (quasi<br />

phonons) similar to those which propagate in an infinite chain. The remaining<br />

modes which do not lie on the branches of the dispersion curves must be associated<br />

with the vibrations localized on the end groups and can be identified by the constant<br />

values of the frequencies when chain length is changed. It follows that:<br />

(i) if I? is the number of repeat units making up the finite chain (neglecting the two<br />

end groups) one expects to find a sequence of bands (or band progression)<br />

which lie on each of the dispersion branches of the corresponding polymer.<br />

Let us take for example the n-alkane n-nonadecane (CH~(CH~)I~CH~). Let us<br />

focus on the vibrations of the segment consisting of y~ = 17 CH2 groups all in<br />

trails conformation. Each CH2 generates 3 x 3 = 9 vibrations (CH2 antisymmetric<br />

and symmetric stretch, CH? bending, wagging, twisting, rocking,<br />

antisymmetric and symmetric C-C stretch, CCC bending, and C-C torsion).<br />

Each of these ‘oscillators’ generates 17 waves, each of which is characterized by<br />

a phase coupling nj/ 17. The corresponding frequencies lie on the dispersion<br />

curves of a single chain of infinite all-trans polyethylene. The band progressions<br />

observed for n-nonadecane will be discussed in Section 3.19.<br />

(ii) The frequency range spanned by each band progression depends on the extent<br />

of intramolecular coupling. If the dispersion of the frequency branch is small,<br />

the progression is squeezed, many bands overlap (e.g., the stretching vibrations<br />

of C-H groups), and no progression can be observed experimentally. When<br />

intramolecular coupling is large, band progressions are clearly observed<br />

(-1 100-720 cm-’ is the frequency range covered by CH2 rockings in all-trans<br />

n-alkanes).<br />

(iii) The intensity in infrared and/or Raman of each of the bands within the progressions<br />

depends on the dipole (polarizability) changes associated with the<br />

corresponding vibration (quasi phonon with y, z jn/y). We have previously<br />

seen that for an infinite chain only k = 0 phonons are active and all other are<br />

inactive, as dipoles or polarizability changes cancel out in the case of perfect<br />

phonon waves. In going from short to longer chains, optical selection rules


tend to the limit of the infinite case and we expect to find in the spectra of the<br />

finite, but long, chains clear (generally, but not always strong) quasi k = 0<br />

phonons while the other members of the band progression will show quickly<br />

decreasing intensity. This is the typical case observed in the series of n-alkanes<br />

by Snyder and Schachschneider [9. 341.<br />

(iv) It is obvious that if normal-mode calculations have permitted calculation of the<br />

dispersion curves for the infinite polymer, the identification of the band progressions<br />

in finite chains with identical structure is made easier. In contrast. if a<br />

systematic study is carried out on a series of inolecules of the same chemical<br />

class, but with increasing chain length, the experimental identification of the<br />

band progressions allows experiinental determination of the phonon dispersion<br />

branches from which a reliable force field can be derived. Vibrational spectroscopy<br />

then becomes a complementary tool (sometimes unique) to the extremely<br />

more expensive and elaborate neutron-scattering techniques. The whole work<br />

has been very clearly and precisely explained and successfully applied by<br />

Snyder and Schachschneider [9, 341.<br />

3.9.3 <strong>Polymer</strong> Chains with Structural Defects<br />

In order to model the reality of a polymeric material we need first to introduce in<br />

the calculations the various energetically possible structural defects. The tjFes of’<br />

defects to be considered are the following:<br />

0 Clzenzical defects, e.g., head-to-head linking in an otherwise head-to-tail chain. A<br />

typical case is the real structure of the chain of polyvinylfluoride ICHz-CFz),,<br />

which contains a sizeable fraction of undesired head-to-head defects [89]. Once<br />

the chemistry of a given polymer is approximately known, other kinds of defect<br />

structures can be envisaged. A typical case is often found for isotopically substituted<br />

chains when substitution is not ideally complete.<br />

Stereocheiiiicnl defects, e.g., syndiotactic configurations in an otherwise isotactic<br />

chain.<br />

Cotiforwmtiond defects, e.g., gauche conformers in an otherwise all-trnizs chain<br />

structure.<br />

It is obvious that, because of intramolecular interactions, the introduction of<br />

some kind of chemical and/or stereochemical defects forces also the introduction of<br />

conformational defects.<br />

Next, the model requires the definition of the concentrafion mid distribiitiori (e.g.,<br />

Bernoullian, Markovian, etc.) of defects. If a small concentration of defects with a<br />

random distribution is considered, defects most probably are isolated in the host<br />

polymer ID lattice. When the concentration increases, even a random distribution<br />

generates both isolated defects and a distribution of ‘islands’ of various lengths (see,<br />

for instance, [90, 911).<br />

When such variables are well defined, calculations require the construction of the<br />

usual dynamical matrix and the solution of the corresponding eigenvalue equation.


3.9 Moving Towards Reality: Froin Order to Disorder 121<br />

It becomes apparent that as the translational symmetry of the chain is removed<br />

by the existence of defects, any periodicity is lost and Eq. (3-431 cannot be used.<br />

Ratter, Eq. (3-17) must be used which corresponds to the dynamical case of a finite<br />

molecule. Since in this case our molecular model is huge and has no symmetry, the<br />

size of the secular equation to be solved becomes extremely large. Moreover, such<br />

types of study require the freedom to try many models with different kinds, concentrations,<br />

and distributions of defects. Last, but not least, in order to play a game<br />

as close as possible to reality in these studies, molecular models must be composed<br />

of as many monomer units as possible.<br />

In solid-state physics, the vibrations of very simple lattices containing a small<br />

concentration of simple defects have been treated with sophisticated analytical<br />

treatments using Green’s fiinctions 1921. Even if some authors have bravely tackled<br />

an analytical solution of the dynamical problem of disordered polymers [93], they<br />

were forced to introduce into the molecule such drastic structural simplifications<br />

that the flavor of chemistry has been lost and the theoretical molecular models have<br />

become, again, too unrealistic.<br />

For our complex molecular systems the problem must be solved by numerical<br />

methods. The problem of the solution of a huge secular equation producing thousands<br />

of vibrational frequencies can be solved by the application of the so-called<br />

‘Negative Eigenvalue Theorem’ (NET) originally proposed by Dean [79] who<br />

aimed at the calculation of the vibrational spectra of disordered ice. We think that<br />

the original work by Dean did not receive due acknowledgement; indeed, his<br />

methods has been widely applied (with little due reference to Dean) in solid-state<br />

physics and in molecular theories whenever the eigenvalues corresponding to vibrational<br />

or electronic states of huge and disordered systems had to be calculated. (As<br />

an example of the calculation on the electronic states of disordered systems, see<br />

[941).<br />

Dean’s method calculates the density of states (vibrational or electronic) in a<br />

given energy range. The whole energy range can be spanned by the calculation and<br />

the complete g(v) is obtained and can be plotted as an histogram to be compared<br />

with the experimental spectrum. The accuracy of the histogram can be improved by<br />

narrowing the steps (Av) in the energy interval in each cycle of the calculation (i.e.,<br />

by improving the resolution of the numerical experiment).<br />

Dean’s method works well for band-matrices which are always found in the case<br />

of polymer chains or in the case of tridimensional lattices with short-range interactions.<br />

The reader is referred to the original papers for a discussion of the method,<br />

its advantages and limitations [91. 951.<br />

Once the histogram of g(v) is plotted, one needs to find the eigenvalues for such<br />

huge matrices corresponding to the approximate eigenvalues comprised in a given<br />

energy range. The problem has been solved with the application of the Wilkinson’s<br />

inverse iteration method [96] (hereafter referred as to IIM; see for instance [97]).<br />

The comparison of the calculated g(v) with the actual experimental infrared,<br />

Ranian and neutron-scattering spectra requires some care and possibly more refined<br />

calculations with improved resolution. g(v) gives one information, namely how the<br />

very many normal modes are clustered in a given frequency range. From IIM we<br />

may also learn the shape of the normal modes at a chosen frequency. Nothing is


128 3 Vibrational <strong>Spect</strong>ra as a Probe of Stvuctzirnl Order<br />

known about the dipole transition moments (infrared intensities) or Raman or<br />

neutron-scattering cross-sections unless specific calculations are carried out using<br />

other models for the electronic distribution (dipole derivatives, polarizability derivatives,<br />

and vibrational amplitudes respectively 1971). Care must then be taken in<br />

carrying out the comparison between calculated g( v) and observed spectra.<br />

3.10 What Do We Learn from Calculations?<br />

We have just outlined the way theoreticians in polymer dynamics face the problem<br />

of the understanding of the vibrational spectra of polymers in their realistic state.<br />

The reader interested in the theoretical aspects will find in the references quoted all<br />

indications which will enable him or her to grasp the theory and to carry out the<br />

calculations.<br />

However, a larger number of reader do not wish to play with calculations, but<br />

simply to know what we have learned about polymer structure and spectra using<br />

these calculations and to apply quickly and directly the knowledge acquired to their<br />

own problems. We report here the concepts of general validity in polymer spectroscopy<br />

which were derived from calculations and which can be applied to any polymer<br />

(for a general discussion, see [ 16, 19, 661).<br />

The reader is certainly familiar with the traditional group-frequency approach<br />

generally used in the past 50 years for the chemical application of the infrared<br />

spectra (Section 3.4). Correlation tables and books have been written which discuss<br />

group-frequency correlations and provide the way to carry out a chemical diagnosis<br />

from the vibrational spectrum (so far, the infrared spectra have enjoyed great<br />

popularity; the reader is advised to extend their interest to the very useful and easily<br />

available Raman spectra).<br />

As already discussed in Section 3.4, from the dynaniical viewpoint chemical correlations<br />

are based on the fact that there are a few vibrations (normal modes) which<br />

are Imyely localized on the functional group and give rise to medium/strong absorption<br />

bands in infrared (scattering lines in the Raman) at specific and wellisolated<br />

frequencies. The occurrence of such bands in the spectrum implies the<br />

existence of such a functional group in the molecule under study. On the other<br />

hand, when the vibrations involve a large molecular domain (collectiue motions) the<br />

observed bands are no longer so characteristic of a given group of atoms.<br />

The vibrations of disordered polymer chains follow the same kind of conceptual<br />

path which, on the other hand, was also found earlier by physicists in the study of<br />

lattice dynamics of very simple disordered lattices [98].<br />

Let us make the problem simpler by considering briefly the small molecular unit<br />

which contains the defect as an isolated entity consisting of n atoms which generate<br />

3n normal vibrations that occur somewhere in the vibrational spectrum. Next, we<br />

insert just one defect unit into the otherwise perfect polymer chain and allow the<br />

whole system to display its own new dynamics through the observed (or calculated)<br />

vibrational spectra.<br />

The dynamics (new frequencies and displacements, see Eq. (3-17)) of the defect


3.10 Wlmt Do We Learn from Cnlcailcitioizs? 129<br />

COz rock<br />

___c<br />

I<br />

GAP W S ~ W ~<br />

/<br />

Figure 3-15. Example of a dynainical situation of a po1yniel;ylene r chain -(CH2)n- which contains<br />

defects consisting of CD2 groups and of CH3 units at either ends. The motion of CD:! rocking<br />

occurs in a gap and does not couple with the host lattice, thus generating localized modes; on the<br />

contrary the frequencies of both external deformation and umbrella motions of the CH3 group<br />

occur in the frequency range spanned by the dispersion curves of the host lattice; coupling takes<br />

place and resonance modes are generated.<br />

embedded in the polymer chain is determined by the masses and the geometrical<br />

arrangement of its vibrating atoms (G matrix, Eq. (3-15)) and by the new intramolecular<br />

forces, i.e., by the new force constants (type, strength, and distance of<br />

intra-unit and inter-unit coupling (F matrix, Eqs. j3-2), (3-3) and (3-16)).<br />

Calculations carried out so far on the simplest prototypical classes of polymer<br />

describe substantially three types of phenomena (Figure 3- 15):


130 3 Vibratiod <strong>Spect</strong>ru as LL Probe of Structural Order<br />

Phonon mode<br />

a<br />

Resonance mode<br />

b<br />

Figure 3-16. Dynamics of a one-dimensional lattice containing one defect. Vibrational displacements<br />

of each unit are rotated of 90” for a better display of the normal modes. (a) uiiperturbcd<br />

phonon of the host lattice; (b) resonance mode: and icj gap/local mode if!-om [66]).<br />

1. Some of the normal modes of the defect strongly couple with the normal modes<br />

(phonons) of the polymer and generate collective motions which lose all the<br />

characteristic of the motion of the isolated unit. These normal modes are generally<br />

labeled as ‘resomince ~zodes’ (Figure 3- 16) because their frequencies are close<br />

enough to those of the host lattice that they generate mechanical resonance with<br />

some of the normal modes of the polymer. This is generally found when the<br />

frequency of the separated unit happens to occur within the range of frequencies<br />

spanned by one of the dispersion curves of the perfect host polymer 1D lattice.<br />

2. When the frequency of the separated unit differs strongly from the frequencies<br />

spanned by the dispersion branches of the polymer chain, the motion of the unit<br />

cannot couple with the motions of the host lattice, it cannot be propagated and<br />

generates a mode strongly localized in space (at the site of the defect within the<br />

polymer) and in energy (at a characteristic specific frequency). These modes are<br />

generally labeled as ‘yap modes’ or ‘out-oflbancl modes’, since the frequency of<br />

the isolated units occurs either in a gap between dispersion branches 01- above<br />

the highest frequency branch respectively. Out-of-band modes in organic polymers<br />

can only be observed in the case of perdeuterated polymers which contain<br />

H atoms as defects. In this case, C-H stretching modes occur at frequencies<br />

higher than any other normal modes of the polymer.<br />

3. Compared with the results from the dynamics of simple impure 3D lattices, in<br />

the case of 1D lattices (polymers) localized modes (or quasi-localized modes) can<br />

sometimes also be generated if the frequency of the defect unit occurs within the<br />

frequencies spanned by a dispersion branch of the perfect host lattice. As the


+<br />

3.11 A Very Siriiple Ccrse: Lattice Dyriurriics of HCI-DCI Mi-xed CrjistLiIs 131<br />

polymer chain is conformationally flexible it may well occur that, for a certain<br />

geometrical arrangement of the atoms. the motion of the defect unit happens to<br />

be orthogonal (or quasi-orthogonal) to the displacements of the atoms during a<br />

certain vibration of the host lattice. It has been shown [99] that in an n-alkane<br />

chain the CH: wagging of the central unit in a conformational sequence such as<br />

-(G)-CH?-(G)- generates a pseudo-localized CH2 wagging motion almost fully<br />

decoupled from the vibrations of the host lattice, even if the frequency occurs in<br />

the frequency range spanned by the CH? wagging dispersion branch [99].<br />

The observation of characteristic gap modes in disordered polymers becomes<br />

then the basis for the use of the vibrational spectra as useful probes for the structural<br />

characterization of these systems.<br />

It should be remembered that the above discussion has been presented by considering<br />

only one single defect embedded in a host polymer chain. When a distribution<br />

of defects is considered, defects may occur close enough to couple mechanically.<br />

Splitting and changes of the gap modes (frequencies and intensities) may occur and<br />

are indicated by the calculations (see, for instance [ 1001).<br />

3.11 A Very Simple Case: Lattice Dynamics of<br />

HCl-DCl Mixed Crystals<br />

It is considered that the simplest possible realistic case on which our theories can be<br />

tested, and which could act as a guide to the reader for the discussions which follow,<br />

are the studies of mixed crystals of HC1 and DCl [97]. Solid HCl and DC1 undergo<br />

a first-order phase transition at 98.4 OK, and 105.0 OK, respectively. Below the transition<br />

temperature, X-ray and neutron diffraction studies show the existence of an<br />

ordered face-centered orthorhombic structure which contains planar zigzag chains<br />

consisting of HCl (DCl) molecules linked to each other by hydrogen bonds (for a<br />

list of references on the physical properties of solid HCl, see [97]). Our model of the<br />

chain consists of a pure infinite isolated chain of the type ... H-Cl.-.H-Cl...H-Cl...<br />

(Figure 3-17). Using the techniques presented in the previous sections we have<br />

Figure 3-17. Geometry and internal<br />

coordinates (in-plane and out-of-plane) for<br />

crystalline HCI or DC1 in the orthorhombic<br />

-<br />

modification. Only one single chain has been<br />

- +<br />

considered. =z - + + -


132 3 T/ihrtitional <strong>Spect</strong>ra us LL Probe of Structural Order<br />

a<br />

Figure 3-18. Dynaniical data<br />

for the perfect chain of HCl<br />

and DCl based on a valence<br />

force field. (a) Dispersion<br />

curves; (b) k = 0 normal<br />

modes; (c) densities of<br />

vibrational states.<br />

calculated dispersion curves (Figure 3-l8a), density of vibrational states g( 11) (Figure<br />

3-l8b) and vibrational displacements for k = 0 phonons both for pure HCI and<br />

DCI (Figure. 3-18c). Vibrational infrared intensities have also been calculated based<br />

on a simple fixed point-charge model [97].<br />

As a second step, we generated models of chains of HCl containing defects of<br />

DC1. Calculatioiis are made easier because isotopic substitution does not change the<br />

force field. Several mixtures were considered and we will focus here ody on three<br />

cases of mixed chains with 3%, 25% and 45Y” DC1 in HCI. Chains were made up by<br />

600 chemical units.


3 I1 A Yey) Siiiiple Case: LLittice Dynnniics of HCl-DCl Mixed Cvvstu1,s 133<br />

Figure 3-18 (b)<br />

b<br />

Suitable experimental procedures were envisaged in order to obtain such materials<br />

and in order to record the infrared spectra to be compared with the calculated<br />

ones.<br />

Figure 3-18a immediately indicates that the HC1 stretching of molecules isolated<br />

in a DCl lattice can generate localized out-of-band modes, while DCl stretching in<br />

a HCl host lattice generate typical localized gap-modes. This can occur for other<br />

lower-frequency modes for both molecules. The 03-co4 branches span frequency<br />

ranges with partial overlapping between HCl and DCI lattices, thus possibly generating<br />

resonance modes. Also, the vibrations of 07-Wg may couple. No coupling<br />

can occur between the out-of plane phonons of HC1 (05-06) and the in-plane<br />

phonons of DC1 (w3-014) since they are orthogonal.<br />

The histograms calculated for the three isotopically impure chains are given in<br />

Figure 3-19 and show a complicated and almost inextricable pattern. Some details<br />

are enlarged in Figure 3-20.


134<br />

3 I 'ibrutioiwl S'Jectra us LI Probe of' Structud O~tler<br />

y (w.<br />

i 56<br />

I<br />

HCf<br />

100<br />

50<br />

C<br />

Figure 3-18 (c)<br />

The object of the next step is to apply IIM to identify the origin of the various<br />

peaks calculated in the g(v) of the mixtures. First, we consider the random distribution<br />

of 25"h DCl defects in a host lattice of HCI (Figure 3-21); the total number<br />

of units in this calculation is 100. Let Z be the number of chemical units within the<br />

defect. Moreover, for sake of completeness, let us also associate qualitatively to<br />

each motion its corresponding dipole moment changes. Let us assume that each bar<br />

in Figure 3-21 gives not only the vibrational amplitude, but also the dipole moment


3.11 A Vq> Siiiiplc Crrw: Luttice Djwmnics of HC1-DCI Mixed Crystals 135<br />

100<br />

glw i<br />

50<br />

0<br />

Figure 3-19. Density of vibrational<br />

states for a single mixed chain of (a) 3%.<br />

jb) 25%, and (c) 45% DC1 in HC1 (chains<br />

of 600 chemical units)<br />

0 100<br />

200 300 LOO 500<br />

Figure 3-20. (a) Experimental<br />

infrared spectrum in the HCI<br />

and DCl stretching regions of<br />

a mixed HCI/DCI crystal<br />

contaiuiiig 25'X of DC1; (b)<br />

g(i8) and (c) dipole-weighted<br />

g[ i t) assuming a Lorentzian<br />

band shape with Ai8,p = 2<br />

cm-'. The observed spectra<br />

have been shifted as indicated<br />

for sake of qualitative coinparison.<br />

A.<br />

L<br />

-A-<br />

19.33 1950 1970<br />

A<br />

2ioo 2i20 2i~o 27ko


136 3 Vilwtrtionrrl <strong>Spect</strong>ra (1s N Probe of Striictiirul Orrlev<br />

I<br />

/I/<br />

11 c)<br />

‘1 ‘1 w=2729.5 cm-l<br />

di<br />

I“ w=270u.5 cm-l<br />

el<br />

Figure 3-21. Sample eigenvectors<br />

in the HC1 stretching region for a<br />

single chain of mixed 25% DCI in<br />

HC1. The top line gives the distribution<br />

of defects. Vibrational displacements<br />

are rotated of 90” for<br />

motions.<br />

changes. The vectorial sum of the dipole moment changes is related to the absorption<br />

intensities. We can then qualitatively predict also the intensity of each of the<br />

calculated modes. The distribution of defects given at the top of the figure indicates<br />

the existence in the chain of 14 DCl isolated units (Z = l), 14 doublets (Z = 8) +1<br />

triplet (Z = 3). Application of NET, IIM, and intensity coinsiderations indicate that<br />

at 1’ = 1948.76 cn-l only the single DCl niolecules as isolated entities perforin fully<br />

localized vibrations at the site where they are located with strong intensity. At<br />

v = 1948.5 and 1948.68 cni-’, the four single units separated only by one HCl<br />

spacer feel some intramolecular coupling and generate localized modes with some<br />

phase coupling, the former has strong intensity and the latter zero intensity. Next,<br />

at 1’ = 1941.5 cm-’ only isolated doublets generate an in-phase motion (strong) of<br />

the two molecules. The out-of-phase motion occurs at Y = 1956.5 cm-] (inactive).<br />

Finally, the motions characteristic of the triplet occur at 1’ = 1938.5 (strong) and<br />

1959.9 cn1-l (weak). It is also apparent that in this frequency range the atoms which<br />

belong to the host lattice do not move (i.e., the motion of the DC1 defect is not<br />

propagated throughout the chain of the host HCl lattice). The concept we wish to<br />

stress here is that the cases just examined are typical examples of modes localized in<br />

space and energy characteristic of particular ‘defects’ or of ‘islands’ or ‘clusters’ of<br />

defects. Moving to the HCl stretching range (near 2700 cm-I) the analysis of the<br />

composition of the chain from localized modes can be completed. For the same<br />

chain considered in Figure 3-22 the distribution of HC1 sequences is the following:<br />

4 (Z = I), 4 (Z = 2), 3 (Z = 3), 2 (Z = 4), 2 (Z = 5), 2 (Z = 6), 1 (Z = 7), 1 (Z = 8)<br />

and 1 (Z = 9). From Figure 3-22, the ‘island analysis’ carried out using the IIM<br />

method allows the following frequencies Characteristic of each island to be located


Figure 3-22. Sample eigenvectors in<br />

the in-plane bending region for a<br />

single chain of mixed 25% DCI in<br />

HCI. The distribution of defects is<br />

the same as that given in Figure<br />

3-21.<br />

dl<br />

Wz360.6 cm-'<br />

unquestionably: Z = 1, 2719.5 cm-' (strong); Z = 2, 2708.5 cm-' (weak or inactive)<br />

and 2729.5 cm-' (inactive); Z = 5-6, 2700.5 cm-' (strong) and Z = 8-9,<br />

2739.5 cm-' (inactive).<br />

An example of resonance modes which cannot provide any signal characteristic<br />

of a given defect is given in Figure 3-23 for the same chain as that considered in<br />

Figure 3-21 and 3-22. In this case the bending modes are considered where (see<br />

above) frequency branches overlap. Because of intramolecular coupling, collective<br />

vibrational waves extending along large domains of the chain are generated and do<br />

not provide any useful and characteristic signal which may contribute with a weak<br />

and broad background.<br />

It is realized that the situation found for impure simple 3D lattices is found<br />

also for the single chain considered here and very likely may be found for other<br />

polymers containing an increasing concentration of defect. In the HCl (or DC1)<br />

stretching region these mixed crystals behave as 'two-mode system' or 'persistencetype'<br />

system. 'Amalgamation-type' behavior is observed in the bending range where<br />

sections of density of states overlap both for HCl and DCl [98].


138 3 Vibrcitional <strong>Spect</strong>ru as a Probe oj Stnrcttiral Order<br />

dl<br />

w=3a1 a cm-‘<br />

el<br />

w=355.0 cm-’<br />

Figure 3-23. Examples of resonance<br />

modes in the bending region for a<br />

single chain of mixed 25% DCI in<br />

HCl.<br />

3.12 CIS-trans Opening of the Double Bond in the<br />

<strong>Polymer</strong>ysation of Ethylene<br />

The method just outlined was used, at the earlier stages of these kinds of study<br />

[loll, for the elucidation of the mechanism of polymerization which was, at that<br />

time, an unsolved puzzle. The chemical problem can be stated briefly in the following<br />

way: when ethylene is polymerized using Ziegler-Natta catalysts a linear<br />

polymer is formed. It was known that the catalysts acts on the double bond of the<br />

ethylene unit and the polymethylene chain can grow on both sides of the ethylene<br />

unit once the double bond has been opened. The opening, however, can occur in cis<br />

or trms configuration.<br />

H H H I H<br />

‘c-c’<br />

‘c-c’<br />

H/I J\H H’ l\H


3.12 CIS-trans Opening @the Double Bond 139<br />

The question which remained to be answered for inore than a decade was<br />

whether the opening of the double bond occurs in cis or trans. The understanding of<br />

this reaction mechanism was of fundamental importance for the development of<br />

other similar reactions for other classes of materials.<br />

One way to solve the problem was to polymerize deuterated derivatives of ethylene<br />

with the same Ziegler-Natta catalysts. If cis di-deutero ethylene (monomer A)<br />

is polymerized and the opening of the double bond is CIS. the resulting polymer<br />

should be rich in the so-called ‘cridzro’ units<br />

HH<br />

c-c 7<br />

DD<br />

c-c-c-c<br />

HDDH<br />

DHHD<br />

a 1 2<br />

If the opening is cis for the case of monomer B (tram-dideuteropolyethylene) the<br />

reaction should yield a polymer rich in the so-called ‘threo’ units.<br />

DH<br />

c-c 7<br />

HD<br />

c-c-c-c<br />

HHDD<br />

DDHH<br />

b 1 2<br />

The opposite should occur if the catalyst forces the double bond to open in trans<br />

configuration.<br />

The polymers were made by chemists, but the identification of the configuration<br />

of the deuterium and hydrogen atoms remained unsettled for a long time. The<br />

vibrational spectra were recorded, but their interpretation was impossible on the<br />

basis of the classical techniques of polymer spectroscopy.<br />

An example of the procedure followed in the identification of the spectra is the<br />

following.<br />

Let the configurations of monomeric units be indicated as follows<br />

HD DH HH DD<br />

c-c- 7<br />

-c-c-<br />

- -c-c- i -c-c- 4<br />

DH HD DD HH<br />

I I1 111 IV<br />

From the viewpoint of calculations a computer program which generates a sequence<br />

of 1-andoni numbers was used in order to generate a random sequence g( 1’)


c<br />

40’ r<br />

<strong>Polymer</strong> 8 (100) .?<br />

Wave number (cm-’<br />

a<br />

Figure 3-24. (a) Calculated<br />

density of vibrational slates of<br />

polymers A and B with 100<br />

monomeric unit. (b) Observed<br />

infrared spectra of poly(ci.v-<br />

CHD=CHD) and poly(rrrrris-<br />

CHD=CHD).<br />

for both chains with 50150 concentration of a and b. The sequence is the following:<br />

bbaabbaabbbbaabaaabababa<br />

bbbbaabaaabbbabbbbaaabba<br />

aaabababbbaabababbbaabaa<br />

ababbabbbabbaaababaaaabbb<br />

The two cases considered here are: (i) polymer A, generated by replacing a with unit<br />

I and b with unit 11; and (ii) polymer B, obtained by replacing a by unit 111 and b by<br />

unit IV.<br />

Each chain contains 600 atoms. The corresponding GR and FR matrices (of dimension<br />

1800 x 1800) were constructed and NET applied for the calculation of<br />

got). The calculated g( v) (Figure 3-24a) compared with the experimental infrared<br />

spectra (Figure 3-24b) clearly shows that the calculated g(v)(A) corresponds to


3.13 Defect Modes cis Structtrr.al Probes in Poljw~eihylriie Chins 141<br />

0<br />

.-<br />

c<br />

I , / /<br />

, I / /<br />

Figure 3-24 (continuedj<br />

I l l , ,<br />

7350 1300' ' I ' 7250 '<br />

Wave number Icm-')<br />

' 1200<br />

1<br />

b<br />

the infrared spectrum of poly(cis-CHD=CHD) and g(v) (B) corresponds to the<br />

experimental spectrum of poly( trans-CHD=CHD).<br />

It can then be concluded that the polymerization by Ziegler-Natta catalysts<br />

occurs with the cis opening of the double bond.<br />

3.13 Defect Modes as Structural Probes in<br />

Polymethylene Chains<br />

The strongest effort by our group and by other authors was aimed at revealing the<br />

structure and molecular dynamical evolution of the very many classes of technologically<br />

and scientifically relevant molecules consisting of n-alkanes or n-alkyl<br />

long chains. We report below some of the results obtained and will show some<br />

applications.


3.13.1 Mass Defects<br />

Let us introduce a CD2 group as a defect in n-alkanes (CH3-(CHl),,CH3) or n-alkyl<br />

chain (CH3(CH?),X or in any polyniethylene chain. This is a typical case of ‘mass<br />

defect’ which leaves the force constant matrix (Eq. 3-31) unchanged because of<br />

Born-Oppenheimer approximation. The site where the CD2 group is located can<br />

be changed along the chain. Chemically, this can be easily achieved and selectively<br />

deutei-ated n-alkane or alkyl derivatives are either coininercially available or can be<br />

(and have been) easily synthesized.<br />

Calculations have shown that the frequency of the rocking motion of the unit<br />

CD? embedded in a CH2 host lattice changes by changing the conformation of<br />

its surroundings [102, 1031. The calculated and observed frequencies for the CD?<br />

rocking motion when the adjacent torsional angles are changed are reported in<br />

Table 3-1. These motions are classical well-localized gap-inodes as they occur in the<br />

frequency gap between the CH2 rocking and the CCC bending dispersion branches<br />

of the host lattice.<br />

Table 3-1. Calculated and experimental CD2 ’gap frequencies’ for selectively deuterated nonadecane<br />

as function of the torsional angle about C-C bonds.<br />

Calculated<br />

Observed<br />

4 3 2 1<br />

-CH2-CH>-CH>-CD>-CH3<br />

T T<br />

G T<br />

T G<br />

G G<br />

615 622<br />

615 632<br />

658 658<br />

654<br />

13 12 11 10 9 8 7<br />

CH2-CH2-CH2-CD2 -CH: -CH:-CH?<br />

T (T T) T<br />

G (T T) T<br />

G (T T) G<br />

G’ (T T) G<br />

T (T G) T<br />

G (T G) T<br />

G (T G) G<br />

G‘ (T G) T<br />

T (G G) T<br />

T (G G) G - 671<br />

G [G G) G<br />

-615 622<br />

- 650 650


3.13 Defect Mo&s cis Structurcil Probe5 iii Poljwietliylene Cliaiiis 143<br />

3.13.2 Conformational Defects<br />

The study of the conformational defects in polymethylene chains has been tackled<br />

some time ago in a systematic way by the school of Snyder by norinal inode analysis<br />

of short-chain molecules (see, for exainplc (1041) and by our group at Milano<br />

using NET [95. 1011. The results by both techniques are very similar. In the previous<br />

discussion in Figure 3-14 we have pointed out that a few peaks in the calculated<br />

g( 1’) for infinite tims polyethylene do not find corresponding infrared and/or<br />

Ranian lines to be assigned to k = 0 motions. These additional experimental peaks<br />

are just an evidence of the existence of extra structures which we can identify as<br />

conformational defects using the theory we are presently discussing.<br />

The main characteristic defect modes calculated (and observed) for tram planar<br />

segments are the following:<br />

1. GG defect. The wagging vibration of a CH2 group located between two adjacent<br />

grrirclie conformations -CH-,(G)-CH?(G)-CH-,- wagging. The calculated (and<br />

observed) frequency lies near 1355 cin-’.<br />

2. GTG‘ defect. CH2 wagging. -CH2(G)-CH,(T)-CH-,(G’)-. This defect introduces<br />

two wagging modes because the two CH2 groups are joined by a bond in<br />

traizs conformation and are partly decoupled from the host lattice by the gauche<br />

bonds. The intra-defect coupling generates out-of-phase and in-phase modes<br />

calculated (and observed) near 1375 and 1306 c1ii-I.<br />

3. GGTGG dc+x-t. This defect is the one which possibly exists on the surface of the<br />

single crystal lainellae of polyeythylene if the chain re-entry forins a tight fold<br />

along the [200] crystallographic plane [ 1051. This defect generates practically the<br />

three lines expected for GG and GTG’ defect in the CH-, wagging region and an<br />

additional quasi out-of band mode (near 714 cm-I) just outside the k = 0 limiting<br />

rocking mode of the CH2 group [99].<br />

4. Heud-to-hecicl defects of the type. -RCH-CH-,-CH-,-RCH-. The CH2 rocking<br />

mode of a sequence of two CH-, groups between two heavy boundaries is calculated<br />

and observed near 850 cm-’ in agreement with empirical correlations<br />

derived from the study of ethylene-propylene copolymers [ 1061.<br />

5. Clzaivi end modes it2 pi-alkyl clznins. When the group -CH-,-CH>-CHl-CH3 in<br />

such chains has TTT conformation, the methyl ‘umbrella deformation’ iU) is<br />

calculated and observed near 1375 c1n-l and the external deforniation mode 6<br />

CH3 occurs near 890 cm-’. According to the previous discussion in this chapter<br />

both motions should be considered end-group modes whose frequency should<br />

remain constant when the length of the chain increases. This is indeed the case if<br />

the chains keep the all-trams structure. However, since both motions occur in a<br />

frequency range spanned by the dispersion cui-ves of the 1D lattice they couple<br />

with the vibrations of the bulk and their coupling extends for a few CH2 units<br />

along the chain. If the conformation of the chains near the ends changes, such<br />

intramolecular coupling changes and the frequencies of the ‘end groups’ may<br />

change. Table 3-2 reports the values of the calculated and observed frequencies<br />

when chain ends are conformationally distorted in the TG and GT conformations.


144 3 Vibrutionnl Spcctrtr as N Prohc OJ Structurnl Order<br />

Table 3-2. Characteristic calculated and expeiimental ‘in band’ frequencies fot the conformers<br />

-CH~’CHL’~’CH~-CHI and -CHP(~)CH~(~’CH~-CH~ at the end of the n-nonadecdnr molecule<br />

v (cm-’)<br />

Calculated<br />

Observed<br />

TG 1348 1342 (IR)<br />

879 866 (R), 866 (IR)<br />

769 766 (IR)<br />

GT 1367<br />

851 844 (R), 845 (IR)<br />

3.14 Case Studies<br />

3.14.1 Conformational Mapping of Fatty Acids Through Mass<br />

Defects<br />

It becomes apparent that the use of selective deuteration becomes a powerful tool<br />

for the mapping, site by site, of the coiifoiniation of polymethylene chains. This has<br />

been applied for the conformational mapping of n-alkane and of fatty acids. The<br />

idea was first proposed by Snyder and Poor for n-alkanes [lo21 and was later<br />

applied to the case of fatty acids [lo71 and has settled the issue on the conformation<br />

in the solid state which has long been matter of great debate [108].<br />

Fatty acids are typical of a class of compounds for which the existence of premelting<br />

phenomena has been claimed with the formation of ‘liquid-like’ structures<br />

for the molecules while still in the solid. The structure of these systems has been<br />

studied by many workers. The molecules in the solid are dimeric since pairs of carboxyl<br />

groups face each other and are held together by hydrogen bonds. Moreover,<br />

complex polymorphic phenomena have been revealed depending on the way crystallization<br />

is carried out (from the melt, from solution, froin the kind of solvent,<br />

etc.) [107, 1091. More recent works based on the study of LAM inodes in the<br />

Raman suggested that the molecules of fatty acids, independent of the crystalline<br />

form, are never all-trans planar and contain somewhere along the chain a gauche<br />

distortion [108]. These information were at odds with our studies on LAM modes of<br />

the same materials which were shown to be fully trans-planar [110, 11 I].<br />

The use of defect gap modes using selectively deuterated materials have definitely<br />

solved the problem. We have recorded the infrared spectra of solid perhydrogenated<br />

stearic and palmitic acids and of a few selectively deuterated derivatives.<br />

For palmitic acid, single CD2 group were placed in position 2, 3, or 4 (carbon atom<br />

1 is that of the carboxyl group); for stearic acid, the single CD? group was placed in<br />

position 2, 3, 4, 5, 6, 7, or 13. As an example of the way we proceeded, Figure 3-25<br />

illustrates the infrared spectra of: (i) normal undeuterated palmitic acid (Figure


0<br />

3.14 Case Studies<br />

N<br />

0<br />

d<br />

- Y<br />

N<br />

o m<br />

orm<br />

v<br />

I<br />

0<br />

U<br />

c-<br />

.-- - 0<br />

I I I I<br />

ln<br />

N<br />

0<br />

m<br />

'9<br />

0 0<br />

e<br />

Y<br />

V<br />

0 L


146 3 Vibratioiiul Sycctm CIS (I Probe of Structural Ordcr<br />

3-25a); (ii) palinitic acid ?-Dz (Figure 3-25b); and (iii) pahnitic acid 3-D2 (Figure<br />

3-25~1. The clear band which appears near 620 cn-' is the gap-niode indicating<br />

unquestionably the existence of the local conformation CH?(T)-CD?IT)-CH? both<br />

in positions 3 and 4. Identical spectra have been observed along the whole chain,<br />

thus confirming the whole lrms structure of palinitic acid in the solid state [107,<br />

110, 1111.<br />

3.14.2 Molecular Mobility and Phase Transitions in Therrnotropic<br />

Liquid Crystals from the <strong>Spect</strong>roscopy of Defect Modes<br />

The issue is the description of the molecular changes accompanying the phase<br />

transitions in systems which show liquid crystalline phases. While much is known<br />

on the morphological changes, little is known on what happens to the molecular<br />

structure through the various phase transitions.<br />

We use defect mode spectroscopy for aiming at the structural changes of prototypical<br />

thermotropic liquid crystals such as alkylcyanobiphenyl and polyesters.<br />

3.14.2.1 Case 1: Cyano-alkyl biphenyls<br />

Let us first consider dodecyl-cyanobyphenyl which shows two phase transitions at<br />

48 "C (crystalline-semectic A) and at 58.5 "C (smectic A-isotropic). The analysis<br />

of the whole spectrum is reported in the original work [112] and we focus here only<br />

on the variation with temperature of the conforiiiational structure of the dodecyl<br />

side chain which we hope to reveal with the study of the temperature-dependent<br />

vibrational spectrum in the 1420-1280 cm-' range where defect modes are expected<br />

to occur.<br />

Figure 3-26 reports such spectra, together with the difference spectra obtained in<br />

the following way. Let s(J + 1) be the spectrum recorded at temperature T(J + 1)<br />

and s(j) the spectrum taken at temperature T(j). Then AT = T(J + 1) - T(j) is the<br />

temperature gradient in our experiment. From Figure 3-26, while drastic changes<br />

are observed from the crystalline to smectic A phase, the spectra do not change<br />

much from sniectic A to isotropic. The difference spectra highlight such changes<br />

and allow the defect modes GTG', GTG, GG, and end-TG to be identified. The<br />

following conclusion can be easily reached: (i) the trans-planar chain exists up to the<br />

crystal-smectic A transition; (ii) at the crystal-smectic A transition the trans-planar<br />

chain collapses in a conforinationally distorted geometry, as indicated by the appearance<br />

of all defect modes listed above; (iii) the observed upward shift of the<br />

umbrella deformation mode of the CH3 group from 1376.5 to 1378 cm-l indicates<br />

that the environment at the end of the chain is changed; and (iv) no great conformational<br />

changes occur from the sniectic A to the isotropic phase; the cyanobiphenyl<br />

core must then supervise this phase transition (as also shown by the<br />

vibratioiial spectrum). Another case treated is that of 4-cyano, 4'octyloxy biphenyl<br />

[113].


3.14 Case Studies 147<br />

OC<br />

60<br />

58<br />

57.<br />

56<br />

55.<br />

54<br />

52.<br />

50<br />

49<br />

47<br />

a.<br />

43<br />

41<br />

39<br />

36<br />

33.<br />

21<br />

Figure 3-26. Dodecylcyanobiphenyl. (a) Temperature-dependent infrared spectra in the defect<br />

mode frequency range. (b) Experimental spectroscopic evidence from difference spectroscopy that<br />

at the phase transition crystal - sniectic the conformation of the dodecyl residue changes froin all-<br />

~YLIII.F to 'liqnid-like' (the most connnon defect modes appear clearly). From sinectic to liquid phases<br />

no drastic confoiiiiational changes are observed (see text).<br />

3.14.2.2 Case 2: Liquid Crystalline <strong>Polymer</strong>s: Polyesters<br />

We consider the case of the polyiner -[(CH2)lo-00C-C~H~-00C-C6HJ-COO-<br />

CcH4-COO],- (hereafter referred as HTH-10) whose phase transitions are the following:<br />

K-S 221 "C; S-I 260°C. From the overall aiialysis of the temperature-


148 3 ViLwntional <strong>Spect</strong>ra us N Probe of Striictural Order<br />

-<br />

5_1_<br />

L<br />

60-513 5<br />

- 5135-575<br />

575-565<br />

-<br />

-<br />

_v_ -<br />

--<br />

56 5-54<br />

54-526<br />

52 6-50 5<br />

505-49<br />

-49 - 47 1<br />

47.1 - 45.5<br />

43-41<br />

004.-<br />

-<br />

41-39<br />

a,<br />

C<br />

0<br />

__c.<br />

'?<br />

39-36<br />

2 002- 1<br />

n<br />

- ---36-33 1<br />

a<br />

-3.1 -27<br />

AT<br />

0.00 t<br />

Figure 3-26. (continued)<br />

dependent Raman and infrared spectra [I 141 it could be concluded that, in the K<br />

phase, most of the decamethylene segments are in the trans-planar conformation<br />

and that upon heating gnuclze rotamers are generated. The identification of what<br />

kind of conformational defects occur in the disordering process when the temperature<br />

increases can only be attempted by the study of the temperature-dependent<br />

spectrum in the traditional CH2 wagging range. We restrict here our analysis to<br />

the temperature-dependent spectra in the usual defect modes range from - 1400 to<br />

- 1300 cm-'. We aim at revealing the conformational evolution through phase<br />

transition of the segments consisting of 10 CH-, groups. Figure 3-27 illustrates the<br />

temperature-dependent infrared spectrum in the CH2 wagging range plotted in


3.14 Case Studies 149<br />

GI A V E N U M BE R 5<br />

Figure 3-27. Liquid crystalline polyester with decamethylene spacers. Temperature-dependent<br />

infrared spectrum in the defect modes frequency range (see text). The spectra and shifted one from<br />

the other as in a tridimensional plot.<br />

three dimensions for the sake of clarity. It becomes apparent that the tvnizs-planar<br />

structure (no defect modes in this frequency range) evolves with increasing teniperature<br />

with the generation of only the defect mode near 1365 cm-' due to GTG'<br />

defects. No absorption due to GG defects (band at 1353 cm-') is observed, i.e.,<br />

sharp kinks are neither generated in the sinectic nor in the isotropic phases. The


150 3 T’ibratioiinl <strong>Spect</strong>ra L ~S u Probe oj StrzictLir.al Order.<br />

intensity of the GTG’ defect inodes increases sharply at the K-S transition, remains<br />

on a smooth platem until the S-I transition occurs, and increases steeply in the<br />

I phase. It has to be noted that even in the I phase the GG defect mode is still<br />

practically not observable.<br />

The mechanism on a molecular level which is derived from the vibrational spectra<br />

can be suniinarized as follows: at room temperature the sample contains a sizable<br />

concentration of ordered material in which the decamethylene spacers are<br />

predominantly in the trms-planar structure. When the I(-S transition is<br />

approached the main spectroscopic signal observed is that associated to the GTC’<br />

modes. It is known that the GTG’ defect introduces a kink in the chain but does not<br />

change the trajectory of the polyniethylene sequence keeping both arms at either<br />

side of the defect parallel. The overall effect is that the length of the decaniethylene<br />

segment simply shrinks. It follows that in the S phase the spacers shrink because of<br />

the introduction of conformational kinks, inducing a sort of shearing motion between<br />

the large aromatic mesogenic groups. The lack of any sizable concentration<br />

of GG structures even in the I phase leads to the interesting conclusion that the<br />

molecular structure in the I phase is not much different from that in the S phase. No<br />

indication is found of ‘liquid-like structures’ observed in other polyniethylene systems<br />

in the liquid phase. It has been proposed [114] that one way to reconcile<br />

the observation of macroscopically isotropic phase with the observation on the<br />

molecular level is to envisage that the isotropic phase consists of a collection of<br />

microdomains whose directors are randomly oriented. However, inside the microdomains<br />

polymer molecules are organized as they are in the S phase.<br />

3.14.2.3 Case 3: Chain Folding in Polyethylene Single Crystals<br />

It is known that chain folding occurs in polyethylene crystals; the fold structure,<br />

however, has been a subject of interest and controversy [105]. It was obvious for<br />

polymer spectroscopists to use vibrational infrared and/or Raman spectra to try to<br />

contribute to the understanding of the shape of the fold (i.e., the molecular conformation)<br />

within the fold at the surface of polyethylene single crystals.<br />

The large amount of spectroscopic work carried out by many authors is<br />

sumarized in [115]. Here, we wish to focus specifically on the contribution given by<br />

defect mode spectroscopy considering both calculations and experiments. The issue<br />

is whether the surface of polyethylene single crystals consists mostly of loosely<br />

looped folds or also of a fraction of chains with a tight adjacent re-entry. In the<br />

former case, ‘liquid like’ defects such as GG, GTG, GTG’ etc. are expected and can<br />

be recognized in the vibrational spectrum. If tight fold re-entry exists defects of the<br />

type GGTGG must occur at the surface of the lainellar single crystal.<br />

The first question which needed to be answered in a less qualitative way was<br />

where to locate the vibrations localized on the GGTGG defect.<br />

Calculations have been carried out on a long polymethylene all-tivm host molecule<br />

containing evenly spaced GGTGG defects. NET and IIM and the ‘island<br />

analysis’ were applied [99]. The most meaningful results are presented below.<br />

Two highly localized modes are calculated to lie near 1361 and 1358 cin-’ and


3.14 Case Studies 15 1<br />

correspond to the out-of-phase motions of the two CH, groups located between two<br />

C-C bonds in gcruchc conformation. These two modes are largely localized at the<br />

defect while the other portion of the chain remains practically still. It is interesting<br />

to note that the defect modes in the CH2 wagging region near 1350 cm-' show that<br />

in this spectral range the GGTGG defect behaves as if it were made up of GG and<br />

GTG defects moving almost independently.<br />

In the lower-frequency region defect modes are calculated near 1105, 900, and<br />

714 cm-'. The mode near 714 cni-' is of particular interest since it lies just below<br />

the limiting k = 0 mode located at 720 cn-l of the CH2 rocking dispersion curve<br />

and is just at the edge of a large energy gap which extends as far as the cut-off of the<br />

branch cu5 (see Figure 3-3). From IIM, we learn that the defect mode near 714 cmP1<br />

originates from a cluster of several CH, groups, including the defect, which is performing<br />

mainly a rocking motion. From the shape of the nornial mode this nearthe-edge<br />

mode cannot yet be considered a localized gap mode since it shows a<br />

nonnegligible degree of cooperativity. Similar cases of near-the-edge resonances<br />

were found for polyethylene in the lower frequency range and for polyvinylchloride<br />

containing head-to-head defects [90].<br />

The modes calculated near 1105 and 900 cm-l are the results of a complicated<br />

mixture of several internal coordinates, but can be described mainly as C-C resonance<br />

modes. They are likely to be observed in the Ranian spectra.<br />

The results of the calculations were checked with the experimental infrared and<br />

Raman spectra of the cyclic hydrocarbon C34H68 whose structure determined by<br />

Newiiian and Kay [116] indicates that two segments of 12 CH? groups in a transplanar<br />

zig-zag conforination are joined by two GGTGG defects. Coincidences<br />

between experiments and calculations are striking. However, a critical analysis<br />

needed to be made ([99, 1151) on whether these signals are characteristic only of the<br />

GGTGG defect and may allow such a defect to be disentangled from the signals of<br />

other conforniational defects possibly existing in a realistic polyethylene single<br />

crystal. The conclusion reached is that calculations on several model systems and<br />

experiments on C34H68 suggest [99] that a simultaneous appearance in the infrared<br />

spectrum of two bands near 1342 cm-' and one near 700 cm-' (certainly below the<br />

in-phase limiting CH2 rocking mode of the tram-planar chain) can be taken as evidence<br />

for the existence of GGTGG defects in polyethylene single crystals. The other<br />

calculated modes occur at frequencies where other modes of the host lattice may<br />

also occur. Calculations were also extended to predict the intensity pattern for<br />

the defect modes in the CH2 wagging range [117, 1181. Figure 3-28 compares<br />

the experimental spectrum with that calculated using the force field proposed by<br />

Shimanouchi [ 1 191.<br />

Experiments of difference spectroscopy combined with band deconvolution processes<br />

were presented and conclusions were not unambiguous [ 1 151. On the other<br />

hand, the arbitrary decisions taken when applying band deconvolution techniques<br />

sometimes make issue more complex. We have tried to avoid band deconvolution<br />

and proceeded directly to careflil experiments of difference spectroscopy on sectored<br />

polyethylene single crystals. The principle on which the experiments was based is<br />

the following. From the study of sectored polyethylene single crystals it is known<br />

that sectors correspond to regions where molecular chains are crystallized in the


1380 1340 1302<br />

Figure 3-28. Coinparison between the calculated and esperimental<br />

infrared spectrum in the CH2 wagging region of the<br />

GGTGG conformational defect. Calculations have been made<br />

the with force field by Shiinanouchi [I 191. The intensity<br />

cm-'<br />

paraiiieters were developed in the laboratory of the author.<br />

{ 1 lo} and { 100) crystallographic planes. By changing the crystallization conditions,<br />

sectored crystals with varying ratios between the { 110) to {loo} sectors can<br />

be obtained. Chain-folding studies has also indicated that tight-fold re-entry with<br />

GGTGG defects can only occur when chain crystallize in the {loo} planes.<br />

Sectored crystals were kindly provided to us by Professor A. Keller with the following<br />

% of { loo} faces: 0, 31, 35,45, and 55. The experiment of difference infrared<br />

spectroscopy we attempted is sketched Figure 3-29. We aimed at isolating the absorption<br />

to be associated with the GGTGG possibly existing on the surface of the<br />

{loo} fraction of the sectored crystals. The experiments were not easy and after<br />

many attempts the final spectrum obtained (Figure 3-30) is compared with that of<br />

the cyclic hydrocarbon C34H60. The matching of the two spectra is striking and does<br />

suggest that, in agreement with calculations, on the surface of polyethylene single<br />

crystals obtained in particular crystallization conditions tight fold re-entry with<br />

GGTGG defects may indeed occur.<br />

3.14.2.4 Case 4: The Structure of the Skin and Core in Polyethylene<br />

Films (Normal and Ultradrawn)<br />

In addition to the defect modes presented in the previous sections it has been shown<br />

recently by both calculation and experiments that generally when yazrclze defects<br />

exist in polyethylene a broad band arises in the CH2 bending range near 1440 cm-'<br />

[ 1201. The band is very broad, possibly because it originates from a distribution<br />

of nonequilibrium gauche structures constrained by the sample morphology (Figure<br />

3-31). Close to such absorption one observes for crystalline orthorhombic n-alkanes<br />

and polyethylene the doublet due to crystal field splitting (see Section 3.6). Let a<br />

and b label the two components observed near 1473 and 1463 cm-' respectively<br />

(Figure 3-31). It is known that the setting angle B between the planes defined by the<br />

tr.arzs-planar chain skeleton in the lattice is approximately 42" [121]. Such value of<br />

0 determines the intensity ratio IJIb = 1.2333 between the two components of the<br />

crystal doublet of the CH2 bending and CH2 rocking in orthorhombic n-alkane and


3.14 Case Studies 153<br />

(I10 + 200) +BULK y (BULK)<br />

a<br />

(I10 .ZOO)<br />

SURFACE (110)<br />

b<br />

(200)<br />

Figure 3-29. Scheme of the experiments of diference spectroscopy on sectored single crystals of<br />

polyethylene. (a) single crystal - bulk = (1 10) + (200) surface of the sectored crystals; (b)<br />

( 110) + (200) surface - ( 1 10) surface = (200) surface.<br />

Figure 3-30. Comparison between the infrared<br />

spectrum in the GGTGG defect mode region of<br />

the crystalline cyclic molecule C34H68 and the<br />

result of the digerence spectroscopy sketched in<br />

Figure 3-29.<br />

I<br />

ll00 1375 1350 1325 1300 1275 1250 1225 1200<br />

WAVE N UM BER5


154 3 Vihtional <strong>Spect</strong>vu as a Probe of Structurd Order<br />

1500 1460 1440 1420 1 uoo<br />

Wa v e n urn be r s<br />

Figure 3-31. Infrared absorption spectrum of a coininercial film of polyethylene before and after<br />

subtraction ofthe spectrum of the defect mode component near 1440 cm-'. After subtraction of the<br />

defect mode the ratio between compoiient a and b of the correlation field splitting approached the<br />

ideal value for a setting angle of42" (see text).<br />

polyethylene [122]. Certainly the wing at the higher frequency side of the defect<br />

mode overlaps with at least line b of the doublet, thus adding more intensity to line<br />

b and removing any physical meaning to be observed intensity ratio between the<br />

two components. However, by difference spectroscopy it is possible to remove the<br />

absorption due to the conformationally irregular fraction.<br />

This observation is the basis for the determination of the amount of material with<br />

pzdw structure in any sample consisting of polymetliylene chains partially crystalline<br />

and partially conformationally disordered amorphous.<br />

This method has been applied to experiments of multiple internal reflection infrared<br />

spectroscopy on thin films of commercial polyethylene [120]. It is known that<br />

'optical microtomy' can be achieved by changing the incidence angle of the inconiing<br />

beam. Structural depth profiling has then be performed thus allowing to probe<br />

the ordered-disordered structure of such films (Figure 3-32).


3.14 Case Studies 155<br />

1500 1480 1460 1 quo 1420 ! 400<br />

Wavenurnoer<br />

Figure 3-32. Multiple internal reflection spectrum in the CH2-bending region at different angle of<br />

incidence of a 56 pi-thick film of commercial polyethylene. The spectrum shows the increase of the<br />

fraction of conformationally irregular material by increasing penetration depth.


156 3 Vibratioiicil Sr,ectru CIS a Probe of Stnictiiid Order<br />

It has been found that the outer skin of the film is much more crystalline (-SO(%)<br />

and the crystallinity decreases when probing inside the film reaching a concentration<br />

of -60%) crystallinity approximately 5-8 pm inside the film [130].<br />

The above analysis combined with defect mode spectroscopy carried out in infrared<br />

(absorption in polarized and unpolarized light, and multiple internal reflection)<br />

and Raman has been applied to the structural characterization at the molecular<br />

level of ultra-drawn films of high-density polyethylene (UD-HDPE) [ 1331. The<br />

comparative study has allowed to describe the structural situation of UD-HDPE<br />

films as follows:<br />

(i) The skin of the film within - 10 pin is highly crystalline and orthorhombic.<br />

(ii) UD-HDPE film is a multiphase system with the topological distribution of<br />

the phases changing in going from the surface into the core.<br />

(iii) The orthorhoinbic phase exists both on the surface and in the core, while the<br />

monoclinic phase starts at a penetration depth d, > - 10 pm.<br />

(iv) Both the orthorhombic and the monoclinic phases are highly anisotropic and<br />

are oriented along the draw direction.<br />

(v) An additional phase is identified from d, > 10 pni into the core. This phase<br />

consists of sequences of CH2 groups in tvrrrzs conformation. The orientation of<br />

the chains is not parallel to the draw direction. This additional phase exists at<br />

low draw ratios and decreases its concentration qualitatively at higher draw<br />

ratios with respect to the orthorhoinbic component.<br />

(vi) GTG’ kinks are organized in anisotropic domains, seemingly as a result of<br />

collective phenomena of orientation during stretching.<br />

(vii) The chain end methyl groups find themselves in a disordered environment<br />

similar to that found in liquid n-alkanes. The existence of liquid-like droplets<br />

or of disordered domains around the CH3 groups may be envisaged.<br />

(viii) GG defects are observed and may originate from intrinsic liquid-like structures,<br />

from the disorder around the CH3 heads and from some kind of chain<br />

folding between crystallites or within polymer lainellae.<br />

(ix) The concentration of conformational defects is comparatively small when<br />

normal solid polyethylene is considered. The concentration of defects decreases<br />

when the draw ratio increases, thus showing that conforniationally<br />

disordered and coiled chains unwind towards a better alignment and perfection<br />

of the tmns-planar molecules.<br />

All these information have been compared with other physical data whch help<br />

in the understanding of the molecular structure of such technologically relevant<br />

materials.<br />

3.14.2.5 Case 5: Moving Towards More Complex Polymethylene Systems<br />

The role of long alkyl chains in natural phenomena and in industrially relevant<br />

systems is well recognized. Due to the flexibility of the polymethylene segments<br />

these systems show many phase transitions which are not yet clearly understood at


3.14 Case Studies 157<br />

the molecular level. Many authors have focused on the vibrational infrared and<br />

Raman spectra as probes for understanding the structure and dynamics of these<br />

systems.<br />

We have tried to analyze a few cases with our spectroscopy of defect modes and<br />

mention here some of the systems studied.<br />

1. Fattj, acids. Generally, fatty acid crystallize as dimers held together by hydrogen<br />

bonds between the two carboxyl groups which face each other in a head-to-head<br />

arrangement. In other sections of this chapter the contributions to the understanding<br />

of the 'static' structure given by the spectroscopy of mass defects and by<br />

LAM spectroscopy have been already presented. The case of fatty acids with<br />

long polymethylene chains has been studied [ 1241 focusing on the conformational<br />

evolution which originates the phase transitions. FTIR spectra clearly and<br />

distinctly show that when thermal energy is provided to stearic acid in the crystalline<br />

state first interlamellar CH? ...CH3 distances increase (as seen in the case<br />

of n-nonadecane; see Section 3.19) and hydrogen bonds then relax. Simultaneously,<br />

intermolecular distances increase with resulting lateral expansion of the<br />

lattice. These phenomena become relevant approximately 10 " below melting.<br />

Few confonnational kinks are generated below melting and only close to the<br />

melting the concentration of GTG, GTG', and GG defects increases rapidly<br />

bringing the system to a conformational collapse in the liquid phase [124].<br />

However, most of the conformational disorder occurs on the lamella surface,<br />

thus adding more evidence to the existence of a general phenomenon of 'surface<br />

melting' which emerges from our studies on many chain molecules [125]. These<br />

results give a detailed description of the mechanism of phase transition in fatty<br />

acids which was vaguely hinted by NMR experiments [ 1261.<br />

2. Bilajw systenis. The bilayer organic perowskytes were obviously used as simple<br />

models of biological bilayered phospholipids and biomeinbrane systems. We<br />

have studied [127] the system [n-CH3( CH2)"NH3]2MnC14 (hereafter referred to<br />

as C14Mn) which in the solid forms two coexisting, but parallel, layers; a similar<br />

molecule with Zn (C14Zn) in place of Mn forins two interpenetrating or 'intercalated'<br />

layers. The conformational flexibility and phase transitions of these two<br />

classes of molecules turns out to be different. DSC data indicate for C14Mn a<br />

main phase transition in the range 74-85" which is described by our conformational<br />

studies as involving the formation exclusively of GTG' kinks which do<br />

not alter the conformational trajectory of each alkyl stem. No GG defects are<br />

observed and clusters of ordered and disordered chains coexist over a large<br />

temperature range (67-80°C). By simulation of the spectra it is found that, on<br />

the average, at 78°C only one GTG' kink per chain is formed; between 65 and<br />

78 "C chains with only one kink are formed and arrange themselves in clusters. It<br />

follows that disorder proceeds cooperatively throughout the system.<br />

The behavior with temperature of C14Zn is totally different [127]. The main<br />

phase transition at 100°C is accompanied by an abrupt generation of conformational<br />

defects at 99 "C. Above melting, chains have reached almost a liquidlike<br />

structure. Correlations between chains do not exist. However, from the<br />

temperature dependence of the factor group splitting of the CH2 bending and


158 3 Plhrutional <strong>Spect</strong>ru CIS Q Probe qf'StructLirri1 Order<br />

CH2 rocking inodes it is observed that the transition is 'prepared' since the<br />

intermolecular forces between chains in C14Zn [weaker than in C14Mn) quickly<br />

weaken and the thermal expansion gives more freedom for torsional fluctuations<br />

of the alkyl chains, thus leading to conformational collapse.<br />

Two classes of mechanisms which lead molecules to phase transition and melting<br />

have been found, namely: (i) nucleation of conforniational disorder from clusters<br />

whose size continuously increases, thus affecting later all chains; and (ii) cooperative<br />

expansion of the whole lattice caused by an increase with temperature of the mean<br />

amplitudes of torsional fluctuations which reaches a threshold level after which<br />

conformational collapse occurs.<br />

In nature, many other polymethylene systems-also of practical relevance-can be<br />

found which show phase transitions that need to be accounted for. More theoretical<br />

and experimental spectroscopic work needs to be done. We have studied the spectra<br />

of intercalated layered alkylaininonium zirconium phosphates [ 128) and of the solutions<br />

of decylammonium chloride [ 1291. The spectroscopic data present U~~LISLI~~<br />

features with temperature which need further theoretical and experimental studies<br />

using the techniques presented in this chapter.<br />

3.15 Simultaneous Configurational and Conformational<br />

Disorder. The Case of Polyvinylchloride<br />

The cases treated in the previous sections were restricted to the study of the few<br />

cases of mass defects and of the very many cases of conformational defects. Most of<br />

the relevant polymers, however, may contain configurational disorder which carries<br />

as a consequence also some conformational disorder. This case has been considered<br />

within the theoretical approach discussed in this chapter.<br />

The mathematical aspects for a general case of polymers with defects in tacticity<br />

were first worked out, formulae are available and can be applied to any case. We<br />

have applied these mathematical techniques to the case of polyvinylcloride (PVC)<br />

which is known to be strongly configurationally disordered. Our work was done<br />

when the problem of the structure of PVC was highly controversial.<br />

First, the dynamics of ordered PVC was treated in terms of the three energetically<br />

most likely ordered chain structures [go], namely: (i) planar extended syndiotactic<br />

with (-TTTT-) conformation; (ii) helical isotactic, threefold helical structure<br />

(-GTGTGT-); and (iii) folded syndiotactic (-TTGGTTGG-) (Figure 3-12). Phonon<br />

dispersion curves and density of vibrational states were calculated for the three<br />

models using the most reliable force fields available at that time (Figure 3-12). The<br />

infrared and Raman activity of the k = 0 phonons for each perfect chain are the<br />

following: model (i), z = 2, t = 1, 12 atoms per crystallographic unit cell, $(n, d/2),<br />

point group, 9A1 (ir 11, R) +7A2 (R) t9B1 (ir I, R) +7B? (ir I, R); model (iii<br />

z = 3, t = 1, 18 atoms per unit cell, $(2x/3,d/3), point group, 16A (ir 11, R) +


17E (ir 1, R): (iii) z = 4, t = 1, 18 atoms per unit cell, D: point group, 17A (R)<br />

+17B, (ir 11, R) +17B: (ir I, Ri +17Bi (ir 1, Ri.<br />

The knowledge of the dispersion curves, of the spectroscopic activity of the k = 0<br />

phonons and their precise frequencies has allowed us to locate the energy gaps,<br />

where to expect localized defect modes, and to predict where the activation of the<br />

density of states singularities (strong peaks in g(v) with no activity in IR and/or<br />

Ranian) due to lack of symmetry because of disorder could generate extra absorption<br />

or scattering.<br />

Since we take for granted that the actual samples of PVC contain geometrical as<br />

well as configurational defects at non-negligible concentrations, calculations were<br />

extended [90] with the introduction of: (i) conformational defects as isolated units<br />

or as possibly interacting multiple units; and (ii) isolated or possibly interacting<br />

configurational defects. Of necessity, when configurational defects are introduced,<br />

local conformational distortions must be considered so as to justify the existence of<br />

an energetically plausible defect. NET was applied for chains consisting of 200<br />

chemical units (i.e., matrices of 3600 x 3600 were constructed), while IIM was<br />

restricted to shorter chains made up by 50 units (900 x 900).<br />

The distribution of defects was random and the concentrations ranged from an<br />

isolated unit to 8'Yo defects. The construction of the large matrices was made from<br />

smaller submatrices associated to each chemical unit in a given conformation or<br />

configuration. The detailed algebra and the complete expressions for such submatrices<br />

are available in [90] and can be applied to any case of polymers made up<br />

by chiral chemical repeat units.<br />

The attention was mainly focused on the C-C1 stretching range of the supposedly<br />

mostly syndiotactic chain. <strong>Polymer</strong>ization in urea clathrates provides the 'purest'<br />

sample of planar syndiotactic PVC [130]. Such a chain shows a large energy gap in<br />

the CCl stretching range (-650 to 820 cm-'). It is very likely that defect modes<br />

from other structures may generate localized gap modes which may be used as<br />

useful structural probes.<br />

While we refer to [90] for a thorough analysis of the complex structural situation,<br />

in Figures 3-33 and 3-34 we give two examples of the frequency and vibrational<br />

displacements for an isolated GG or an isolated isotactic diad defect in a fillly<br />

syndiotactic planar zig-zag host chain. It must be noted that gap frequencies are<br />

different, but that the vibrational displacements associated to these gap modes involve<br />

many chemical units (-20). The complexity of the problem is shown in Figure<br />

3-35 where the density of vibrational states of a realistic model of configurationally<br />

disordered PVC is compared with the experimental spectrum.<br />

3.16 Structural Inhomogeneity and Raman <strong>Spect</strong>roscopy<br />

of LAM Modes<br />

The problem raised in Section 3.14.1 requires further discussion. From longitudinal<br />

accordion motion (LAM) spectroscopy of a variety of fatty acids Vergottin et 211.


160 3 Vibmtioriul <strong>Spect</strong>ra us a Probe of Strircturul Order<br />

Scale<br />

GG-defect<br />

gap mode<br />

GG<br />

0.05<br />

0<br />

Figure 3-33. Gap mode (described in terms of total displacements of the atoms) for a GG defect<br />

in a syndiotactic PVC chain. Notice that the gap mode is localized over a large section of the<br />

chain.<br />

Figure 3-34. Gap mode (described in terms of total displacements of atoms) for an isotactic isolated<br />

dad. Notice the breadth of the mode which extends over at least 20 carbon atoms.


3.16 Stnictural Znlioi~zogeneity and Ranmn <strong>Spect</strong>roscopy qf LAM Modes 161<br />

Figure 3-35. Density of vibratioiial states for<br />

a realistic model of configurationally disordered<br />

PVC (Bernoullian paramater<br />

P, = 0.47). Comparison with the experimental<br />

spectrum (see text).<br />

[lo81 suggested that the alkyl chains of fatty acids contain a few conformational<br />

defects. This conclusion was shown not to agree with the conformational mapping<br />

side-by-site obtained by defect-mode spectroscopy with the introduction of CD2<br />

defects in the polymethylene alkyl chain.<br />

Such discrepancy seemed either to weaken the relevance of LAM spectroscopy<br />

for the measure of the length of the all-trans segment in a polyethylene chain, or to<br />

cast doubts on the reliability of defect-mode spectroscopy. Some further studies<br />

were in order.<br />

Since dispersion curves have been clearly discussed in Sections 3.5 and 3.6, the<br />

concepts which form the grounds of LAM spectroscopy can be easily defined. Let<br />

us consider an all-tmizs polymethylene chain and focus our attention on the branch<br />

(08 of the phonon dispersion curve. The branch starts as mostly C-C stretching and<br />

ends as CCC bending. For a finite chain of N chemical units we expect to locate on<br />

this branch a band progression of N vibrational modes characterized by the phase<br />

coupling pJ = jn/N. The ‘most in-phase mode’ occurs at very mall frequencies; its<br />

frequency approaches zero when the chain tends to be of infinite length. Let<br />

1,2,3 . . . N label the components of the band progression of the modes which lie on<br />

the Cllg branch.<br />

Schaufele and Shimanouchi [ 13 I] have shown that for centrosymmetric all-from<br />

polymethylene chains the lowest frequency mode performs an accordion-like motion<br />

(LAM) with the atomic displacements describing a wave with one node at the


center of the chain [ 1321. Such a mode has been labeled as LAM 1 mode. The higher<br />

frequencies components have been indicated analogously as LAM?, LAM3, etc.<br />

The odd modes give origin to a strong band progression in the Raman spectrum,<br />

with LAMl extremely strong and the other components with intensity rapidly<br />

decreasing as the number of nodes increases.<br />

Schaufele and Shiinanouchi have considered the longitudinal stretching motion<br />

of an elastic rod of length L, elastic modulus E and density p and have shown that<br />

the frequency of its ‘according’ motion can be written as<br />

(3-53)<br />

The same authors have shown that if E and p are suitably chosen for polymethylene<br />

chains, VLAM~ of tram-planar finite polymethylene chains also follows the<br />

same law. In this case L is the length of the all-miis polymethylene chain which can<br />

be approximated as<br />

L(A) z 1.25(A)(N - 1) (3-54)<br />

where N is the number of CH2 units of the chain.<br />

LAM spectroscopy gained much popularity in polymer characterization and<br />

morphology since it has been possible to measure from Raman spectra the length of<br />

the tmns-planar stems in polyethylene laniellae and in many other relevant cases.<br />

As mentioned above, these concepts were applied in a straightforward manner<br />

to the case of fatty acids by the French group [108], but the results were highly<br />

controversial, thus casting some serious doubts on the applicability of LAM<br />

spectroscopy.<br />

The dynamical problem has been re-examined with the purpose of evaluating the<br />

effects of chain ends on LAM modes. Simplified models were employed which<br />

consider chemical repeat units as point masses joined by elastic springs. Various<br />

cases of perturbations of the LAM modes were considered namely:<br />

(i) The effect of heavier masses at one or at both ends of the polymethylene chain<br />

mimicking systems such as X-(CH2),-X where X = -CH3, F, C1, Br, I, etc. It<br />

transpires that \’LAMI decreases when the mass of X increases [133].<br />

(ii) The effect of joining a polymetliylene chain with another chain with varying<br />

length and with different and variable spring constants. The realistic case<br />

studied was the class of eniifluorinated n-alkanes F(CFZ)~ . (CHZ)~H where N<br />

and M change in a broad range of values. Contrary to expectations, the two<br />

segments do not generate two independent LAMl modes, nor can a simple<br />

heavy mass (X as (i)) at the end of a chain account for the observed spectra.<br />

Calculations show that the molecule vibrates as a whole and generates only one<br />

LAMl mode whose frequency is critically dependent on the length of both<br />

segments. Moreover, the node of such LAMl inode shifts along the chain<br />

depending on the length N and M of the two arms [134].<br />

(iii) The above results opened the way to the study of dimeric systems where the<br />

two monomers are joined head-to-head by springs with varying spring con-


3.1 7 Fermi Resonancrs 163<br />

stants. This model aimed at reproducing the class of dimeric fatty acids or fatty<br />

alcohols. This is the most important case both for its dynamical implications<br />

and for the consequences on structural determination as indicated at the beginning<br />

of this section. The modeling [133] considers two chains with one<br />

heavier mass at one end joined head-to-head by a spring with variable spring<br />

constants. Such a spring mimics the hydrogen bonds which link the two carboxyl<br />

groups.<br />

The results of the calculation indicate that LAMl extends over the whole dimer,<br />

with the node at the center of the dimer between the two carboxyl groups. The frequency<br />

of LAMl has to be searched at much lower frequencies and the frequencies<br />

of the band progression of higher-order LAM modes are totally different from those<br />

identified in the spectra of these materials by the French group. The new band<br />

progression is easily found in the low frequency Raman spectra (Figure 3-36), thus<br />

giving further experimental support to what was shown by defect-mode spectroscopy<br />

(see Section 3.14.1). namely that fatty acids in the crystal do not contain conformational<br />

kinks and are fully trans-planar.<br />

3.17 Fermi Resonances<br />

The next step in OUT analysis is to focus on additional spectroscopic manifestations<br />

of molecular order and disorder which are not only intellectually interesting as<br />

physical phenomena, but also become of extreme importances as markers for the<br />

experimental routine characterization of the very many systems containing segments<br />

of polyinethylene chains.<br />

We are faced with the following experimental facts which need to be understood<br />

and possibly used for analytical and structural determinations. The Raman spectrum<br />

of polyethylene (or practically of any polymethylene chain) in the solid alltrans<br />

structure is very similar (or identical) to that given in Figure 3-37a; upon<br />

melting, the overall appearance of the Raman spectrum is that given in Figure<br />

3-37b.<br />

3.17.1 Principles<br />

The experimental fact is that whenever a molecule contains a few CH? groups the<br />

Ranian spectrum shows a very characteristic pattern of lines in the CH2-stretching<br />

frequency range. Since they are strong they can be observed in a variety of materials,<br />

even under the most awkward experimental conditions. Indeed, they can be<br />

observed for monolayers of organic or biological inaterials spread on surfaces, in<br />

aqueous solutions, in microscopic crystals, etc. Moreover, the spectrum shows<br />

drastic changes in frequencies and intensities when the molecular structure or the<br />

environment changes. For this reason the Raman spectrum is at present a very


164 3 li'hratioiziil <strong>Spect</strong>ra as a Probe oj'StrzictLira1 Order<br />

+<br />

+<br />

+<br />

400 200 0<br />

Frequency (cm-' )<br />

Figure 3-36. Rainan spectra (0-400 cm-' ) of (a) hexatriacontane, n-C36H74; (b) stearic acid,<br />

n-Cl7H35COOH and (c) strearyl alcohol, n-ClgH17OH. Notice that since fatty scids and alcohols<br />

form head-to-head hydrogen-bonded diniers LAM modes extend throughout the whole dirneric<br />

system. Thus, the LAM progression shifts toward much lower frequencies than those expected for<br />

one isolated chain, approaching the values of hexatriacontane. The lowest frequency strong LAM1<br />

mode has its node at the center of the molecule, inside the hydrogen-bonded diineric structure (see<br />

text).<br />

powerful and essential diagnostic tool on the conformatioilally ordered/disordered<br />

structure of complex organic, biological and industrially relevant polyinethylene<br />

materials which contain polymethylene chains.<br />

The reason of such peculiar aid specific spectroscopic manifestations is that<br />

complex Fermi resonance phenomena occur in molecules containing polymethylene<br />

chains. We are concerned in this chapter on the Ferini resonances associated with<br />

the vibrations of polymethylene systems in the trans or gauche states, considered<br />

either as isolated entities and/or when they are organized in ordered and/or disordered<br />

supermolecular structures. Let us remember that Ferini resoiiaiice derives


3.1 7 Femi Resorinnces 165<br />

2<br />

L 1<br />

2900 2800<br />

frcgurlcr I cm-' 1<br />

I<br />

[ , , I , !<br />

3000 2900 28oC<br />

Figure 3-37. Examples of the changes with physical states of the Raman spectrum of polyniethylene<br />

systems: (a) Raman spectrum of crystalline orthorhombic polyethylene in the CH stretching region;<br />

(b) Raman spectrum of polyethylene in the molten state in the same frequency range.<br />

from the mixing, because of anharmonic terms in the potential, of two (or more)<br />

eigenstates (one fundamental and one (or more) overtone( s) or combination( s))<br />

which belong to the same symmetry species. As a result, generally the two (many)<br />

interacting levels repel each other and the generally weak overtone(s) (or combination(s))<br />

borrows (borrow) intensity from the much stronger fundamental [2, 31.<br />

Fermi resonance has for a long time been a well-known and explicable fact in the<br />

case of small molecules [135]; the case of hydrocarbon chains was simply and<br />

clearly presented by Lavalley and Sheppard [ I361 for the molecule of hexadeuteropropane<br />

(CD3-CH?-CD3), which consists of an isolated CH? group between two<br />

CD3 groups with which little or no mechanical coupling is likely to occur, in a first<br />

approximation. The molecule belongs to the symmetry. Figure 3-38a gives the<br />

survey Raman spectrum of this molecule, while Figure 3-38b and 3-38c show an<br />

expanded section of the Raman spectrum polarized ( 1 1 ) and depolarized (I)[137].<br />

Normally, in the CH? stretching range we should reasonably expect two lines to<br />

be associated with the antisymmetric (d-) and symmetric (d') CH? stretching of<br />

species BZ (1) and A1 ( 1 1 ) respectively. Instead, we find two strong lines certainly<br />

with A, symmetry (polarized) and one certainly with B7 symmetry (depolarized).<br />

The dynamical situation which justifies the observed Raman spectrum is the fol-


CO]CH$O]<br />

a<br />

CD3CHzC03 MELT, I I I<br />

300.K<br />

L<br />

I(<br />

J<br />

177.K<br />

L<br />

101.K<br />

1400 1500 2800 2900 3000<br />

AAYAN FREPUENCY (ern-'<br />

b<br />

0 1500 2800 2900 xxx)<br />

RAMAN FREPUENCY (crn-'<br />

C<br />

Figure 3-38. (a) Survey Raman spectrum of hexadeuteropropane (CDj-CHz-CDl) in the solid<br />

state at 15 "K. The species of the vibrational transitions are indicated; (b) expanded temperaturedependent<br />

Raman spectrum in the 1400-3000 ern.-' range in polarized light (11 polarization) exhibiting<br />

the transitions of A, species; (c) expanded temperature-dependei~t Raman spectrum in the<br />

1400-3000 cn1-l range in polarized light (1) showing the transitions of B1 species, free from Fei-mi<br />

resonances near 2925 cin-'. The dashed scattering is due to 'bleeding' of the 11 polarization due to<br />

experimental difficulties.


lowing. The CHl-bending mode 6 [certainly of A1 symmetry ( 1 1 ) )<br />

3.17 Fermi Resonaiices 167<br />

(Figure 3-37b)<br />

observed at 1460 cm-~' should have its first overtone 126 : 2 x 1460 = 2920 cni-',<br />

A1 x A1 = Al) which can enter Fermi resonance with the d+ mode of A1 symmetry.<br />

Eigenstates mix, frequencies are split apart with intensity borrowing froin the<br />

fundamental to the overtone producing two strong lines at 2876 and 2942 cni- '<br />

(Figure 3-38a-c'i.<br />

It should be remembered that the first overtone 26 should generally have vanishing<br />

intensity, but because of Fernii resonance its frequency is pushed upward and<br />

it borrows a lot of intensity from the fundamental d+ observed at 2876 cm-' .<br />

A similar, but seemingly more complex case, occurs for polymethylene chains.<br />

Let us consider the dispersion curve of polyethylene or the corresponding vibrations<br />

of finite polymethylene chains (Figure 3-39 j whose frequencies lie on these dispersion<br />

curves at finite values of the phase coupling (Section 3.5). The vibrations of the<br />

lower portion of the dispersion branch of CH2 rocking have their first overtone<br />

levels ( - 720 x 2 zz 1440 cnrl) close to the fundamental levels of the CHZ-bending<br />

motions. On their turn, the first overtones of the levels along the dispersion curve of<br />

the CH2-bendings occur just in the frequency region where the fhdaniental CH<br />

stretching modes occur (2 x -1440 z 2880 cnirl). When symmetry allows, Fernii<br />

resonances occur and levels repel each other with drastic intensity changes. The<br />

issue becomes formally more complex since (Section 3.6) when molecules crystallize<br />

in an orthorhonibic lattice (with two molecule per unit cell) the number of frequency<br />

branches doubles, thus doubling the number of levels which can generate<br />

both overtones and combinations which may produce Fermi resonances in the<br />

proper frequency ranges, as just indicated.<br />

These Fernii resonance phenomena have found detailed mathematical treatment<br />

first tackled by Snyder et al. [138, 1391 for single all-tram chains and later reformulated<br />

by Abbate et al. for ii) single chains, (ii) an orthorhombic lattice [140], (iii)<br />

all-trans, and (iv) conformationally distorted chains [ 141, 1421. The most interesting<br />

features to be accounted for are the Raman scattering of polymethylene chains in<br />

the frequency regions centered near 1450 and 2900 cnirl.<br />

For the sake of simplicity in the discussion which follows we label with (+) and<br />

(-) the symmetric and antisymnietric combinations respectively either of internal<br />

coordinates (e.g., C-H stretches within a CH? group) or of 'group coordinates' between<br />

two CH2 units within the 'chemical repeat unit' made up by two CH? groups.<br />

First, we discuss the case of true k = 0 modes for an infinite chain (or the case of<br />

k 4 0 for finite long chains). Let us, for the sake of simplicity, consider the spectrum<br />

of traiis polyethylene both as a single chain and crystallized in an orthorhombic<br />

lattice with two molecules per unit cell. Since the spectrum of polyethylene<br />

as single chain is not available for this discussion we necessarily use the calculated<br />

spectra as simulation of reality [140]. Let us first compare the general features of the<br />

calculated Raman spectra in the CHz-bending centered near 1435 cm-' for a single<br />

chain and for crystalline polyethylene (Figures 3-40a, b) (point group D?!,). The<br />

single very strong line near 1435 cm-' (Figure 3-39) is to be assigned to the limiting<br />

k = 0 totally symmetric in-phase CH?S+ mode of A, species. For the crystalline<br />

material, such a mode splits into two levels near 1420 (Ag) and 1435 an-' (BI?); an<br />

additional broad scattering (with a wing towards higher frequencies) is calculated


168<br />

3 Vibrational <strong>Spect</strong>ra as u Probe oj. Strurtrrrcil Ordu<br />

30(<br />

V<br />

cm -1<br />

I<br />

25C<br />

15C<br />

1oc<br />

Figure 3-39. Dispersion curves of<br />

single-chain polyethylene, indicating the<br />

scheme of Fermi resonances which can<br />

occur between phonon of different<br />

branches.<br />

near 1464 cm-'. Following the previous procedure we look first at the first overtone<br />

of the infrared active CH? rocking P- mode for a single chain (line group symmetry)<br />

(2P-), BzU x Bzu =A,. For the line-group, Fermi resonance can take place<br />

between 2P- of A, species and the A&+ flindamental, giving rise to the weaker line<br />

near 1445 an-'. Extension of the treatment to the case of the orthorhombic lattice<br />

with two molecules per unit cell needs to consider that each level splits into two


3.17 Ferini Re.sonrriire.s 169<br />

a<br />

Figure 3-40. (a) Calculated Raman spectra in the CHZ-bending region of single-chain polyethylene.<br />

The BZ was divided into 72 intervals of 5". Each Lorentzian function has AI~,/~ = 5 cn-': (b)<br />

Calculated Ranian spectrum in the CHz-bending region of crystalline polyethylene as n convolution<br />

of Lorentzian lines as in case (a).


15<br />

Figure 3-41. Calculated Raman spectrum for crystalline polyethylene in the CH-stretching region.<br />

The spectrum is a convolution of numerous combination and overtone levels in Ferini resonance<br />

with the nonresonant CH2 antisyinmetric stretchings. This spectrum should be compared with the<br />

experimental one of Figure 3-37.<br />

levels; moreover, in addition to the overtone levels, we should also consider their<br />

combinations. For the space group we can use the observed experimental frequencies:<br />

(2P-)-, B3u x B3,, = A,, 2 x 731 = 1462 an-'; 2(P-)+, B?,, x Bzu = A,,<br />

2 x 719 = 1438 cm-'; (P-)- + (P-)', B3u x B?, = Bl,, 731 + 719 = 1450 cn-'. It<br />

follows that for the space group the mode (af)+ can couple with 2(P-)- and<br />

2(P-)+ while (6')- can interact with the combination (P-)- + (P-)'. Raman<br />

spectra of stretch-oriented (extruded rods) in various polarization directions 128,<br />

1231 show that the symmetry species of the lines at 1464 and 1440 cm-' is certainly<br />

Big, while the line at 1416 cm-' is certainly A,.<br />

The full treatment needs to consider in addition to the k 40 modes also the<br />

k # 0 modes (throughout the whole BZ) due to either the finiteness of the chain, or<br />

to the whole continuous dispersion curve for an infinite chain. The population of<br />

overtones and combinations becomes extremely large and many of these levels can<br />

enter Fermi resonances with a large transfer of intensity from the fundamental<br />

strong modes.<br />

Once the bending region has been understood the same approach must be taken<br />

by considering all line-group or space-group overtones and combinations of the d<br />

modes which are in Fermi resonance with the CH2-stretching modes. All details of<br />

the calculations are provided in [140]. Figure 3-41 gives the result for crystalline<br />

polyethylene.<br />

The most important and useful feature in the Raman spectra of polymethylene<br />

chains is that reported in Figure 3-37, which is easily observed every time a sequence<br />

of CH2 units occurs in a molecule. If the system forms an orthorhombic<br />

lattice, generally one observes a broad and strong line near 2850 cn-' with a broad<br />

wing which extends towards higher frequencies foiming a broad background scattering<br />

that reaches -2950 cm-'. Broad and sometimes ill-defined bands appear<br />

near 2930 and 2900 cin-l. For the orthorhoinbic system, such a background scat-


3. I7 Fernii Resoiiances 17 1<br />

tering has been shown experimentally to have A, symmetry [28, 1231 and originates<br />

from a convolution of a large population of overtone and combination levels which<br />

are in Fermi resonance with the symmetric CH?-stretching mode which, in principle,<br />

should occur as an unperturbed strong level near 2880 cm-', but has been<br />

pushed downward by Fernii resonances with the manyfold of overtones and combinations<br />

to which it has also lent intensity (Figure 3-41). Floating on top of such<br />

a broad A, scattering we find the strong and sharp line associated with the CH?<br />

antisymmetric stretching which remains unperturbed because it does not enter in<br />

any Fermi resonance coupling.<br />

3.17.2 Applications<br />

We need now to translate the above theoi-etical results into practice. First, let us<br />

start from an orthorhombic lattice with two molecules per unit cell and lower the<br />

symmetry to a supermolecular organization in which the cell can be considered as<br />

isolated and decoupled from its neighbors. This situation can be achieved in the<br />

following case: (i) isotopic solid solutions; (ii) molecule as a guest in a chemically<br />

different host lattice (e.g., urea clathrates) [ 1431: (iii) liquid crystalline-like systems;<br />

and (iv) systems with confomiationally distorted chain ends, but with an ordered<br />

bulk [1031. The following changes of the Raman spectrum are expected (and<br />

observed):<br />

(i) The triplet of lines in the bending region becomes almost a singlet (@), since the<br />

crystal field splitting disappears and the convolution of overtone and conibination<br />

bands for finite chains (two phonon states for polyethylene) in Fermi<br />

resonance with the crystal field split fundamentals are removed (Figure 3-40).<br />

(ii) As the number of overtone and conibination states in Fermi resonance with the<br />

A, fundamental state (CH2 symmetric stretch) decreases we expect the lowering<br />

of the total A, scattering and possibly a shift toward higher frequencies<br />

of the A, fundamental. Generally, the scattering in the valley between the<br />

antisymmetric and symmetric CH? stretch 2980 and 2850 cm-' respectively)<br />

is observed to decrease and the two fundamental lines sharpen somewhat.<br />

Attention should be paid to the fact that for chains with CH3 end groups<br />

the symmetric CH-stretching mode shows up as a weak line within the same<br />

frequency range.<br />

The next step is to interrupt the long trans chain by gauche defects. The following<br />

changes are expected (and observed) in the Rainan spectrum [ 141, 1421.<br />

(iii) As both P and 6 inodes of CH? groups in gauche conformation occur at different<br />

frequencies the population of vibrational levels which enter Fernii resonances<br />

in the trans interaction scheme decreases. We expect changes of the<br />

Raman spectrum both in the CH?-stretching and bending regions. Qualitatively,<br />

the vibrational spectrum should change from that of a full. all-fr~nis<br />

single chain molecule toward that of a fully decoupled structure, thus tending


172 3 Vibrutionul <strong>Spect</strong>ra as n Pvohe of StrirctLiral Order<br />

to the situation described in Figure 3-38 for hexadeuteriopropane. It should<br />

be noticed that necessarily the 2850 cn1-l line should move towards higher<br />

frequencies approaching the unperturbed value near 2880 cni-' and the 2s line<br />

near 2930 cm-' should eventually become stronger as it does not share with the<br />

manyfold of binary combinations and overtones the intensity borrowed from<br />

the fundamental originally observed near 2850 cm-'.<br />

(iv) A drastic increase in concentration of gaziche structures occurs when the polymethylene<br />

system melts. In that case one can assume that the statistics of<br />

Flory's Rotational Isomeric State Model [48] works and we can describe the<br />

system as coiisisting of a distribution of gauche and short trans sequences.<br />

Upon melting, the d+ band near 2850 cm-' becomes the strongest feature in<br />

the spectrum and d- collapses into a very broad band with its maximum<br />

shifted upward of 4-10 cni-I. The overtone level near 2940 cm-l increases<br />

in intensity. This is a very useful marker when the chain is confoimationally<br />

collapsed (liquid like) (Figure 3-37b).<br />

(v) There are two possible interpretations of the broadening of d- band observed in<br />

the liquid phase. The first is that the broadening is the result of a convolution<br />

of many transitions associated to d- mode of CH? groups in constrained nonequilibrium<br />

near gauche conformations [143, 1441. Another explanation of the<br />

broadening refers to the overall dynamics of the whole chain which may occur<br />

in the liquid (see Section 3.18) [145].<br />

(vi) A detailed study has been carried out by Abbate et al. on the molecule<br />

CD3-CH2-CHa-CD3 11411 as a model in which (a) the possible coupling with<br />

the methyl end groups are removed by deuteration and (b) only one degree of<br />

torsional freedom between CH? groups is allowed. It is immediately apparent<br />

that in gauche conformation the symmetry of the molecule is lowered, thus<br />

changing the resonance schemes which may involve also the d- mode which<br />

is unperturbed in the truns case. Calculations show that both d- and d+<br />

are allowed to mix, thus reshuffling the intensities in a way that is difficult to<br />

predict.<br />

(vii) The experimental intensity ratio R = I2~50/12940 has been proposed as an experimental<br />

way to determine the relative concentration of trans versus gauche<br />

conformers in a sample. R has been matter of strong debate (see for instance<br />

[ 1401). As a result of the calculations by Abbate on hexadeuteriobutane [141],<br />

because of a mixing of the modes and the unpredictable changes of the extent<br />

of Feimi resonance, we do not advise the use of R as analytical probe in liquid<br />

polymethylene systems or in partially disordered solid polyethylene samples.<br />

3.18 Band Broadening and Conformational Flexibility<br />

The experimental spectroscopic facts which need to be accounted for in a comprehensive<br />

way in relation to order-disorder are the following (we report here the<br />

most striking manifestations, recorded in various ways, for oligo and polyniethylene<br />

systems):


3.18 Band Broadening and Confonnational Flexibility 173<br />

Figure 3-42. Hexadeuteropropane:<br />

10-<br />

n<br />

0 " d+(CHZ)<br />

I I I<br />

Figure 3-43. Ranian spectrum in the C-Hstretching<br />

range of C~lH44: (a) solid and ib) melt.<br />

a<br />

3050 3000 2950 2900 2850 2800 cm-I<br />

1. For short chains, the antysimmetric d- mode broadens in the Raman spectrum<br />

by increasing temperature and sometimes shifts upward in frequency (Figure<br />

3-42), while d+ neither shifts substantially nor broadens. (See also Figure 3-37.)<br />

2. For long polymethylene chains and polyethylene at melting, the strong d-<br />

modes collapses into a featureless very broad scattering and by increasing temperature<br />

shifts its maximum towards higher frequencies (Figure 3-43).<br />

3. For crystalline n-alkanes the width of d- does not increase with teniperature up<br />

to a few degrees below melting (Figure 3-44).<br />

4. When n-alkanes are inserted as clathrates in urea channels, the width of d- is


174 3 Vilmitionul Slirctru us ci Probe of Stnictiirul Ortlev<br />

Figure 3-44. Evolution with temperature of the Raman<br />

spectrum of solid Cl,Hd, in the CH stretching rangc.<br />

Notice the behavior with temperature of d+ and d- inodes.


3.18 Bmd Brourleiziriy md Cmforrnational Flrxibilitjj 175<br />

IL-180°C<br />

-1<br />

cm<br />

3050 3000 2950 2900 2850 2800 2750<br />

Figure 3-45 Raman spectrum of C~jH52 in urea clathrate at 3°C and -180°C. Notice that for<br />

d-At,,,,2 = 6 cn-' at T= -180°C and Ai~,p = 28 cm-' at T= 25°C; Ai~,p of d' remains practically<br />

constant at the same two temperatures.<br />

narrow at low temperatures and broadens substantially at higher temperatures<br />

(Figure 3-45).<br />

3.18.1 Principles<br />

In Section 3.17, experiments have been presented showing that in the case of polyinethylene<br />

molecules (including polyethylene), in going from the solid all-tlvrizs to<br />

the liquid phases, the pattern of Raman scattering changes dramatically but in<br />

a consistent and characteristic way. We focus our attention on the temperaturedependent<br />

broadening observed for the d- and 8 modes; in contrast, the width and<br />

shape of d+ modes are practically temperature-independent. In Section 3.17.2 it was<br />

suggested that one possible explanation of the band broadening of d- could be a<br />

convolution of bands associated with d- modes in different quasi g~iziclie conformations<br />

existing in the liquid or 'amorphous' material.<br />

This interpretation may be only partial since Snyder has shown that band<br />

broadening of d- and 6 modes is certainly observed also for hexadecane in uiea<br />

clathrates where the molecule is believed to be (on average) in the tu.m~-planar


176 3 Lri/ibvational <strong>Spect</strong>rci cis (1 Probe c?f'Stncc.tLiral Ovtlev<br />

conformation [ 1461. Some additional or alternative processes then must exist. We<br />

wish to discuss the results of band-broadening studies which may shine more light<br />

on the conformatioiial dynamics in systems where order-disorder may exist.<br />

A full discussion of the theory and experiments on band-broadening effects in<br />

n-alkanes is reported in [137]. We wish to outline here a few basic concepts which,<br />

in a first approximation, are the tool to be used in the understanding of the complex<br />

dynamics in conforinationally flexible systems.<br />

Band broadening [147] both in infrared and Raman spectra of polyatomic molecules<br />

in the liquid phase may arise from vibrational relaxations (VIBR) (i.e., some<br />

kind of process dissipates the vibrational energy stored in a vibrational transition)<br />

and/or from reorientational motions of molecules (REORIENT) (i.e., the energy<br />

stored in a vibrational transition is used to generate a reorientation in space of<br />

the molecule). Vibrational relaxation can be due to vibrational energy dissipation<br />

to the rotational-translational degrees of freedom of the thermal bath, to intramolecular<br />

energy transfer between vibrational modes, or to several intermolecular<br />

phenomena such as resonance vibrational energy transfer and vibrational<br />

dephasing. Theory states that the amplitude-dependent vibrational relaxations are<br />

temperature-independent, while the angular-dependent reorientational relaxations<br />

depend strongly on temperature.<br />

Raman scattering provides a nice way to gather information on these processes.<br />

For liquid samples, it is possible to separate experimentally the isotropic spectrum<br />

(ISOT) from the anisotropic one (ANIS), namely<br />

I(IS0T) = I(ll) - (4/3)I(i)<br />

(3-55 j<br />

I(AN1S) = I(i) (3-56)<br />

From these measurable quantities, under certain approximations [ 1371, it is also<br />

possible to give precise expressions for the relaxation times:<br />

where<br />

z(V1BR) = 1 /(KcAu~~)(ISOT) (3-57)<br />

z(RE0RIENT) = ~/(KcAv~/~)(REORIENT) (3-58)<br />

It is coinnionly known that the depolarization ratio for linearly polarized light is<br />

given by<br />

p = I(I)/I(II) 1 3P2/(45C(' + 4p2j (3-60)<br />

where a is the mean polarizability describing the isotropic component of the Ranian<br />

polarizability tensor and P, the anisotropy, describes the anisotropic component.<br />

It follows that for totally symmetric modes p 5, 0, I(il)= IfISOT), and the vibra-


3.18 Band Bvoaderiing arid Conjonnationril Flexibility 177<br />

tional relaxation contribution can be easily evaluated. For all antisymmetric modes<br />

a = 0 and p = $; only the contribution from coupled vibrational-reorientational<br />

motions can then be obtained from Raman experiments. The evaluation of the pure<br />

reorientational contribution can be obtained by suitably subtracting the contribution<br />

from vibrational relaxations (Eq. (3-59)).<br />

3.18.2 Short and long n-alkanes<br />

The Raman spectra of a series of n-alkane suitably chosen have been studied [137]<br />

with the precise aim of extracting from d- (antisymmetric, - depolarized) and d'<br />

(symmetric - polarized) modes of the CH2 group information of general validity on<br />

intra- and intermolecular dynamics using the techniques discussed in Section 3. IS. 1.<br />

The following molecules have been studied [ 1371 in the solid and liquid states:<br />

propane-d6 (CD3CH2CD3) butane-d6 (CD3CH2CH2CD3), and pentane-dlo<br />

(CD~CH~CHZCH~CD~) and n-nonadecane. The following conclusions, which may<br />

be of general interest, have been derived:<br />

1. For d+ p = 0, thus allowing us to distinguish the contribution by vibrational<br />

relaxation processes; for d- CI = 0 and p = %, thus the coupled vibrationalreorientational<br />

processes are active simultaneously. The reorientational contribution<br />

was derived by suitable deconvolution methods (Eq. (3-59)).<br />

2. The temperature independence of the band width of d+ modes has been verified<br />

and provides an average value of T(VIBR) - 1.3 x 10l2 s for the three small<br />

deuterated molecules. An amplitude-dependent vibrational dephasiiig process<br />

(inhomogeneous broadening) best accounts for the observed temperatureindependent<br />

broadening.<br />

3. A ~ J of ~ / d- ~ is strongly teniperature-dependent and increases rapidly with<br />

increasing temperature (Figure 3-41); correspondingly, r(RE0RIENT) decreases<br />

by at least one order of magnitude from low to higher temperatures (see<br />

Table 3-3). It is found that, at a given temperature, identical for the three deuterated<br />

molecules, the band width increases with increasing chain length. Moreover,<br />

at a constant temperature, the reorientational relaxation time increases<br />

with increasing chain length. This trend is consistent with the view that the contribution<br />

of reorientational relaxation decreases with increasing chain length, as<br />

might be expected based on inertial and frictional effect. Indeed, it is conceivable<br />

that while the molecule of propane in the liquid phase may easily tumble in<br />

all directions about its axes, for longer molecules the end-over-end tumbling<br />

becomes progressively less likely while some sort of rotation about the long<br />

molecular axis may still occur even for longer n-alkanes.<br />

4. For chains larger than propane (for which no torsional motion about<br />

(CHl)-(CH?) bonds exists) it is quite conceivable that the increasing hinderance<br />

to molecular reorientation which broadens d- mode comes from torsional<br />

motions of the molecular skeleton, i.e., its overall torsional flexibility. In other<br />

words, when a flexible molecule tries to tumble it necessarily drives the motions<br />

of the torsional angles. We expect these contributions to increase with iiicrcasing


178 3 Vihmtional <strong>Spect</strong>ra ITS CI Probe of Structural Order<br />

Table 3-3. Frequencies, bandwidths (full widths at half inaxiriiumi and reorieiitntioiiizl-rel~ixalion<br />

times of d- of propane-dh, butane-d6 and pentane-dl,] as a function of temperature.<br />

Temperature A1'1,2(d-) s(RE0RIENT) d-<br />

iK) (cm-') (cm-l) x 1012 s<br />

CD3CH:CD? 101<br />

177<br />

235<br />

288<br />

300<br />

CD1 CHzCH2 CD3 135<br />

185<br />

231<br />

300<br />

CD3CD2CH2CD2CD3 143<br />

191<br />

240<br />

300<br />

2919.0 15.7<br />

3920.5 24.8<br />

2925.0 70.0<br />

2923.5 79.0<br />

2896.5 15.0<br />

2896.5 24.2<br />

2899.5 40.0<br />

2898.0 65.0<br />

2896.5 18.1<br />

2898.0 20.1<br />

2899.5 24,s<br />

2901.0 30.9<br />

1.3<br />

0.58<br />

0. I7<br />

0.15<br />

1.7<br />

0.66<br />

0 33<br />

0.1s<br />

1.2<br />

0.93<br />

0.67<br />

0.48<br />

temperature and chain length. The threshold temperature at which for a given<br />

chain length coupled rotations and torsion are able to generate conformational<br />

kinks in a more or less nonequilibrium structure (either pinned at a given site or<br />

mobile along the chain) remains an unsolved problem faced by many authors<br />

who tackled the problem by numerical simulations [ 1481.<br />

5. For the short-chain models it has been found that, by increasing the temperature<br />

(i.e., when the coupling between rotations and torsions increases), the frequency<br />

of d- increases consistently by a few wavenumbers (see Table 3-3); this suggests<br />

that either the diagonal or off-diagonal force constants within each CH2 unit<br />

change when the niolecule is (even slightly) conformationally distorted or the<br />

onset of torsional distortions activates iiitraniolecular coupliiig which extends at<br />

an unknown distance s at either side of the CH2 group (see Eq. (3-45)).<br />

3.18.3 Collective Chain Flexibility<br />

In this section our attention is focused on attempts to describe in molecular terms<br />

the broadening and the upward shift of d- mode as discussed in Sections 3.18.1 and<br />

3.18.2.<br />

Let us first consider the case of a long 11-alkane molecule in the solid crystalline<br />

phase. Froin lattice dynainical calculations it is known that collective librational<br />

phonons (about the chain axis) exist in the lattice, some giving rise to scattering<br />

and/or absorption in infrared [70]. Let us take a n-alkane molecule with an even<br />

number of carbon atom which belongs to the C?h point group. The rigid rotation<br />

about its axis is of B, species, the same as the Raman active d- mode. When the


3.18 Band Broadening and Confornmtioiud Flesibility 179<br />

molecule is embedded in a medium which transforms the free rotation into a librational<br />

motion d- and the libration can, in principle, couple because they belong<br />

to the same symmetry species. A similar observation was made by Snyder for the<br />

CH2 rocking mode 11461.<br />

The contribution toward the understanding of chain flexibility comes from the<br />

following experiments and calculations which are described in detail in 11371 and<br />

[145].<br />

1. The first experiment has been that of recording the Rainan spectra of long<br />

n-alkanes in urea clathrates [ 1451 and in another host lattice (perhydrotriphenylene)<br />

where each guest n-alkane molecule sits in a channel, isolated from<br />

the other chains and retaining, on average, the all tram-planar structure. Similar<br />

experiments are also reported in [146] and [149]. By changing the temperature,<br />

the host lattice changes the average diameter of the columns where the guest<br />

molecule sits, thus allowing a varying degree of librational freedom for the guest<br />

molecule. While at low temperature the librational motion is strongly hindered,<br />

at relatively higher temperatures the lattice expands anisotropically allowing the<br />

guest molecule to perform librational motions of large amplitude.<br />

We expect the band of the d- mode to be narrow at low temperatures and to<br />

broaden at higher temperatures. Figure 3-45 shows that this is indeed the case.<br />

The band not only broadens but its frequency also shifts upward by a few<br />

wavenumbers. It is concluded that the almost rigid (and possibly small amplitude)<br />

collective librations about the chain axis at low temperature evolve with<br />

increasing temperature towards a collective libro-torsional (or libro-twisting)<br />

motion with nonnegligible angular distortions more or less hindered by the<br />

environment.<br />

2. It must be pointed out that the broadening and upward frequency shift of the d-<br />

mode cannot be ascribed to the generation of gauche defects inside the host lattice.<br />

It has been verified experimentally that the frequencies of LAMl, LAM3,<br />

and LAM5 of the guest molecule embedded in the host lattice are identical to<br />

those observed for the same n-alkane in the crystalline state [145]. Hence, the<br />

molecule does not change its length from that of the tvaizs structure [145].<br />

3. In going from the solid to ‘softer’ systems, certainly the frequency of the d-<br />

mode shifts upward by a few wavenumbers. No shifting (nor broadening) is<br />

observed for crystalline n-alkanes until the so-called soft ‘rotatory phase’ is<br />

reached before melting. In the so-called a phase (see Section 3.19) some shifting<br />

and broadening is observed.<br />

Levin [150] and Gaber and Peticolas [151] have shown that the upper shift<br />

and broadening occurs also in bioinembrane materials. Such a shift has been<br />

used by Gaber and Peticolas to monitor the ‘lateral interactions’ in phospholipids.<br />

We notice that for these phospholipid-type material the long-chain alkyl<br />

residues are fixed at one end to polar head groups and thus cannot perform rigid<br />

librational motions about their axes. Forcefully they can only perform librotwisting<br />

or libro-torsional motions.<br />

4. The last step is to account for the dynamical origin of the fact that when the long<br />

polyniethylene chain increases its libro-torsional flexibility the frequency of d-


180 3 Vibrational <strong>Spect</strong>ra as ci Probe of Strtrctirrcil Ordu<br />

modes shifts very selectively upward, while the frequency of d+ mode remains<br />

unchanged. There must be some specific process which should be discovered.<br />

With this purpose various dynamical possibilities have been considered [ 1451.<br />

It has been concluded that a rational explanation is that there exists a selective<br />

intramolecular coupling between d- modes and the torsional modes of the carbon<br />

backbone. Symmetry consideration tell us that for a trans-planar molecule<br />

the existence of such coupling will affect d- and not d+ modes. In other words,<br />

within the quadratic approximations interactions of the type fCH stretch/C-C torhions<br />

seem to exist and are modulated by the amplitude of the collective torsional<br />

flexibility.<br />

Another contribution to the upper shift may come from a complex scheme<br />

of Fermi resonances which are activated because of the lowering of the instantaneous<br />

symmetry during the libro-torsional motion. This difficult problem<br />

remains unsolved.<br />

5. The final observation, relevant for the discussion which follows in Section 3.19,<br />

is that in crystalline a-alkanes the band width of d- is relatively narrow and the<br />

frequency does not change throughout the temperature range of existence of the<br />

crystalline lattice. This suggests that libro-torsional motions are not allowed in<br />

the crystal and only small amplitude collective librational phonons seem to be<br />

allowed. Some additional observations will be added in Section 3.19.<br />

3.19 A Worked Example: From n-Alkanes to<br />

Polyethylene Structure and Dynamics<br />

3.19.1 Description of the Physical Phenomena<br />

We wish to guide the reader through a realistic case worked out step-by-step using<br />

most of the concepts and methods discussed in this chapter. We consider the case of<br />

n-nonadecane (CH~-(CH~)I~-CH~) as a prototypical case and try to derive from<br />

the vibrational spectra all possible information about its structure and dynamics.<br />

Both order and disorder will be considered. The methods applied and the conclusions<br />

reached form the background knowledge to be used for the understanding ot<br />

(i) the structural properties of many molecules containing long-chain alkyl derivatives;<br />

and (ii) the structural features of polyethylene or other polymers containing<br />

polymethylene segments.<br />

The basic physics we wish to understand is the mechanism at the molecular level<br />

which generates the phase transitions of relatively short n-alkanes. The experimental<br />

nonspectroscopic facts to be accounted for are the following (for a collection of<br />

references and a full discussion of many of the data presented in the first part of this<br />

section see [ 1031): orthorhombic n-alkanes first undergo a solid-solid phase transition<br />

(Toicc) to a phase variously described as 'rotatory phase' or 'c1' phase whose<br />

structure has been matter of controversy in the literature. This phase characterizes a


3.19 A Worked E.utinple 181<br />

pre-melting state for n-alkanes; just a few degrees above ToLx the crystal melts. The<br />

temperature range of existence for the t( phase decreases with increasing the length<br />

of the chain and is also pressure-dependent. Even polyethylene seems to show a<br />

similar pre-melting state just a few degrees before melting.<br />

The generality of the phenomenon to be understood was the reason which has<br />

convinced us of the need to perfom a detailed spectroscopic study on the simplest<br />

system, n-nonadecane which shows the largest temperature range for the existence<br />

of the a phase.<br />

We have studied the following molecules: n-nonadecane, n-nonadecane-2D2 and<br />

n-nonadecane- 1 OD? (hereafter referred to as C19, 2D2 and 1 OD2 respectively).<br />

The known structural features of C19 are the following: crystalline C19 below<br />

22.8 "C crystallizes in the orthorhonibic lattice with two molecules per unit cell<br />

[152]. Chains are trans-planar and perpendicular to the plane formed by the methyl<br />

groups, thus being perpendicular to the lamella surface. At 22.8 "C a phase transition<br />

is observed to the a phase and melting occurs at 32 "C. The measured heat of<br />

transition is 3300 cal/mol, while the heat of melting is 10950 cal/mol [153]. The<br />

volume changes determined with a dilatometer at the solid-solid phase transition<br />

and melting are 2.5% and 10.5% respectively. The structure of the r phase below<br />

melting is reported to be 'hexagonal' with some indication of rigid rotations of the<br />

chains [154].<br />

We add here inforination for the two selectively deuterated materials which were<br />

prepared in the authors' laboratory. There is no reason to think that 2D10 may<br />

not have the same structure as the parent molecule. The perturbations by the two<br />

deuterium atoms should be negligible regarding the static structure while they will<br />

affect the vibrational spectra. As a result of the 'regular' packing, the lamellae<br />

formed by these all-trans 10D2 molecules contain precisely in the middle a 'layer' of<br />

CD2 groups. From the viewpoint of vibrations we expect normal crystal field splittings<br />

for the vibrations of the CD2 since intermolecular phase coupling within the<br />

lattice may occur regularly. In contrast, 2D2 molecules in the crystal, while retaining<br />

the orthorhombic structure, are arranged at random with the heads labeled<br />

with CD2 randomly distributed up or down within the crystal. Within Borii-<br />

Oppenheimer approximation the intermolecular potential cannot distinguish between<br />

deuterated or undeuterated ends of the hydrocarbon chain. However, because<br />

of the random mechanical perturbations of the vibrations at the site of chain heads<br />

in this molecule, crystal field splitting of the CD2 vibrations is not expected (in<br />

spite of the orthorhombic packing). These predictions are indeed verified in the infrared<br />

and Raman spectra of both lOD2 and 2D2.<br />

3.19.2 Vibrational <strong>Spect</strong>ra of pure n-Nonadecane<br />

N = 59 is the number of atoms in the C19 molecule. The symmetry point group of the<br />

single trcms-planar chain of C19 is C?" and its 59 x 3 - 6 = 171 vibrations can be<br />

classified in the following irreducible representations with their spectroscopic activity:


182 3 T7ibmtional <strong>Spect</strong>ra CIS a Probe of Structurd Order<br />

I- P<br />

u+w<br />

I I I I I 1<br />

IW 1400 1200 loo0 000 600<br />

v cm-'<br />

Figure 3-46. Survey infrared spectrum of n-nonadecane in the orthorhonibic phase. Band progressions<br />

are indicated. For the labeling of the modes see text. R = skeletal stretching, p = CHj<br />

deformation.<br />

The infrared spectrum of C19 in the solid orthorhonibic phase is shown in Figure<br />

3-46; we use this spectrum as reference in the present analysis. The Raman spectrum<br />

of the same material is given in Figure 3-47.<br />

The identification of the large number of normal inodes in the infrared and<br />

Raman spectra of these molecules becomes indeed unfeasible. Life becomes easier<br />

if we consider C19 as a segment of polyethylene made up by 17 CH2 units and try<br />

to located the 17 normal modes lying in each frequency branch of the dispersion<br />

curves of infinite polyethylene (Figure 3-1). Each mode is then characterized by a<br />

phase shift q, between adjacent CH2 units, qj = 2nj/17, (here j = 0,. . . , 17 - 1).<br />

The possibility of observing all noniial modes depends on their symmetry and on<br />

their intrinsic intensity in infrared or Raman, as well as on the dispersion of each<br />

phonon branch. As discussed in Section 3.9.2, we expect to observe in the infrared<br />

spectrum band progressions with decreasing intensity for branches with reasonable<br />

dispersion, while for flat branches band progression will be strongly compressed<br />

with strong overlapping.<br />

The experimental observation predicted by theory has been discussed previously<br />

in detail by Snyder and Schachtaschneider [9, 341. From the dispersion curves<br />

of Figure 3-1 it follows that 110 clear band progressions are expected for CH2<br />

stretching (d, overlapping) and CH2 bending (6, overlapping), while progressions<br />

are expected for CH2 wagging (w, very weak in IR); in the CH? wagging progression<br />

we clearly locate Ws, W7, WS, and W11; the other CH2 wagging motions are


3.19 A Wovh-ecl E.uanzple 183


184 3 T/ihrutionul <strong>Spect</strong>ra as u Probe of Strzicturul Order<br />

overlapped by the stronger CH3 U mode near 1375 cm-' . The expected progression<br />

of CH2 twisting (t) must be very weak in IR while CH? rocking (PI must be strong/<br />

medium and clearly noticeable from 720 to near - 1100 cni-'. C-C stretchings are<br />

weak in IR. For the band progression of the P modes we expect to observe nine<br />

components of B2 species corresponding to quasi-phonons with odd j values.<br />

J = I, 3, and 5 turn out to be almost overlapping and the others must he observed in<br />

the frequency range -750-1 100 cm-' . The observed band progressions are clearly<br />

displayed in Figure 3-46.<br />

The infrared bands at 1375 and 890 an-' are unquestionably assigned to the<br />

umbrella (U) and rocking ([I) modes of the methyl groups respectively (Figure 3-46).<br />

The Raman spectrum provides further information. It is generally known that in<br />

the Raman spectra only the limiting k i 0 modes for the infinite chain appear as<br />

strong or easily observable transitions. This is also the case for C19 where only the<br />

k 2 0 modes are observed.<br />

The most important information conies from the observation of the band progression<br />

of the LAM modes (Section 3.16) with k = 1,3, and 5 components<br />

observed at 124.5, 342, and 490 cni-'. Using equations 3.53 and 3.54, the frequency<br />

of LAM1 at 124.5 c1n-l proves that the chain comprises exactly 19 C atoms.<br />

The conclusions are as follows:<br />

Conclusion no. 1<br />

The infrared and Raman spectra provide unquestionable evidence that C 19 below<br />

22.8"C consists of trnns-planar molecule with 17 CH2 units capped by two CH3<br />

groups. Further observation of the infrared spectrum for the orthorhombic phase<br />

shows the splitting (factor group splitting) of the 6 modes and of many components<br />

of P modes into doublets with a measurable intensity ratio (Figure 3-48). Analogously,<br />

the Raman line associated to k - 0 of the 6 mode near 1460 cni-' is split<br />

into a doublet (Figure 3-47). The number of components of the splitting is related to<br />

the number of molecules (2) in the unit cell. If the orthorhombic lattice is accepted,<br />

the intensity ratio of the components of the splitting in the infrared provides the<br />

value of the setting angle of the hydrocarbon chains in the unit cell (8 e 42"). The<br />

shape of the Ranian scattering in the 3000 and 1450 cni-' range can be interpreted<br />

in terms of Fermi resonances (Section 3.17) uniquely between trans-planar chains.<br />

Conclusion no. 2<br />

IR and Raman spectra provide evidence that two trms-planar molecules are<br />

organized in the orthorliombic lattice, with setting angle 8 = 42".<br />

3.19.3 Temperature-Dependent <strong>Spect</strong>ra<br />

Next, we study the temperature-dependent infrared and Raman spectra of C19<br />

through phase transition (Ta) (T, = 22.8 "C) and melting (T,) (TI,, = 32 "C). We<br />

first focus on the spectroscopy of C19. Each experimental observation is labeled and<br />

will be included in the overall analysis. In the IR spectrum:


3.19 A Worked Example 185<br />

19 5°C<br />

215°C<br />

28°C<br />

Figure 3-48. Temperature dependence of<br />

the infrared spectrum of a-nonadecane<br />

in the CW2 rocking region showing that<br />

the crystal field doublets coalesce into<br />

singlets at the orthorhombic 4 c( phase<br />

transitioiis indicating the existence of 780 760 740 720 700<br />

‘single chains’.<br />

v cm-1<br />

(a) At T, the band progression of the P modes which shows factor group splitting<br />

at lower temperature loses all the splitting, but remains as prominent feature in<br />

the IR spectrum until T, (Figure 3-48).<br />

(b) The progression of the P modes disappears at T,, and is replaced by a broad<br />

band near 722 crn-l.


186 3 Vibrational <strong>Spect</strong>ra cis a Probe of Structural Order<br />

I I I I I I<br />

1000 900 800 700<br />

v cm-'<br />

Figure 3-49. Band progression<br />

due to CH2 rocking<br />

modes for n-nonadecane in<br />

the orthorhornbic phase (at<br />

0°C) and in the a phase.<br />

Notice at 25 "C the appearance<br />

of a few weak bands<br />

and the disappearing of the<br />

crystal field splitting.<br />

(c) At T, the mode at 1375 cni-' (U) shifts to 1378.5 cm-'.<br />

(d) At T, a weak, clear and broad absorption appears at 766 cni-* and remains a<br />

clear, weak and broad band in the melt (Figure 3-49).<br />

(e) At T, additional peaks are observed in the infrared at 766, 804, 845, 866, 877,<br />

940, and 956 cm-' (Figure 3-49). Some of these modes may be assigned to the<br />

even j values of the P band progression.<br />

(f) At T, in the CH2 wagging progression we clearly locate WS, W7, WS, and W11<br />

as in the case of the solid, but the frequencies of these components are rigidly<br />

shifted toward higher frequencies for a few wavenumbers (Figure 3-50). The<br />

other CH2 wagging motions are overlapped by the CH7 U inode which shifts to<br />

-1378 cm-l.<br />

(8) At T, the CH; ,B mode shifts from 891 to 889 cin-l.<br />

(h) At T, a weak, but unquestionable peak arises at 1342 c1n-l (Figure 3-50).<br />

Upon melting:<br />

(i) At T, the band progression of the CH2 wagging disappears (Figure 3-50).<br />

(j) At T,ll the peak at 1342 c1ii-l gains considerable intensity (Figure 3-50).<br />

(1) At T,, the two well-known defect modes absorptions near 1353 cm-' and 1370<br />

cni-' appear together with the broad peak centered near 1306 cm-' (Figure<br />

3-50).<br />

In the case of the Rainan spectrum the most meaningful observations are the following:<br />

(m) Above T, the features of the k = 0 modes of the infinite trans-planar molecule<br />

are still the dominant features (Figure 3-51b).<br />

(n) At T, the lattice band at 98 cm-' disappears.<br />

(0) At T, the LAMI frequency shifts from 124.5 to 123.2 cm-' (Figure 3-52).


3.19 A Workrd Example 187<br />

Figure 3-50. Ternperaturedependent<br />

infraied spectra of<br />

n-nonadecane showing changes<br />

through the solid-cc (T - 22OC)<br />

and r-liquid (T - 29 "C) phase<br />

transitions<br />

i I , I<br />

1400 1360 1320 I290 1240<br />

v cm-'<br />

(p) At T, the A, component of the factor group splitting of the CH2-bending<br />

mode near 1417 cm-' practically disappears, even if not completely (Figure<br />

3-53).<br />

(q) At T, the valley between the A, (2845 cm-') and Bl, (2882 cm-') modes of the<br />

C-H-stretching region becomes emptier and the band at 2845 cm- ' sharpens.<br />

(r) At T, a slight increase in the scattering near 1090 cm-l is observed.<br />

(s) At T, on the lower frequency side of the well-isolated peak near 890 cm-' (/3<br />

mode) two weak bands are observed at 866 and 844 cm-' (Figure 3-54).


188 3 Vibrational <strong>Spect</strong>ra as a Probe ojStrticttirci1 Order<br />

ORTH.<br />

MELT<br />

I , 1 I<br />

2wQ 2800 ” 1200 800 coo 0<br />

v cm-1<br />

I<br />

Figure 3-51. Changes of the Raman spectrum of n-nonadecane in the various phases: (a) ortliorhombic,<br />

(b) a, (c) liquid.<br />

The Raman spectrum in the melt shows the following features:<br />

(t) At Tm the LAM1 mode disappears completely and is replaced by the so-called<br />

‘pseudo LAM’, a broad and weak scattering near 320 cn1-l common to all<br />

liquid n-alkanes, melt polyethylene or any molecule containing long CH2<br />

chains in the ‘melt’ state (Figure 3-51c).<br />

(u) At T, the Raman lines at 866 and 844 cm-’ gain in intensity and, together<br />

with the line near 890 cm-l, form a strong characteristic triplet with components<br />

of similar intensity (Figure 3-51c).<br />

(v) At Tm the Fermi resonance line at 1461 cmP1 decreases while the Fermi resonance<br />

line at 2930 cm-‘ increases.


3.19 A Worked E.rcnniplr 189<br />

LAM I<br />

I<br />

Figure 3-52. Details of the changes of the<br />

Ranian spectra of ti-nonadecatie at 0 "C<br />

(orthorhombic) and 29.7 "C ( M phase). The<br />

I<br />

lowering of the frequency of LAM 1 shows 200 170 140 110<br />

that interlamellar forces have weakened.<br />

v crn-1<br />

I 1 I I I I I 1<br />

1470 1440 1410 1380 1350 1320 1290 1260<br />

cm-1<br />

Figure 3-53. Details of the Raman spectrum of n-nonadecane just above and just below the solid<br />

- il phase transition. The crystal field splitting near 1430 cnr' disappears and only the single chain<br />

bending deformation remains at 1440 cin-'. The Fernii resonance line near 1460 cm-' is also<br />

slightly perturbed.


j&890 860 830<br />

29,JT<br />

~.~<br />

Figure 3-54. Details of the Raman spectrum of n-nonadecane<br />

in the orthorhombic (0°C) and c( phases 179.7‘C). The line<br />

near 890 cn-’ indicates that all chain ends are in the TT confomiations.<br />

In the u phase a small poptilation of chaiii~ has<br />

tilted their heads in TG arid GT conlormations.<br />

(w) At TI, in the CH stretching range the line at 3882 cin-’ is submerged by the<br />

broad scattering near 3000 cn-’ from which two peaks emerge at 2850 (d’)<br />

and 2875 cm-‘ (d-). The typical Raman spectrum of the “melt” is observed. It<br />

should be compared with Figure 3-43.<br />

(x) At T, near 720 cmP1 a clear stepwise increase of the diffuse broad background<br />

scattering is observed.<br />

(y) At T, the line near 1090 cin-’ becomes strong at the expenses of the k = 0<br />

singularities of the infinite trans-planar chain at 1135 and 1074 cm-’ in the<br />

orthorhombic and a phases. This line is characteristic of the stretching of the<br />

C-C bonds in gauche conformation (Figure 3-51c).<br />

Conclusioii no. 3<br />

At the orthorhombic to a phase transition the disappearance of the factor group<br />

splitting (observations (a) and (n)) indicates that the orthorhombic lattice with two<br />

molecules per units cells disappears and is replaced either by a unit cell with one<br />

molecule per unit cell or by a lattice where the tridimensional order is practically<br />

lost.<br />

Chains, however, remain mostly straight tmns-planar (observations if), (in),(o),<br />

(p) and (9)). Perturbations appear at chain ends (observations (c), (g) and (o)),<br />

indicating that the medium surrounding the CH3 group has changed, in particular<br />

the inter-lamellar forces have weakened (observation (0)). The following structures<br />

appear at the end of the molecule -CH?(T)-CH2(G)-CH?-CH3 and<br />

-CH?(G)-CH~IT)-CHI-CH~(observation (s)). This accounts for the lowering of<br />

the overall symmetry which causes the appearance of a few of the coinponents of<br />

the P modes inactive for the orthorhombic stiucture (observation (e)). It seems<br />

likely that some sort of disordering at the surface of the lamella takes place with<br />

some chains having their ends tilted in gauche conformations (observation (r)). The<br />

occurrence of a few GTG, GTG’ and GG conformational defects (or unspecified<br />

gauche defects) is indicated (observations (11) and (1)); the topology of these defects<br />

along the inoIecuIes cannot, so far, be defined.


3.19 A Worked E.xuniple 191<br />

Figure 3-55. Sketch of assumed<br />

locations of selectively deuterated<br />

chains of n-nonadecane (2D2 arid<br />

10D2, see text) in the orthorhombic<br />

crystal.<br />

-<br />

-5Hz- $ D z - ~ H ~<br />

Conclusion no. 4<br />

The process of melting can be described in the following way: the tvuizs-planar<br />

structure of the fraiu chains collapses (observations (i), (j) and (t)) and conformational<br />

disorder takes place accompanied by the appearance of the following<br />

conformational defects: GG, GTG, GTG’ (observations (j) and (l)), end-gauche<br />

(observation (u)) and gauche in general (observations (b), (t), (y), (v), (w), and (x)).<br />

In practice we find characteristic signals of liquid-like structures.<br />

With the aid of the selectively deuterated species 2D2 and 10D2 further details on<br />

the processes of solid-solid phase transition and melting can be derived.<br />

First, we consider the 2D2 molecule. Of necessity, the process of packing of traizsplanar<br />

chains to forni an orthorhonibic crystalline lamella cannot distinguish<br />

deuterated and undeuterated chain ends; the resulting lamella can be described by<br />

a random distribution of ‘up’ or ‘down’ chains (Figure 3-55). With such chains we<br />

can distinguish the conformational behavior of the heads and/or the tails of the<br />

chains. It becomes thus possible to be sure of the origins of the signals coming from<br />

conformational defects.<br />

The most direct information comes from the frequencies of the CD? rocking<br />

which have been calculated with normal coordinate calculations and are reported in<br />

Table 3.1.<br />

The infrared spectrum of 2D2 in the 670-600 cm-’ range shows only a band<br />

near 620 cm-’ below To (Figure 3-56). The band is not split, as mentioned above,<br />

because the precise phase coupling is lost by the random distribution of ‘up’ and<br />

‘down’ chains. At the solid-solid transition a very weak additional band appears<br />

near 655 cm-’ (Figure 3-56) indicating the formation of gauche conformations at<br />

the deuterated chain ends, thus reiterating what has been already learned from C19.<br />

From a quantitative measure of the absorption intensities of the these modes it<br />

transpires that approximately 25% of chain ends are conformationally distorted.<br />

The infrared spectrum of 10D2 provides further information (Figure 3-57]. For<br />

the orthorhonibic phase the line near 620 c1n-l is split because in this case the phase<br />

coupling is achieved since translational symmetry is obtained within the crystal. At


30.5 'C<br />

Figure 3-56. Temperature dependence of the infrared<br />

\..:-A<br />

3, 6.C spectrum of n-nonadecane-2D2 through orthorhombic<br />

+ c( + phase transitions. Notice that (i) in the a phase a<br />

small number of deuterated chain heads has taken up the<br />

L.:-32Pc<br />

TG conformation which increases dramatically in the<br />

I " " ' J<br />

melt, (ii) in the melt a certain number of deuterated chains<br />

700 660 620 580<br />

v cm-1 do not tilt their heads but still keep the TT conformation.<br />

T, the doublets disappear since chains move and become uncorrelated (isolated).<br />

Since no absorption is observed near 655 and 677 cn-' it means that in the middle<br />

no conformational distorsions are generated in the a phase. The liquid-like structure<br />

is, however, observed at the melting.<br />

Conclusion no. 5<br />

The final model emerging from this part of the study is the following: starting from<br />

a regular and ordered orthorhombic structure in which all tmns-planar chains are


3.19 A Worked E.vnn~yle 193<br />

Figure 3-57. Temperature dependence of the infrared<br />

spectrum of n-nonadecane-1 OD: through orthorhonibic<br />

+ c( phase transitions. Notice that (i) the orthorhombic<br />

crystal disappears at 22.4 "C, (ii) in the c( phase at the center<br />

of the lamella no confomiational kinks are formed<br />

while they certainly appear in the melt (33.2"C) Combination<br />

of the data of this figure and of Figure 3-56 are direct<br />

evidence of 'surface melting'.<br />

arranged in a crystal with a setting angle of 42" at T, the lattice expands anisotropically,<br />

interlamellar forces weaken and a conformational disordering takes<br />

place at the surface of the lamella. Such disordering is generated by the conformational<br />

deformation of approximately 25% of the chains which tilt their head with<br />

gauche conformation either in position 2-3 or 3-4. The inside of the lamella is not<br />

perturbed and all chains are still trans-planar (Figure 3-58).<br />

The next question which remains to be answered is whether surface disordering is<br />

generated by the longitudinal motion of the chains. The following experiment has<br />

been made [155].


I94<br />

3 k'ibrcitionul Specfra CIS a Probe of Structural Order<br />

a<br />

ORTH<br />

a<br />

\N\ 3 ~<br />

y.\\\ \ \ '<br />

b<br />

Figure 3-58. Sketch of the mechanism of phase transition of n-nonadecaiie based on detailed<br />

spectroscopic evidence. (a) Schematic overall view; (b) details on a molecular level.<br />

Let us take two samples of crystalline orthorhombic n-alkanes with an odd<br />

number of carbon atoms. They show properties identical to those observed for C19.<br />

Let us make by suitable methods a mechanical mixture say of C21 and C23. The<br />

infrared spectra of the mechanical mixtures at low temperature are just the superposition<br />

of the spectra of the two independent components. Let us focus on the<br />

crystal field splitting, say of the doublets near 940 cn-' (C23) and 925 cin- ' (C21 j<br />

(Figure 3-59). If the mechanical mixture is brought near the solid-solid phase transition<br />

the infrared spectrum evolves with time. Time- and temperature-depeiideiit<br />

spectra show that the doublets become singlets thus showing that chains have<br />

moved (Figure 3-59). The final product is a mixed crystal of C21 and C23 as also<br />

proved by parallel DSC studies. These studies were first reported by Ungar and<br />

Keller [ 1561 and were re-examined and interpreted as reported in [ 1551.


3.19 A Worked E.vnrqde 195<br />

009.1<br />

0084<br />

0<br />

n<br />

b 0074<br />

$00641<br />

00541 I I 1 I I I<br />

960 950 94C 930 92C 910 900<br />

&a ve 1 eng t n (c m ’’ 1<br />

Figure 3-59. Time-dependent infrared spectruni in the 960-900 cm-’ range showing the Pi5 component<br />

of the CHz rocking progression band of n-CrlH44 and the P17 component of n-Cr3HdZ.<br />

Evolution with annealing time. Annealing temperature 30T, A(t = 0); B (t = I hr); C (t = 2 h); D<br />

(t = 4 h); E (t = 6 h). At t = 0 the sample consists of a mechanical mixture of the two materials. At<br />

t = 0 both solids are crystalline orthorhombic as shown by the correlation field splitting doublet. By<br />

increasing the time of annealing both doublets coalesce into a singlet indicating that chains have<br />

migrated and formed a mixed crystal.<br />

Figure 3-60 shows the kinds of kinetic studies of the mixing process which have<br />

been made.<br />

Conclusion no. 6<br />

In the a phase chains move longitudinally and are ‘squeezed out of the lamella’; the<br />

onset of such motion occurs near the solid-solid transition temperature.<br />

This is in agreement with recent measurements of quasi-elastic neutron scattering<br />

by Gullaume et al. on C19 at 300°K who determined the existence of translational<br />

longitudinal diffusion with discrete steps of one or two CH2 units with a measured<br />

diffusion coefficient D = 6 x lop6 cm’s-l [157]. Statistical thermodynamics calculations<br />

give theoretical support to this kind of phenomenon [ 1581. These conclusions<br />

find a striking support by the calculations of molecular dynamics by Klein [159]<br />

which describe precisely all the details that we have found from spectroscopy.<br />

Moreover, recent experiments of temperature-dependent scanning tunnelling microscopy<br />

by Rabe et al. [160] give just ‘photographic’ support at the molecular level of<br />

the process as described by spectroscopy, namely: (i) first the chains organized in<br />

an ordered lattice; (ii) then the onset of longitudinal diffusion which generates


196 3 IG!wtioiinl Specfrri us N Probe of Sfiw-tiird Order<br />

tan,, t Hours)<br />

Figure 3-60. 1 : 1 molar mixture<br />

of n-CzJHqg and n-C.5HF2. Kinetics<br />

of the process of diffusion<br />

at four different temperatures.<br />

Data from calorimetric<br />

measurements<br />

conformational deformations at chain ends; (iii) diffusion which generates the<br />

interpenetration of chains; and (iv) the full collapse towards a liquid-like structure.<br />

Conclusion no. 7<br />

As a consequence of such longitudinal motion, if crystallites of n-alkanes are suitably<br />

brought together thermally activated transport of matter can take place and<br />

diffusion of molecules from one crystallite to another can occur. The kinetic of<br />

diffusion is function of temperature, length of the chain, contacts between crystallites<br />

(ie., pressure put on the sample), etc.<br />

We can proceed further and try to understand the mechanism at the molecular<br />

level which drives the chains outside their own lamella by a thermally activated<br />

process.<br />

We refer the reader to [ 1031, [ 1551 and [ 1451 for a first analysis of the various data.<br />

We simply mention the line of thought followed by various authors on this rather<br />

complex problem.<br />

Several theoretical models have been proposed in an attempt to account for the


3.19 A Worked Example 197<br />

existence of the c( phase and for the longitudinal motion. First, the general idea<br />

was that in the a phase chains rotate rigidly [152-1541. Next, models of rotational<br />

diffusion coupled with longitudinal translations of one CH2 were proposed based<br />

on NMR data [ 161, 1621. Elastic neutron-scattering experiments introduced the<br />

idea of diffusional cooperative rotations or of independent translational motions<br />

with jump length of one CH;! unit followed by 180" rotational jump [163]. A<br />

translation of two CH? units was proposed [164] in agreement with calculations<br />

by McCallough [165]. Most of the models proposed consider that in the solid state<br />

chains perform large-amplitude collective rototranslational motions. McCullough<br />

has presented an interesting model which shows by calculations that chain can find<br />

a favorable energy path if they librate rigidly with large amplitude librations and<br />

translate along the chain axis.<br />

The knowledge at the molecular level of the mobility of chain molecules in the<br />

solid state is of basic importance for the understanding of the general processes of<br />

annealing of chain molecules. Attempts have been made to derive information using<br />

band shapes in the Ranian spectra, especially focusing on the CH-stretching vibrations<br />

(Section 3.18).<br />

As already stated at the end of Section 3.18, we have measured the temperaturedependent<br />

band widths and the frequency of the CH2 d- mode for several crystalline<br />

n-alkanes and polyethylene and found no sizable broadening or upward frequency<br />

shifts until they approach the melting point. All the theoretical models that<br />

imply large amplitude librotorsional motions are not supported by the spectroscopic<br />

criteria discussed in Section 3.18. We found evidence of large-amplitude<br />

libro-torsional oscillations only when the n-alkane chains are included as clathrates<br />

in various systems.<br />

Conclusion no. 8<br />

Based on the analysis of teniperature-dependent Ranian band shape and frequency<br />

shift of the Raman active antisymmetric CH1-stretching modes (Section 3.18),<br />

contrary to the case of urea clathrates, in the solid crystalline state (including the CI<br />

phase), n-alkane chains do not perform large-amplitude libro-torsional motions.<br />

We must search for alternative molecular models which may account for longitudinal<br />

diffusion with transport of matter without implying large amplitude librotorsional<br />

oscillations. A conceptually and, at least, intellectually exciting model<br />

which may help in this search has been presented by Mansfield and Boyd [166] who<br />

attempted to account for molecular relaxation phenomena. We label this model as<br />

Utah-twist. The main message by these authors is that calculations suggest that an<br />

orthorhombic lattice of n-alkanes can host a chain distorted by a gentle and long<br />

twist. The long twist is achieved with a uniform succession of small torsions about<br />

the C-C bonds which is extended over approximately 12-15 CH2 units. At the end<br />

of the gentle twist the chain is flipped by 1800 with respect to the head of the twist.<br />

Such a long twist (which we label as 'twiston') can propagate along the polymethylene<br />

chain without finding a barrier to its propagation. Based on the aton-atom<br />

potential chosen by the two above authors the activation energy for generating a


198 3 JJiber-ulioiiiil <strong>Spect</strong>ra as a Probe oj Structural Ordeer-<br />

Figure 3-61. Projection onto the ab face of the structuic<br />

I 1<br />

of the orthorhombic cell of polyethylene (two molecules<br />

per unit cell, setting angle o = 42").<br />

twiston is estimated to be -10 kcal/mol. The twiston model implies for a polymer:<br />

(i) the initial formation of a disordered region near the surface of the crystal; (ii) the<br />

propagation of the twiston through the crystal along the chain axis; and (iii) the<br />

disappearance of the twiston at the opposite surface leaving the chain rotated by<br />

180" and translated by one CH2 unit.<br />

The idea is formally very appealing since, from the spectroscopic viewpoint,<br />

it implies a collective mobility of the polymethylene chain which accounts for the<br />

observed longitudinal mobility, surface melting, or disordering and transport of<br />

matter without requiring large-amplitude libro-torsional motions which are not<br />

detected by vibrational spectroscopy, as previously discussed.<br />

The twiston model has been further expaiided by Mansfield [I671 and later by<br />

Skinner and Wolynes [168]. We shortly mention the physics behind since it may be<br />

applied to other cases in polymer physics. The whole problem of polynielhylene<br />

chains is reviewed in [169].<br />

Let us take the structure of the orthorhombic unit cell of polyethylene (Figure<br />

3-61) and consider the CH2 units as point masses of mass 117 at a distance ci from the<br />

axis of the molecule. The units can rotate by an angle 0 about the molecular axis.<br />

When they rotate they are subject to the elastic response C by the torsions between<br />

two adjacent (CH?) units. From Figure 3-61 it is apparent that when the chain<br />

rotates as a rigid body about its axis in an orthorhombic lattice it probes a twofold<br />

periodic potential V(0) from the crystal field.<br />

The Hamiltonian proposed by Mansfield has the following form<br />

In Eq. (3-61) the first term represents the contribution from the crystal potential,<br />

the second term the kinetic energy, and the third is the quadratic elastic contribution.<br />

In a gentle twist it can be assumed that Oi do not change much from one unit<br />

to the other. According to Mansfield one can make the continuous approximation<br />

and reach a solution of the type:<br />

@(x - xo - vt) = 4 arctan exp [ k y(x - xo - vt)] (3-62)


3.1 9 A Worked Example 199<br />

The soliton solution gives the values of Oi which describes a twist in the chain<br />

which, in the absence of friction, propagates keeping the same shape and velocity v.<br />

The Utah-twist (twiston) is then approximated to a solitonic wave whose velocity<br />

of propagation is estimated by Mansfield (using some molecular parameters) to be<br />

- 10j cm s-I.<br />

The final concept which emerges is that in the orthorhonibic crystal one can<br />

conceive the existence of conforniational collective excitations thermally generated<br />

which extend over approximately 20 CH2 units and move along the chain with a<br />

certain velocity.<br />

The main problem is to prove the existence of such mobile twistons. The idea of<br />

using spectroscopy implies the fact that we should be able to detect somewhere in<br />

the vibrational spectrum absorption and/or scattering associated to vibrations<br />

localized on the twiston. Calculations have been made (i) of the density of vibrational<br />

states of a long polymethylene chain containing a random distribution of<br />

twistons and (ii) of the infrared spectrum (frequencies and intensities) of the molecule<br />

C19 containing a twiston. A comparison has been made with the infrared<br />

spectrum of C19 in the all-frans structure [169] (Figure 3-62).<br />

in these calculations the twiston is assumed not to be mobile, but pinned at a<br />

given site of the chain. The fact that the twiston is mobile along the chain in principle<br />

does not deny the possibility, under certain conditions, to observe its vibration<br />

in the infrared/Raman spectra. The conditions are the following:<br />

(a) Since the band width is inversely proportional to the vibrational lifetimes of<br />

each normal vibration, if the velocity of the twiston is small the vibrational<br />

lifetime is long enough to allow the formation of its own vibrational mode and<br />

can give rise to an observable IR or Raman band. if the twiston moves too fast,<br />

the vibrational lifetimes are too short and the band broadens and flattens in<br />

a broad continuum. Simple calculations show that if a band of the twiston has<br />

to be observed with Ai~,p M 20 cin-' it sets an upper limit to the velocity of the<br />

twiston of = lo4 cm-' which is not too different from the velocity estimated by<br />

Mansfield.<br />

(b) The concentration of twistons has to be large enough to generate a detectable<br />

absorption in IR or scattering in the Raman. The concentration of twistons<br />

depends on their energy of formation which has to be low if twistons are to<br />

be seen spectroscopically. The energies calculated for the Utah-twist [ 1661 are<br />

much too high and would make the twiston undetectable by spectroscopic<br />

techniques.<br />

The results of the theoretical calculations of the spectra of twistons and a few<br />

attempts to catch spectroscopic signals possibly associated to twistons are reported<br />

in [169]. The evidence collected so far is neither compelling nor completely negative.<br />

The model needs to be supported or dismissed by other physical techniques.<br />

The model of twistons which move as 'solitary waves' was also proposed by<br />

Dwey-Aharon et al. [ 1701 in the case of polyvinyljdene fluoride (CHz-CF?),, which<br />

upon poling transforms from the helical structure (TGTG') (labeled either as p<br />

form or form 11) into a planar all-tram one (a, or form I). Form I is technologically


200 3 Vibrutional <strong>Spect</strong>ra (IS CI Probe of Str'zictural Order<br />

Figure 3-62. Searching for twistons: section of the calculated spectruin of ti-nonadecane (-) in the<br />

all-trrrns planar conformation; (- - -) with the conformational distortion like a soliton according to<br />

Mansfield and Boyd. Both frequency and intensity were calculated.


3.20 References 201<br />

very important, since having a strong dipole perpendicular to the chain axis shows<br />

remarkable pyro and piezoelectric properties widely used in modern technology.<br />

According to Dwey-Aharon when the conformationally disordered material existing<br />

in form I1 is placed in a strong electric field (poling) conformational twistons are<br />

generated which propagate and yield form I. The spectroscopy of polyvynilidene<br />

fluoride and the search for twistons has been treated in [171].<br />

Finally, the concept of twiston has been again introduced in a spectroscopic work<br />

to account for the set of peculiar experimental data collected in the study of the<br />

phase transition in ethylene-tetrafluoroethylene alternating copolymers [ 1721.<br />

Acknowledgments<br />

The contributions to knowledge presented in this chapter are the result of an<br />

enthusiastic collaboration of many researchers and students who worked in our<br />

group at the Politecnico of Milano. Our warmest thanks are extended to Dr. M.<br />

Gussoni, who laid the background for the development of this science.<br />

3.20 References<br />

[l] H. H. Nielsen, Ha?idbzrclz der Plzvsik, Springer, Berlin, Vol. 37/1, 1959.<br />

[2] G. Herzberg, Infrared and Ranian <strong>Spect</strong>ra of’ Polyatonzic Molecules, Van Nostrand, Princeton,<br />

USA, 1963.<br />

[3] E. B. Wilson, J. C. Decius and P. C. Cross, Moleculur I/ibrutioiis, McGraw Hill, New York,<br />

1955; V. Volkenstein, M. A. Eliashevich and B. Stepanov, Kolebuni.vu Molecul 11, Moscow,<br />

1949; S. Califano, Vibrational Stutes, Wiley, New York, USA, 1976.<br />

[4] For a general discussion see: W. Person and G. Zerbi (Eds.), Vibrotional61teizsities in Infrared<br />

and Ranm <strong>Spect</strong>roscopy, Elsevier, Amsterdam, The Netherlands, 1984; L. A. Gribov, Intensity<br />

Theory for IIfrared <strong>Spect</strong>ra of Polyatoniic Molecarles, Consulting Bureau, New York,<br />

USA, 1964.<br />

[5] M. Gussoni in Aduances in hfiared and Ru~nan <strong>Spect</strong>roscopy, (Eds. R. J. H. Clark and R. H.<br />

Hester) Heyden, London, England, 1979, vol. 6 p. 96.<br />

[6] M. Gussoni, C. Castiglioni and G. Zerbi, J. Mol. Strzrct, 1989 198, 475 M. Gussoni, C.<br />

Castiglioni, M. N. Ramos, M. Rui and G. Zerbi, J. Mol. Strirct. 1990 224, 225 C. Decius, J.<br />

Mol. <strong>Spect</strong>ry, 57, 384 1975; A. J. Van Straten and W. Smit, .I Mol. <strong>Spect</strong>ry., 1976 62, 297 M.<br />

Gussoni, C. Castiglioni and G. Zerbi, J. Phys. Clienz. 1984 88, 600.<br />

[7] S. Trevino and H. Boutin, J. Macronzol. Sci., 1967 Al, 723.<br />

[8] E. Curtis, Thesis, University of Minnesota, 1958.<br />

[9] J. H. Schachtschneider and R. G. Snyder, <strong>Spect</strong>rocki/n. Acro, 1963 19, 17.<br />

[lo] The most known set of original computing programs which has later originated many other<br />

programs for dynamical and force constant calculations is that written by J. H. Schachtschneider,<br />

Shell Rep. 57/65.<br />

[ 111 See, for instance, J. L. Duncan, Force Curisfunt Calcirlcrtions ik hfokecirles, Specialist Report,<br />

Chemical Society, London, 1975, vol. 23, p. 104: G, Fogardsi and P. Pulay, in I’ihr.trtiorm/<br />

<strong>Spect</strong>ru cind Strircfure, (Ed. J. R. Durig) Elsevier, Amsterdam, 1985, Vol. 14, p. 125.


202 3 Vibrutioncil <strong>Spect</strong>ra CIS a Probe of Striicturnl Odcr<br />

[I21 M. Gussoni, C. Castiglioni and G. Zerbi, J. Phys. Clzenr., 1984 88, 600.<br />

[I31 M. Gussoni, M. Ramos, C. Castiglioni and G. Zerbi, J. Mol. Struct., 1988 174, 47.<br />

[ 141 P. Pulay, X. Zhou and G. Fogarasi, in Rewit E~~~~ei~inientrd and Co/iif~/ittrtion[/I Atluirrrc,cc. i~7<br />

Molecular <strong>Spect</strong>roscopy, (Ed. R: Fausto), NATO AS1 Series C, Vol. 406, p. 99; W. B. Person.<br />

K. Szczesniak and J. E. Del Bene, ibidem, p. 141.<br />

[I51 K. Palmo, L.-0. Pietila and S. Krimm, Conzpurers Clienz., 1991 f>, 249; K. Palmd, L.-0.<br />

Pietila and S. Krirniii. .I. Comp. Cl~eni., 1991 12, 385; K. Palmo, N. G. Mirkin, L.-0. Pieti13<br />

and S. Krimni, M~momulecziles, 1993 26, 6831, K. Palmo, L.-0. Pietila and S. Krimm,<br />

computer.^ Clieni., 1993 17, 67; K. Palmo, L.-0. Pietila and S. Krimni, J. Comp. Clwnr.. 1992<br />

13, 1142.<br />

[16] G. Zerbi, Vihrcitional <strong>Spect</strong>ra qf High Pol~wers, (Ed. E. G. Brame), Applied <strong>Spect</strong>roscopy<br />

Reviews. Dekker, New York, USA, 1963, vol. 2, p. 193; H. Tadokoro; Srrirc.trrw uf'Crj~stir1-<br />

lii7e Pol~~ers, Wiley, New York. 1979.<br />

[I71 P. C. Painter, M. M. Coleman and J. L. Koenig. Tlie Tlreury of t'ihutiond S)c,c.fro.cc'opj7' ctr7d<br />

its Applications to <strong>Polymer</strong>ic Materials, Wiley, New York, USA, 1982.<br />

[I81 For a general discussion of the dynamics of simple and simplified chain molecules see: R.<br />

Zbinden, Znjiaved <strong>Spect</strong>roscopy of High Poljvners, Academic, New York, USA. 1964.<br />

[ 191 G. Zerbi, Vibrational Sliectroscopy of Very Largt, molerirles, in Advances in Infrared and<br />

Rainan <strong>Spect</strong>roscopy (Eds. R. J. H. Clark and R. H. Hesterj Wiley-Heyden, New York,<br />

USA, vol. 11, p. 301.<br />

[20] J. C. Decius and R. M. Hexter, Moleciilar I4bratians in Crystcils, McGraw-Hill, New Yorb,<br />

1977; S. Califano, V. Schettino and N. Neto, Lattice Dynnmics of i2.loleciilrrr Cr~.sfa/s. Lectut-e<br />

Notes in Chemistry, Springer, Berlin, 1981.<br />

1211 M. Gussoni, G. Dellepiane and S. Abbate, J. Mol. <strong>Spect</strong>ry., 1975 57, 321.<br />

[22] For a discussion on the problem of redundancies see: M. Gussoni and G. Zerbi, Accademia<br />

Nazionale dei Lincei, Rendiconti serie VZZI 1966 40, 843; M . Gussoni and G. Zerbi, ibid.<br />

p. 1032; M. Gussoni and G. Zerbi, Clieni. PIiys. Lett., 1968 2, 145: I. M. Milk Clwni.<br />

Pliys. Lett. 1969 3, 267; G. Zerbi, in <strong>Modern</strong> Trends in I'ibrationul <strong>Spect</strong>roscop?', (Eds.<br />

A. J. Barnes and W. J. Orville-Thomas) Elsevier, Amsterdam. 1977. p. 261: M. V. Volkenstein,<br />

L. A. Gribov, M. A. Eliashevich and B. Stepanov, Kolehzniyci Molt~kul, Moskow, 1972,<br />

p. 203.<br />

[23] N. B. Colthup, L. H. Daly and S. E. Wiberley, Introdwtioii to Znfifi.arei/ and Xtnncirr S,ut*c./roscopy<br />

Academic, New York, 1975.<br />

[24] R. N. Jones and C. Sandorfy, Clienrical Applicntions uf Spec~troscopj~, in Teclmiques of<br />

Organic Chemistry, (Ed. A. Weissenberger) Interscience, New York, USA, 1956, vol. IX.<br />

[25] L. Bellaniy, The hfr~irt~l Slwctriz of COn711kY Moleciiles, Wiley, New Yo]-k, USA, 195s.<br />

1261 K. Nakamoto, /nj?ared <strong>Spect</strong>ru of Polyatornic Molecules, Wiley, New York, 1963.<br />

[27] See the series of papers such as, for example, T. Shinianouchi, Tables qf Moleczrltrr Frcquencies,<br />

Part 7, J. Physicul and Cheniical Reference Dota, 1973 2, 225.<br />

[28] For the case of a polymer see, for instance, G. Masetti, S. Abbate, M. Gussoni and G. Zerbi.<br />

J. Clzern. Pliys. 1980 73, 4671.<br />

[29] M. Gussoni and G. Zerbi. J. &lo/. <strong>Spect</strong>ry., 1968 26, 485.<br />

[30] G. Dellepiane, M. Gussoni and G. Zerbi, J, Chern. Plqis. 1970 53, 3450; M. Gussoni, G.<br />

Dellepiane and G. Zerbi, in Molecular Structures and I'lbrations (Ed. S. J. Cyvin) Elsevier,<br />

Amsterdam, 1972, p. 101.<br />

[31] P. Torkington, J. Cliern. Phylys., 1949 17, 347.<br />

[32] Y. Morino and K. Kuchitsu, J. cheni. Plz-vs., 1952 20, 1809.<br />

[33] J. H. Schachtschneider and R. G. Snyder, <strong>Spect</strong>rochini. Actu 1963, 19, 17.<br />

[34] R. G. Snyder and J. H. Schachtschneider, Sj7ectrocliim. Acta. 1963 19, 117.<br />

[35] J. H. Schachtschneider and R. G. Snyder, <strong>Spect</strong>rocliim. ACIU, 1965 21, 1527.<br />

[36] R. G. Snyder, J. Clwrn. Phys., 1967 47, 1316.<br />

1371 R. G. Snyder and G. Zerbi, <strong>Spect</strong>roclrini. Acta, 1967 23 A, 391.<br />

[38] W. T. King and B. L. Crawford, J. Mol. <strong>Spect</strong>ry., 1960 5, 429.<br />

[39] J. Overend and J. R. Scherer, Spc.c.trocAini. Actrt, 1960 16, 773.<br />

[40] Handbook of Conducting Polyniers, (Ed. T. A. Skotheini) Dekker? New York. 1986. Vols. I<br />

and 2.


3.20 References 203<br />

[41] M. Gussoni, C. Castiglioni and G. Zerbi. in SjJNti'o.SCOpy qf rlduarrced Mciterinls (Eds. R. J.<br />

H. Clark and R. E. Hester) Wiley, New York, 1991, p. 251.<br />

[42j G. Zerbi. M. Gussoni and G. Castiglioni. in Coi+tyaterl Po/ym,r.s, (Eds. J. L. Bredas and R.<br />

Silbey) Kluwer, Netherland, 1991, p. 435.<br />

[43] C. Castiglioni, M. Del Zoppo and G. Zerbi, .I R~mm <strong>Spect</strong>ry.. 1993 24. 485.<br />

1441 We consider as almost unique example of the existence of long range interactions in covalent<br />

systems the superlinear increase with conjugation length of the second order hyperpolarizability<br />

y in long chain polyenes as reported by: I. D. W. Samuel. 1. Ledoux, C. Dhenaut, J.<br />

Zyss, H. M. Fox; R. R. Scrock and R. J. Silbey, Science, 1994 256, 1070.<br />

[45] G. Natta and P. Corradini, Nuotjo Cinzeiito, Siqpl., 1960 15, 40.<br />

[46] B. Wunderlich. hlcicrortzoleciilur Physics, Vols I and 2, Academic, New York, 1976.<br />

(471 See, for instance, P. De Sanctis, E. Giglio, A. Liquori and A. Ripanionti, J. Po/yrn. Sci., 1963<br />

Al; 1383.<br />

[48J P. Flory. Statisticcil Mechanics of Chciirz Mokecdes, Interscience, New York, 1969.<br />

[49] M. Tasumi and T. Shimanouchi, J. Cheiii. Phys, 1965 43, 1245; M. Tasumi and S. Krimm, J.<br />

Cherii. Phys., 1967 46, 755.<br />

[50] L. Piseii and G. Zerbi, J. Mol. Spctry., 1968 26, 254.<br />

[51] A. A. Maradudin, E. W. Montroll and G. H. Weiss, Solid State Pliys., Si/pp/., 1963 3, 1.<br />

[52J L. Brillouin, HJaoe Propagation in Periodic Strrrctures, Dover, New York, 1953.<br />

[53] G. Zerbi, in Advances in Infrared and Raman Sprcfroscop-v, (Eds. R. J. H. Clark and R. E.<br />

Hester) Wiley, 1984, p. 301.<br />

[S4] G. Zerbi and M. Sacchi, Mucromolecrilrs, 1973 6, 692.<br />

[55] P. Bosi, R. Tubino and G. Zerbi, J. Clzeriz. Phys.? 1973 59, 4578.<br />

[56] L. Piseri and G. Zerbi, J. Chern. Phys., 1968 4S, 3561; G. Zerbi and L. Piseri, J. Chew?. Phys.,<br />

1968 49. 3840.<br />

1571 For Calculations of dispersion curves of Hydrogen bonded systems in one dimension: R.<br />

Tubino and G. Zerbi, J. Chem. Phj..s., 1969 51, 4509; A. B. Dempster and G. Zerbi, J. Cliem<br />

Pliys., 1971 54, 3600; R. Tubino and G. Zerbi, .I. Cheni. Phys. 1970 53, 1428.<br />

[58] G. Zerbi, C. Castiglioni and M. Del Zoppo. in Conducting <strong>Polymer</strong>: The Oligorner Approuch<br />

(K. Mullen and G. Wegner Eds. ) VCH, Keinheim (1998).<br />

[59] C. Castiglioni, M, Gussoni and G. Zerbi, J. Cheni. Phis., 1991 95, 7144.<br />

[60J R. Tubino, L. Piseri and G. Zerbi. J. Chenr. Phys., 1972 56, 1022.<br />

1611 P. W. Higgs, Proc. Roy. Soc., London, 1953 11220, 472.<br />

[62] T. Miyazawa, J. Clieni. Phys., 1961 35, 693.<br />

[63] G. Masetti. S, Abbate, M. Gussoni and G. Zerbi. J. C/zenz. P/z,vs., 1980 73, 4671.<br />

[64j B. Orel, R. Tubino and G. Zerbi, Mo1ec~rlrrr Plz~x, 1975 30. 37.<br />

[65] G. Zerbi, F. Ciampelli and V. Zaniboni, J. PO/.VJJ. Sci., Ptrrt C, 1964 141.<br />

[66] G. Zerbi, Probiiig the Real Struetiire of Chciin Molecules by Vibrational <strong>Spect</strong>roscopy? American<br />

Chemical Society, Specialist Reports. 1983 203, 487.<br />

I671 T. Miyazawa and T. Kitagawa, J. Polwn. Sci., 1965 B2, 395; S. Enomoto and M. Asahina,<br />

Pol.vm. Sci., 1964 A2, 3523.<br />

[68] R. G. Snyder, J. Mol. <strong>Spect</strong>ry., 1961 7, 116; R. G. Snyder, in Methods of E.xperinieiitd<br />

Physics (Ed. R. A. Fava). Academic, New York, Vol. 16, part A, p. 73.<br />

[69] S. Krimm, Achmce.r 117 Poljvrier Scieme, 1960 2, 51.<br />

[70] D. Dows, in Plivsics mid Chrrnisfry of Orgarlic Solid State, (Eds. 1. Fox, M. M. Labes and<br />

A. Weissberger) Wiley-Interscience, New York, 1963, Vol. 1; W. Vedder and D. F. Hornig,<br />

Aehmzces in Spec~troscoly, 1962 2, 189; M. I. Bank and S. Krimm, J. Appl. Pliy,~., 1969 10,<br />

248; S. Kriinin and T. C. Cheam. Fmwluy Disc. Chem. Soc. 1979 68, 243.<br />

[71J V. Zamboni and G. Zerbi, J. Po!,?n. Sci., 1964 C7, 153: G. Zerbi and G. Masetti, .J. Mol.<br />

<strong>Spect</strong>ry, 1967 22, 284; G. Zerbi and J. P. Hendra. J. Mol. <strong>Spect</strong>ry., 1968 27, 17.<br />

[72] For Polytetrafluoroethylene: F. J. Boerio and J. L. Koenig, J. them. Phj.s., 1971 54, 3667<br />

and C. J. Peacock, P. J. Hendra, H. A. Willis and M. E. A. Cudby, J. C/ic/i/. Soc. A. 1970<br />

2943.<br />

[73] For Isotactic Polypropylene, ref. 66, p. 501, and G. Masetti, unpublished work.<br />

1741 M. Peraldo, Giizz. Chi~tr. Ztal, 1959 89, 798; M. Peraldo and M. Farina, Chi///, Did iMilaii),<br />

1960 4-3, 1349.


204 3 Vibrational <strong>Spect</strong>ra [is (i Prohe of Structiirul Ordeer,<br />

[75] G. Zerbi, M. Gussoni and F. Ciampelli, <strong>Spect</strong>r~ochini. Actci, 1967 23, 301.<br />

[76] G. Masetti, F. Cabassi and G. Zerbi, <strong>Polymer</strong>, 1980 21, 143.<br />

[77] G. Natta and P. Corradini, Atti Accad. Nazle. Lincei, Mein. Ckrisse Sci. Fis. Mu/. Nor., 1955<br />

4, 73: G. Natta, P. Corradini and M. Cesari, ibici.. 1960 21, 365.<br />

[78] G. Zerbi, Sperialist Reports 1983, 203, 501.<br />

1791 H. JeKreys and B. S. Jeffreys, Method3 of Mertliemuti~d Physics, Cambridge, New York,<br />

1950, p. 40; P. Dean and J. L. Martin, PIX Roy. Soc. (London), 1960 A259, 309; P. Dean,<br />

Rev. Mod. Pliys., 1972 44, 127.<br />

[80] A. Rubcic and G. Zerbi, il/lnrrorrio/ecirles. 1974 7, 754.<br />

[81] W. Myers, J. L. Donovan and J. S. King, J. Cheni. Phys., 1965 42, 4299; G. C. Summerfield,<br />

J. Chem. Pliys., 1965 43, 1079.<br />

[82] A. Sjolander, Arkio. Fysi/c, 1958 14, 315.<br />

[83] L. V. Tarasov, Soviet P1iys.-Solid State, English Transl., 1961 3, 1039.<br />

[84] S. Trevino and H. Boutin, J. Macrornol. Sci., 1967 Al, 723.<br />

[Sj] W. Myers, G. C. Summerfield and J. S. King, J. Chem. Phys., 1986 44, 184.<br />

[86] S. F. Parker J. Chem. Soc., Faradq Tram> 1996 92, 1941.<br />

[87] Faraday Discussions of the Chemical Society Orpnixifion of Miicronio/ec,/iles in the Condensed<br />

Phase, No. 68, Roy. Soc. Chem. 1979.<br />

[88] G. Zerbi in Phonons, (M. A. Nusimovici Ed.), Proceedings of the international conference,<br />

Rennes, France 1971, p. 248.<br />

[89] G. Zerbi and G. Cortili, Chern. Comnt, 1965 13, 295; G. Cortili and G. Zerbi, <strong>Spect</strong>rochini.<br />

Actrr, 1967 33A, 285.<br />

1901 A. Rubcic and G. Zerbi, hfncrornolecules, 1974 7, 759.<br />

[91] M. Tasumi and G. Zerbi, J. Cheni. Phys. 1968 48, 3813.<br />

[92] B. Fanconi, Ann. Rev. Pliys. Clien?. 1980 31, 265.<br />

[93] C. Schmid and K. Holzl, J. Polyni. Sci., Pliys. Ed.., 1972 10, 1881; C. Schmid and K. Holzl, J.<br />

Phys. C, 1973 6, 2401; V. N. Kozyrenko, L. Kumpanenko and V. Mikhailov, J. Polyrn. Sci..<br />

Polpi. Pliys. Ed., 1977 IS, 1721.<br />

[94] G. Zannoni and G. Zerbi, J. Mol. Struct., 1983 1000, 385. See also J. J. Ladik, Qucrntuni<br />

Theory qfSolids, Plenum, New York, 1987, p. 140.<br />

[95] G. Zerbi, L. Piseri and F. Cabassi, Mo/. Plzys., 1971 2-3, 241.<br />

[96] J. H. Wilkinson, Computer J. 1958 I , 90.<br />

[97] M. Gussoni and G. Zerbi, J. Cheni. Phys., 1974 60, 4862.<br />

[98] I. F. Chang and S. S. Mitra, Atlo. Phys., 1971 20, 359.<br />

[99] G. Zerbi and M. Gussoni, <strong>Polymer</strong>, 1980 21, 1129.<br />

[ 1001 A. Rubcic and G. Zerbi, Macromolecules, 1974 7, 759.<br />

[I011 M. Tasumi and G. Zerbi, J. Cheni. Phys., 1968 48, 3813.<br />

[I021 R. G. Snyder and M. Poore, Macromolecules, 1973 6, 709.<br />

[103] G. Zerbi, R. Magni, M. Gussoni, K. Holland-Moritz, A, Bigotto and S. Dirlikov, J. Chem<br />

Phys., 1981 75, 3175.<br />

[I041 For the works by the school of R. G. Snyder in this field see, for instance, M. Maroncelli, S.<br />

P. Qi. H. L. Strauss and R. G. Snyder, J. Am. Chem. Soc., 1982 104, 6237.<br />

[I051 A. Keller, Furaday Disc. Cllem. Soc. 1979 68, 244.<br />

[lo61 G. Bucci and T. Simonazzi, J. Polynz. Sci., port C. <strong>Polymer</strong> Symposia, 1964 7, 203.<br />

[107] G. Zerbi, G. Minoni and A. P. Tulloch, J. Chn. Phys.. 1983 78, 5853.<br />

[I081 G. Vergottin, G. Fleury and I. Moschetto, Advvrrnces i7 hqiared and Ramari <strong>Spect</strong>roscopy,<br />

(Ed. R. J. H. Clark and R. E. Hester) Heyden, London, 1978, vol. 4, p. 195.<br />

[I091 G. Zerbi, G. Conti, G. Minoni, S. Pison and A. Bigotto, J. Phys. Chem., 1987 91, 2386.<br />

[I101 G. Minoni and G. Zerbi, J. Phys. Cliem., 1982 96, 4791.<br />

[ 11 I] M. Del Zoppo and G. Zerbi, Poljmer, 1990 31, 658.<br />

[I 121 E. Galbiati and G. Zerbi, J. Chem. Phys. 1986 84, 3500.<br />

[I131 E. Galbiati and G. Zerbi, J. C/iem. Plzys., 1987 87, 3653.<br />

[114] E. Galbiati, G. Zerbi, E. Benedetti and E. Chiellini, <strong>Polymer</strong>, 1991 3-3, 1555.<br />

[ 1151 S. J. Spells, S. J. Organ, A. Keller and G. Zerbi, <strong>Polymer</strong>. 1987 28, 697.<br />

[116] H. Kay and B. A. Newman, .4cta Crystollogr. 1968 B24, 615.


3.20 Refeeveiicrs 205<br />

[I 171 P. Jona, M. Gussoni and G. Zerbi. J. Appl. Phys., 1985 57, 834.<br />

[118] M. Gussoni, P Jona and G. Zerbi. J. Mol. Struct., Theochem, 1985 119, 329; ibid. p. 347.<br />

[119] T. Shimanouchi, J. Phys. Chem. Ref: Data, 1973 2, 225.<br />

[120] G. Zerbi, G. Gallino, N. Del Fanti and L. Baini, <strong>Polymer</strong>.. 1989 30, 2324.<br />

[I211 G. Avitnbile, R. Napolitano, R. Pinozzi, K. D. Rouse, W. Thomas and B. T. M. Willis,<br />

Polvn1. Lett.. 1975 13, 371.<br />

[I221 S. Abbate, M. Gussoni and G. Zerbi, J. Cheni. Phj~x, 1979 70, 3577.<br />

[I731 E. Agosti, G. Zerbi and I. M. Ward, <strong>Polymer</strong> 1992 33, 4219.<br />

[124] G. Zerbi, G. Conti, G. Minoni, S. Pison and A. Bigotto. J. Phys. Chenz. 1987 91, 2386.<br />

[125] M. Del Zoppo and G. Zerbi, <strong>Polymer</strong>, 1990 31, 658.<br />

[ 1261 D. Chapmann, Thc Structi/re of Lipids, Methuen, London, 1965.<br />

11271 C. Almirante, G. Minoni and G. Zerbi, J. Phys., Cheni., 1986 YO, 852.<br />

11281 L. Basini, A. Rafhelli and G. Zerbi, C11ein. Motel-. 1990 2, 679.<br />

[129] G. Ferrdri and G. Zerbi, J. Rrmiczn <strong>Spect</strong>roscopy, 1994 25, 713.<br />

[I301 For a discussion on the structure of PVC see, for instance, F. A. Bovey, F. P. Hood, E. W.<br />

Anderson and R. L. Kornegay, J. Phys. Clzem. 1967 71, 312; L. Cavalli, G. C. Borsini, G.<br />

Carraro and G. Confalonieri, J. Po/jm Sci., Part A-1. 1970 8, 801.<br />

11311 R. F. Schaufele and T. Shimanouchi, J. Chem. Phys. 1967 47, 3605.<br />

11321 T. Shimanouchi, Piwe and Appl. Chem., 1973 36, 93.<br />

11331 G. Minoni and G. Zerbi, J. Phys. Chem., 1982 86, 4791.<br />

11341 G. Minoni and G. Zerbi, J. Pdym. Sci. Polyni. Lett., 1984 22, 533.<br />

[I351 E. Fermi, Z. Physik, 1931 71, 250.<br />

[136] J. Lavalley and N. Sheppard, <strong>Spect</strong>rochim. Acta, Part A. 1972 28, 845.<br />

[137] S. Wunder, M. Bell and G. Zerbi, J. Chem. Phys., 1986 85, 3827.<br />

11381 R. G. Snyder, S. L. Hsu and S. Krimm. <strong>Spect</strong>roclzini. Am, Part A, 1978 31, 395.<br />

11391 R. G. Snyder and J. R. Scherer, J. Chem. Phys., 1979 71, 3271.<br />

[140] S. Abbate, G. Zerbi and S. L. Wunder, J. Pkjx Chenz., 1982 86, 3140.<br />

[I411 S. Abbate, S. L. Wunder and G. Zerbi, J. Phys. C/zeni.. 1984 88, 593.<br />

11421 L. Ricard, S. Abbate and G. Zerbi. J. Phys. Cheni., 1985 89, 4793.<br />

[143] G. Zerbi and S. Abbate, Chem. Phys. Lett., 1981 80, 455.<br />

[144] M. Del Zoppo and G. Zerbi, <strong>Polymer</strong>, 1991 32, 1581.<br />

11451 G. Zerbi, P. Roncone, G. Longhi and S. L. Wunder, J. Chem. Phys., 1988 89, 166.<br />

11461 R. G. Snyder, J. R. Scherer and P. Gaber. Biochinz. Biophjx Acta, 1980 601, 47.<br />

[147] W. G. Rotschild, Dyirtiniics of'Molecukrr Liqiiitls. Wiley, New York, 19S4.<br />

[148] B. G. Sumpter, D. W. Noid and B. Wunderlich, J. Chem. Phys. 1990 93, 6875; G. L. Liang,<br />

D. W. Noid, B. G. Sumpter and B. Wunderlich, Actu <strong>Polymer</strong>., 1993 44, 319.<br />

[149] M. Kobayashi, T. Kobayashi, Y. Cho and F. Kaneko, J. Chem. Phys. 1986 84,4636.<br />

[I501 1. W. Levin, in Adoances in IR arid Rantan <strong>Spect</strong>roscopy, Wiley, New York. 1984, Vol 11.<br />

[151] B. P. Gaber and W. Peticolas, Biochim. Biophys. Acta 1977 465, 260.<br />

11521 A. Muller, Proc. Royrl Soc., London Ser. 1930 A127, 417.<br />

11531 M. Broadhurst, J. Res. Nut/. Btrr. Stand Sect 1962 A66, 241.<br />

[ 1541 K. Larsson, Nature (London) 1967 213, 383.<br />

[155] G. Zerbi, R. Piazza and K. Holland-Moriytz, <strong>Polymer</strong>, 1982 23, 1921.<br />

[156] G. Ungar and A. Keller, Colloid Pol~vi. Sci., 1979 257, 90.<br />

[157] F. Guillaume, C. Sourisseau and A. J. Dianoux, J. Cheni. Phys., 1990 93. 3536.<br />

[158] P. Jona, B. Bassetti, V. Benza and G. Zerbi, J. Chenz. Phys., 1987 86, 1561.<br />

11591 M. L. Klein, J. Client Soc. Farada~ Trans., 1992 13, 1701.<br />

11601 L. Askadskaya and J. P. Rabe, Phys. Reo. Lett., 1992 69, 1395.<br />

[lhl] For review on model calculations see D. W. McClure, J. C/iern. Phys., 1968 49, 1530.<br />

[162] H. G. Olf and A. Peterlin, J. Poly. Sci., 1970 A2, 791.<br />

[ 1631 B. Ewen and D. Richter, J. Cheni. Pliys. 1978 69, 2954.<br />

[164] J. D. Barnes, J. Chern. Phys., 1973 58. 5193.<br />

[165] R. L. McCullough, J. Macromol. Sci.. Phys., 1974 9, 97.<br />

11661 M. Mansfield and R. H. Boyd, J. PO/~YJI. Sci.. Polyni. Pliys. Ed.., 1978 16, 1227.<br />

11671 M. L. Mansfield, CIIenz. Plijx Lett., 1980 69, 383.


206 3 Vibratiord <strong>Spect</strong>ra cis a Prohr of StrirctLiral Order<br />

[168] J. L. Skiiiiiei- and P. J. Wolynes, Chm. Phys, 1980 73, 4015.<br />

[169] G. Zerbi aiid M. Del Zoppo, J. Clzern. SOC. Fur.crday T~nr7s., 1992 88, 1835.<br />

[ 1701 H. Dwey-Aharon, T. S. Slucking, P. L. Taylor and A. J. Hopfinger, Phys. Rev. Sw. B, 1980<br />

21.<br />

11711 G. Borionetti, G. Zaiiiioni and G. Zerbi, J. Mol. Srvircf. 1990 12.1, 425<br />

[I721 S. Radice, N. Del Fanti and G. Zerbi, Po/ynirr, 1997 35, 2753.


4 Vibrational <strong>Spect</strong>roscopy of Intact and Doped<br />

Conjugated <strong>Polymer</strong>s and Their Models<br />

Y. Firrukawa and M. Tusumi<br />

4.1 Introduction<br />

Most organic polymers are electrical insulators, a property which distinguishes<br />

them from metals. However, the development of a new class of organic polymers<br />

with quasimetallic electrical conductivities has been actively pursued during the<br />

past 18 years, following the discovery in 1977 of high electrical conductivities<br />

for doped polyacetylenes [ 1, 21. The novel concept of a conducting organic polymer<br />

has aroused the interest of a large number of researchers in various areas such as<br />

chemistry, physics, electrical engineering, inaterial science, etc. Detailed discussions<br />

in each area can be found in many review articles [3-171. In particular, a new field of<br />

physics has been opened for the purpose of understanding their electrical properties.<br />

A goal of basic research on conducting polymers is to understand the metallic<br />

properties of such polymers. Since conducting organic polymers have conjugated<br />

7c-electrons in common, chemists intuitively believe that studies on the physical and<br />

chemical properties of conjugated systems would lead to an elucidation of the<br />

mechanism of charge transport. On the other hand, new concepts, solitoris [18, 191,<br />

yolui.oizs [20-221, and bipolaroils [21-231 have been proposed by solid-state physicists<br />

as elementary excitations in conjugated polymers, in order to explain the<br />

physical properties of these polymers. They are collectively called self-loculized<br />

e-xcitutions. These concepts and terminologies are unfamiliar to molecular spectroscopists.<br />

Thus, it seems desirable to build a bridge between solid-state physicists and<br />

molecular spectroscopists.<br />

The structures and properties of conducting polymers have been studied by<br />

various spectroscopic techniques. Ainong them, vibrational ( Rainaii and infrared)<br />

spectroscopy is a powerful tool for elucidating the molecular and electronic structures<br />

of conducting polymers as described in previous reviews [13-151. In spite of<br />

numerous spectroscopic studies, discussions on polarons, bipolarons, and solitons<br />

were not based upon reliable evidence from vibrational spectroscopy until very<br />

recently. Thus, the aims of this review are: (i) to provide a general introduction to<br />

the concepts of polarons, bipolarons, and solitons from the standpoint of molecular<br />

spectroscopists; (ii) to describe the methodology of Raman studies on these selflocalized<br />

excitations; (iii) to review the results of studies on poly( p-phenylene) and<br />

other polymers; and (iv) to discuss the mechanism of charge transport in conducting<br />

polymers.


208 4 T/ihmtional <strong>Spect</strong>roscopy of Ivitact arid Doped Coiijirgnted <strong>Polymer</strong>s<br />

4.2 Materials<br />

The conductivity of iodine-doped polyacetylene first reported by Shirakawa et al.<br />

[I] in 19'7'7 was 30 S cm-'. Since then, the conductivity reported for doped polyacetylene<br />

has kept increasing, the highest conductivity obtained so far for an iodinedoped<br />

stretched polyacetylene film [1?] being > 10' S cm-', a value comparable<br />

with that of copper (6 x 10' S cni-I).<br />

A film of intact polyacetylene usually shows a conductivity lower than lo-'<br />

S cm-' . However, the conductivity increases dramatically when the film is exposed<br />

to oxidizing agents (electron acceptors) such as iodine, AsF5, H'S04, etc. or reducing<br />

agents (electron donors) such as alkali metals. This process is referred to as duping,<br />

by analogy with the doping of inorganic semiconductors. The polymers that are not<br />

doped are referred to as ititact polymers in this review. The main process of doping<br />

is a redox reaction between the polymer chains and acceptors (or donors). Upon<br />

doping, an ionic complex consisting of positively (or negatively) charged polymer<br />

chains and counter ions such as 13-, AsF6-, etc. (or Naf, K+, etc.) is formed.<br />

Counter anions or cations are generated by reduction of acceptors or oxidation of<br />

donors, respectively. The use of an acceptor causes the p-type doping, and that of<br />

a donor the n-type doping. The electrical conductivity can be controlled by the<br />

content of a dopant. A sharp increase in conductivity is observed when the dopant<br />

content is < 1 mole% per C2H2 unit. After this sharp increase, the conductivity becomes<br />

gradually higher with further increase in the dopant content. At low doping<br />

levels, polyacetylene does not exhibit metallic properties, whereas its conductivity is<br />

high. When the dopant content is more than about 13 mole% per C2Hz unit, polyacetylene<br />

shows metallic properties such as a Pauli susceptibility [24, 251, a linear<br />

temperature dependence of the thermoelectric power [26], and a high reflectivity in<br />

the infrared region [27], though the temperature dependence of conductivity is not<br />

like that of a metal. The origin of such properties of heavily doped polyacetylene is<br />

not yet fully understood.<br />

Following polyacetylene, a large number of conducting polymers have been<br />

reported. The chemical structures of typical conducting polymers are depicted in<br />

Figure 4-1. These conducting polymers have conjugated n-electrons in common,<br />

(a) (b) (C) (d)<br />

pq..J-@--q)n<br />

s \ i n<br />

(e)<br />

(f)<br />

p, //<br />

aniline<br />

Figure 4-1. Chemical structures of conducting<br />

polymers. ia) Trurrs-polyacetylene;<br />

(b) cis-polyacetylene; (c) polyj p-<br />

phenylene); (d) polypyrrole; (e) polythiophene:<br />

if 1 poly( p-phenylenevinylene);<br />

(g) poly(2,5-thienylenevinylenei; (h) poly-<br />

(leucoemeraldine base form); (i j<br />

(h) (0 polyisothianaphthene.


4.3 Georrietvy oj Intuct Po1vniev.r 209<br />

and all are either insulators or semiconductors in their intact states, and they<br />

become conductors upon doping. Metallic properties are observed for poly( p-<br />

phenylene) [28], polythiophene 129-331, polypyrrole [3 11, and polyaniline [34, 351 at<br />

heavy doping levels, although reported data depend on samples and preparation<br />

methods. The origin of the metallic states of heavily doped polymers is one of the<br />

major unresolved problem in the field of conducting polymers.<br />

4.3 Geometry of Intact <strong>Polymer</strong>s<br />

The question of whether the C-C bonds in an infinite polyene chain are equal or<br />

alternate in length has been discussed by quantum chemists since the 1950s [36, 371.<br />

Various experimental results on oligoenes have shown the existence of alternating<br />

single and double bonds. The bond alternation in rrnns-polyacetylene has been<br />

confirmed by X-ray [38] and nutation NMR [39] studies. However, it is difficult to<br />

determine exact structure parameters for conducting polymers from X-ray diffraction<br />

studies, because single crystals of the polymers are unavailable. It is then useful<br />

to examine the geometries of the polymers and model oligomers by the molecular<br />

orbital (MO) method, especially at nb biitio Hartree-Fock levels or in higher<br />

approximations.<br />

Conducting polymers can be divided into two types, degenevnte and nondegenerate,<br />

according to the structure of the ground-states of intact polymers. The<br />

total energy curve of a degenerate system in the ground state is shown schematically<br />

as a function of a structural defoiination coordinate, R, in Figure 4-23, and that<br />

for a nondegenerate system in Figure 4-21 Let us consider an infinite transpolyacetylene<br />

chain as a prototype of the degenerate ground-state polymers. There<br />

are two degrees of freedom in bond lengths: rc-c and YC-C, which correspond, respectively,<br />

to the lengths of the alternating single and double CC bonds. The coordinate<br />

for structural deformation is simply expressed by the following equation.<br />

This coordinate reflects the degree of bond alternation. This coordinate is used in the<br />

effective-conjugation-coordinate model proposed by Zerbi et al. [15] for the purpose<br />

Figtire 4-2. Total energy as a function<br />

of a structural deformation coordinate.<br />

(a) Trans-polyacetylene<br />

><br />

u<br />

LT<br />

(degenerate ground-state system); ib) +<br />

poly( p-plienylenei (nondegenerate<br />

ground-state system).<br />

6<br />

\ A. /


210 4 Vihvatiotial <strong>Spect</strong>i~oscopy of hztuct ~ nd Doped Corijuyuted Po!vnrers<br />

of explaining the Raman spectra of intact polymers and doping-induced infrared<br />

bands. (This model will be mentioned again in Section 4.5.) There are two stable<br />

structures (A and B in Figure 4-2a) with alternating C-C and C=C bonds, while the<br />

structure with equal C-C bond lengths (C in Figure 4-2a) is unstable. These two<br />

structures are identical to each other and have the same total energy; in other<br />

words, they are degenerate. According to a nutation NMR study [39], the lengths of<br />

the C-C and C=C bonds in tv~ztzs-polyacetylene are 1.44 and 1.36 A, respectively.<br />

An ab initio MO calculation at the Hartree-Fock level with the 6-31G basis set for<br />

a model compound, C22H24, has shown that the C-C and C=C bond lengths of the<br />

central unit are 1.450 and 1.338 A? respectively [40]. Similar results have been<br />

obtained from calculations which take into account the effects of electron correlation<br />

[41]. From these experimental and theoretical results, Ro (equilibrium value of<br />

R, see Figure 4-2a) is calculated to be 0.06 and 0.079 A, respectively.<br />

A nondegenerate polymer has no two identical structures in the ground state. Let<br />

us consider poly( p-phenylene) as an example of nondegenerate polymers. In order<br />

to express the structural defoimation in this polymer, the following coordinate may<br />

be chosen, as proposed by Castiglioni et al. [42].<br />

where T,S are shown in Figure 4-2b. The total energy of this polymer has only one<br />

minimum at Ro as shown in Figure 4-2b. If we use the bond lengths, vl = v3 =<br />

v4 = 1’6 = 1.388 A, 12 = 1’5 = 1.382 A, and v7 = 1.492 A, of the central unit of neutral<br />

p-terphenyl obtained by an ab initio MO calculation at the Hartree-Fock level<br />

with the double-zeta quality 3-21G basis set [43], Ro is calculated to be 0.490 A. As<br />

shown in Figure 4-2b, an increase in R means the structural defoimation toward a<br />

quinoid structure from a benzenoid structure.<br />

4.4 Geometric Changes Induced by Doping<br />

The main process of doping is a charge-transfer reaction between an organic polymer<br />

and a dopant. When charges are removed from (or added to) a polymer chain<br />

upon chemical doping, geometric changes occur over several repeating units and<br />

the charge is localized over this region. Structure parameters such as bond lengths,<br />

bond angles, etc. are changed in this region. For example, the bond lengths [44] of<br />

neutral 1.7-terphenyl, its radical anion, and its dianion calculated by the PM3 method<br />

are shown in Table 4-1. From this table, we can see readily that the bond lengths<br />

of the neutral and charged species are different from each other. The phenyl and<br />

phenylene rings in neutral p-terphenyl have benzenoid structures. In the radical<br />

anion and dianion, an increase in R occurs as the result of shortening of the 1’23, v45,<br />

and v67 bonds and lengthening of the ~ 1 21-34, , and 1’56 bonds; in other words, geometric<br />

changes from benzenoid to quinoid occur upon ionization. The geometric


1'45<br />

4.4 Geometric Chunges I?zditced bv Doping 21 1<br />

Table 4-1. Bond lengths (in A) calculated by the PM3 method for p-terphenyl, its radical anion,<br />

and its dianion.<br />

Species Bond length" Coordinateb<br />

1'11 123 1'34 745 r56 y67 R<br />

Neutral 1.391 1.390 1.397 1.469 1.397 1.389 0.507<br />

Radical anion 1.397 1.385 1.417 1.426 1.423 1.366 0.580<br />

Dianion 1.400 1.371 1.438 1.388 1.446 1.351 0.640<br />

Numbering of carbon atoms is shown below. Dihedral angles between neighboring benzene rings<br />

are 48", O", and 0" for y-terphenyl, its radical anion, and its dianion, respectively. The data are<br />

taken from [44].<br />

The structural deformation coordinate defined in Eq. 4-2 is calculated by using the equation,<br />

R = 11 fi(4~56 ~ 2167 ~ ).<br />

changes from the neutral species are larger for the dianion than for the radical<br />

anion. Since the localized charges, which are accompanied by geometric changes,<br />

can move along the polymer chain, they are regarded as charge carriers in coiiducting<br />

polymers. These quasiparticles are classified into polarons, bipolarons, and<br />

solitons according to their charge and spin.<br />

4.4.1 Polarons, Bipolarons, and Solitons<br />

Self-localized excitations and corresponding chemical terminologies are listed in<br />

Table 4-2. Schematic structures of the self-localized excitations in poly( p-<br />

phenylene) and trans-polyacetylene are depicted in Figure 4-3. In these illustrations,<br />

the charge and spin are localized on one carbon atom. In real polymers, however,<br />

they are considered to be localized over several repeating units with geometric<br />

changes.<br />

When an electron is removed from an infinite polymer chain, charge +e and spin<br />

1/2 are localized over several repeating units with geometric changes. This is a selflocalized<br />

excitation called a positive polaron (Figure 4-3a, e). Since a positive<br />

polaron has charge +e and spin 1/2, it is a radical cation in chemical terminology.<br />

When an electron is added to a polymer chain, a negative polaron, which is a<br />

radical anion, is formed (Figure 4-3b, f).<br />

When another electron is removed from a positive polaron, charge +2e is localized<br />

over several repeating units. The charge +2e is considered to be localized in a<br />

region narrower than that of a positive polaron. This species is called a positive<br />

bipolaron (Figure 4-3c). Since a positive bipolaron has charge 1% and no spin, it<br />

is a dication in chemical terminology. A negative bipolaron corresponding to a<br />

dianion can also be considered (Figure 4-3d).


2 12 4 T4hvritionnl <strong>Spect</strong>voscopy of Intact nnd Doped Conjugrited Poljmiers<br />

Table 4-2. Self-localized excitations and chemical terminologies.<br />

Self-localized excitation Chemical term Charge Spin<br />

positive polaron<br />

negative polaron<br />

positive bipolaron<br />

negative bipolaron<br />

neutral soliton<br />

positive soliton<br />

negative soliton<br />

radical cation<br />

radical anion<br />

dication<br />

dianion<br />

neutral radical<br />

cation<br />

anion<br />

+e<br />

--E<br />

+ 2e<br />

- 2e<br />

0<br />

+e<br />

-e<br />

Figure 4-3. Schematic structures of self-localized excitations in poly( p-phenylenej and tucms-polyacetylene.<br />

(a) Positive polaron; (b) negative polaron; (c) positive bipolaron; id) negative bipolaron;<br />

(e) positive polaron; (fj negative polaron: (g) neutral soliton: (11) positive soliton; (i) negative soliton;.<br />

D, donor; A, acceptor; +, positive charge: -, negative charge; 0, unpaired electron.<br />

In trans-polyacetylene having a degenerate ground-state structure, soliton excitations<br />

can be formed between the A and B phases (see Figure 4-2a). Solitons are<br />

classified into neutral, positive, and negative types according to their charge. A<br />

neutral soliton has no charge and spin 1/2 (Figure 4-3g). A positive (or a negative)<br />

soliton has charge +e (or -e) and no spin (Figure 4-3h, i). A neutral soliton, a<br />

positive soliton, and a negative soliton correspond, respectively, to a neutral radical,<br />

a cation, and an anion of a linear tvans-oligoene with an odd number of carbon<br />

atoms, CnHn+2 (where n is an odd number).<br />

Bond-alternation defects or misfits in polyenes were studied using the Huckel MO<br />

method in the early 1960s [36, 45, 461, though these are nothing but solitons proposed<br />

by Su, SchrieHer, and Heeger [lS, 191 about two decades after. These workers<br />

have studied solitons in tvnns-polyacetylene by using a simple tight-binding Hamil-


4.4 Geometric Chmges Iiid~lrcecl bjj Dopiiiy 21 3<br />

Figure 4-4. Schematic molecular orbital energy levels. (a) Neutral conjugated system; (b) its radical<br />

cation; (c) its dication; (d) neutral radical of an odd oligoene. HOMO, highest occupied molecular<br />

orbital; LUMO, lowest unoccupied molecular orbital; SOMO, singly occupied molecular orbital;<br />

0. electron.<br />

tonian with an electron-lattice interaction teiin (SSH model), their model being<br />

essentially the same as the Huckel approximation. The continuuin versions of the<br />

SSH niodel are very useful for describing polarons and bipolarons as well as solitons,<br />

because analytic solutions can be obtained for physical parameters (creation<br />

energy, electronic absorption, etc.) relating to the excitations [21, 47-49].<br />

4.4.2 Electronic States of Poiarons, Bipoiarons, and Solitons<br />

We will first discuss the electronic states of linear conjugated molecules at the<br />

Huckel level with electron-lattice interaction [7], as models of self-localized excitations.<br />

The MO energy levels of these systems are schematically shown in Figure 4-4.<br />

In a neutral conjugated molecule (e.g., p-terphenyl), the bonding and antibonding<br />

levels are formed as the result of interaction between n-electron levels (Figure 4-4a).<br />

In its radical cation, geometric changes lead to an upward shift of the highest<br />

occupied molecular orbital (HOMO) and a downward shift of the lowest unoccupied<br />

molecular orbital (LUMO), and one electron is removed from the HOMO<br />

level (Figure 4-4b); the singly occupied molecular orbital (SOMO) is then formed.<br />

In its radical anion, the HOMO level is occupied by two electrons and the LUMO<br />

has one electron. Since geometric changes in the dication, where another electron<br />

is removed from the HOMO level (Figure 4-4c), are larger than those in the radical<br />

cation, the HOMO and LUMO levels shift further. In the dianion, each of the<br />

HOMO and LUMO levels is occupied by two electrons. In the MO energy levels of<br />

an odd trans-oligoene radical, CllHJ1+2' (where n is an odd number), a nonbonding<br />

level, which is occupied by one electron, is formed in the center of the band gap as<br />

shown in Figure 4-4d [36]. In its cation and anion, the nonbonding level IS occupied<br />

by null and two electrons, respectively.


CB<br />

I<br />

B;<br />

VB<br />

r-7<br />

Figure 4-5. Schematic electronic band<br />

structures. (a) Neutral polymer; (b) posi-<br />

IWS tive polaron; icj negative polaron; id)<br />

positive bipolaron; (e; negative bipolaron:<br />

!f! neutral soliton; (g) positive soliton; (11<br />

negative soliton. CB, conduction band:<br />

VB. valence band; 0. electron; arrow,<br />

(e) (f) (9) (h) electronic transition.<br />

Next, we discuss the electronic transitions due to polarons, bipolarons, and solitons<br />

in a polymer chain on the basis of a theoretical study reported by Fesser et al.<br />

[48] on a continuum electron-phonon-coupled model. The electronic energy levels<br />

of a neutral infinite polymer and those of polarons, bipolarons, and solitons are<br />

shown schematically in Figure 4-5.<br />

In an infinite neutral polymer, interaction between repeating units leads to the<br />

formation of electronic energy bands. The bonding and antibonding levels constitute,<br />

respectively, the valence band (VB) and the conduction band (CB) as shown in<br />

Figure 5-5a. The band gaps (WI in Figure 5-Sa) of most intact polymers are in the<br />

range between 35000 and 8000 cm-' (4.3 and 1.0 eV, 1 eV = 8066 cm-I).<br />

For a positive polaron (Figure 5-5b). two localized electronic levels, bonding and<br />

antibonding, are formed within the band gap. One electron is removed from the<br />

polaron bonding level. Thus, a positive polaron is expected to have the following<br />

three intragap transitions:<br />

01, polaron bonding level +- valence band<br />

02, polaron antibonding level + polaron bonding level<br />

(03, polaron antibonding level t valence band, and conduction band t polaron<br />

bonding level.<br />

When a negative polaron is formed, one electron is added to the polaron antibonding<br />

level. In this case also, three transitions are expected (Figure 4-5c).<br />

The electronic structure of a positive bipolaron is shown in Figure 4-5d. Since the<br />

geometric changes for a bipolaron are larger than those for a polaron, localized<br />

electronic levels appearing in the band gap for a bipolaron are farther away from<br />

the band edges than for a polaron. Two electrons are removed from the bipolaron


4 5 Mer1iodolog.v of Raiii~11 Studies of Polarons, B@olmons, and Solitons 215<br />

bonding level. Thus, a positive bipolaron is expected to have the following two<br />

iiitragap transitions:<br />

01 ’, bipolaron bonding level t valence band<br />

(oj’, bipolaron antibonding level - valence band<br />

When a negative bipolaron is formed, two electrons are added to the bipolaron<br />

antibonding level (Figure 4-5e) and thus two transitions are again expected. The<br />

intensities of the electronic transitions for polarons and bipolarons will be discussed<br />

in Section 4.7.1.2.<br />

When a soliton is formed, a nonbonding electronic level is formed at the center of<br />

the band gap (Figure 4-5f-11). For a neutral soliton, the nonbonding electronic level<br />

is occupied by one electron, while for a positive soliton and a negative soliton, the<br />

level is occupied by null and two electrons, respectively. Thus, all the neutral, positive,<br />

and negative solitons are expected to have only one intragap transition, us, as<br />

shown in Figure 4-5f-h.<br />

Polarons, bipolarons, and solitons are different from each other as described<br />

above. It is expected that we can detect these excitations separately by electronic<br />

absorption, ESR, and vibrational spectroscopies. In particular, geometric changes<br />

induced by doping can be studied by vibrational (Raman and infrared) spectroscopy.<br />

4.5 Methodology of Raman Studies of Polarons,<br />

Bipolarons, and Solitons<br />

Doping induces strong infrared absorptions that are attributed to the vibrational<br />

modes arising from charged domains generated by doping 110, 14, 15, 501. In the<br />

case of trans-polyacetylene, Horovitz et al. [5 1-53] have proposed the aniplitudemode<br />

model for explaining doping-induced infrared absorptions. According to<br />

their theory, the normal vibrations that induce oscillations in bond alternation are<br />

strongly observed in the infrared spectra of doped polyacetylene. Castiglioni et al.<br />

[ 541 have reformulated the amplitude-mode theory in tenns of the GF matrix<br />

method [55] used in molecular spectroscopy. In their treatment, they have proposed<br />

an effective conjugation coordinate (expressed by Eq. 4- 1) which reflects bond<br />

alternation. They have also explained the doping-induced infrared absorptions<br />

of other conducting polymers such as polypyrrole, polythiophene, poly( p-phenylenevinylene),<br />

etc., by using effective conjugation coordinates [50]. However, the<br />

types of self-localized excitations (polarons, bipolarons, and solitons) existing in<br />

these doped polymers have not been identified.<br />

Raman specti-oscopy of conducting polymers has mainly treated the structures of‘<br />

intact polymers (geometry of polymer chains, conjugation length, force field, etc.)<br />

[ 13-15]. Although the Ramaii spectra of doped polymers have been reported<br />

113-151, the observed Raman bands have not been correlated to the types of self-


2 16 4 Vilmitioizal <strong>Spect</strong>roscopy cf Intact nnd Doped CorzjLiyuted Poluviner=r<br />

localized excitations created by doping. Zerbi et al. [15] have drawn a conclusion<br />

from the effective-conjugation-coordinate model that the Raman bands arising<br />

from the charged domains are extremely weak and would never be observed.<br />

However, a new approach to Raiiian identification of the self-localized excitations<br />

will be described in this review.<br />

Most of intact polymers show electronic absorptions in the region from ultraviolet<br />

to visible. Upon doping, new absorptions with several peaks appear in the<br />

region from visible to infrared. These new absorptions are associated with selflocalized<br />

excitations created by doping. T~LIS, we can observe vibrational spectra<br />

arising from the charged domains, by using resonance Raman spectroscopy with<br />

a wide range of excitation wavelengths from visible to infrared [56, 571. So far,<br />

although most of the Raman spectra of doped polymers have been measured by<br />

using visible laser lines for excitation, it is also desirable to use laser lines with<br />

longer wavelengths.<br />

A useful method for analyzing the vibrational and electronic spectra of polymers<br />

is to compare them with the spectra of oligomers. In chemical terminology,<br />

polarons, bipolarons, and charged solitons are nothing but radical ions, divalent<br />

ions, and ions of oligomers, respectively (see Section 4.4.1). If these charged<br />

oligomers have geometries similar to those of self-localized excitations, these compounds<br />

would give rise to the electronic and Raman spectra similar to those of the<br />

self-localized excitations. We can identify self-localized excitations on the basis of<br />

the spectroscopic data for charged oligomers (charged oligomer approach) [56, 571.<br />

4.6 Near-Infrared (NIR) Raman <strong>Spect</strong>rometry<br />

Since the 1970s, Raman spectrometry has been using visible laser lines as the most<br />

common excitation sources. It was generally believed that NIR laser lines were not<br />

suitable for obtaining high-quality Raman spectra, though since Hirshfeld and<br />

Chase [58] and Fujiwara et al. [59] reported successful NIR Raman spectrometry<br />

measurements in 1986, this technique has advanced dramatically and NIR Raman<br />

spectrophotonieters are now commercially available.<br />

We will describe two NIR Raman spectrophotonieters, one with a diffraction<br />

grating, and the other with an interferometer. A Ti : sapphire laser (Coherent Radiation<br />

890) pumped with an argon ion laser (Coherent Radiation Innova 90-6) is used<br />

as an NIR source for Raman excitation. This laser covers the wavelength range<br />

between 680 and 1000 nm. Since the output intensity of the Ti: sapphire laser is very<br />

stable for a long time in comparison with a NIR dye laser, it is a good excitation<br />

source for NIR Raman spectrometry. Continuous-wave lasers are used for measuring<br />

the Raman spectra of doped polymers, because the doped samples are liable<br />

to damage due to high peak powers of pulsed lasers. Raman spectra are usually<br />

measured on a single polychromator (Spex 1870) equipped with a charge-coupleddevice<br />

(CCD) detector (Princeton Instruments LN/CCD-l024TKBS, back-illuminated<br />

type, 1024 x 1024 pixels) operated at -120 "C. The CCD detector functions


4.7 Poly(p-pheiq.lene) 217<br />

with an ST-135 controller (Princeton Instruments) linked with an NEC PC-9801FX<br />

computer through an IEEE-488 interface. The spectroscopic response of the CCD<br />

detector extends to about 1000 nin. Optical filters (holographic notch filters) are<br />

used to eliminate the Rayleigh scattering. The advantage of this system lies in its<br />

high optical throughput due to the use of the single polychromator and optical<br />

filter.<br />

We have modified an NIR Fourier-transform spectrophotometer (JEOL JIR-<br />

5500) for the nieasurements of Rarnan spectra excited with the 1064-nm line of a<br />

continuous-wave Nd : YAG laser (CVI YAG-MAX C-92) [60]. Rainan-scattered<br />

radiation is collected with a 90" off-axis parabolic mirror in a back-scattering<br />

geometry. Collected radiation is passed through two or three long-wavelength-pass<br />

filters to reduce the Rayleigh scattering.<br />

4.7 Poly( p-phenylene)<br />

As a typical example, we will first discuss the results of the electronic absorption<br />

and Raman spectra of Na-doped poly( p-phenylene). Poly( p-phenylene) consists of<br />

benzene rings linked at the para positions as shown in Figure 4-lc. Since poly(pphenylene)<br />

has a nondegenerate structure, polarons and/or bipolarom are expected<br />

to be formed by chemical doping. As model compounds of poly( p-phenylene),<br />

y-oligophenyls (abbreviated as PPn, 11 being the number of phenylene and phenyl<br />

rings) such as biphenyl (PP2), p-terphenyl (PP3). p-quaterphenyl (PP4), p-quinquephenyl<br />

(PP5), and p-sexiphenyl (PP6) are used. It has been shown by X-ray<br />

diffraction studies that the two phenyl rings of PP2 [61, 621 in crystal at room temperature<br />

are coplanar, and PP3 [63] and PP4 [64] in similar conditions deviate<br />

slightly from coplanarity. However, these deviations are not significant in analyzing<br />

their vibrational spectra. Accordingly, the phenylene and phenyl rings of PP5 and<br />

PP6 in solids, and poly( p-phenylene) are assumed to be coplanar.<br />

4.7.1 Electronic Absorption <strong>Spect</strong>ra<br />

4.7.1.1 Intact and Doped Poly(ppheny1ene)<br />

The electronic absorption spectra of undoped and electrochemically Bu4N+(ntype)-doped<br />

poly( y-phenylene) films are shown in Figure 4-6a, b, respectively [65,<br />

661. The undoped poly( p-phenylene) film shows the electronic absorption band at<br />

27400 cm-' (365 nm), which is associated with the interband transition from the<br />

valence band to the conduction band. The value of the band gap (0101 in Figure 4-521)<br />

is estimated to be 24200 cm-' from the onset of the electronic absorption [66]. The<br />

electronic transition energies of PPn [67, 681 are plotted in Figure 4-7. The transition<br />

energy decreases as the chain length becomes longer, an observation explained<br />

by the fact that increased conjugation occurs for longer chains. The observed tran-


30 20 10 0<br />

ENERGY / lo3 cm-l<br />

Figure 4-6. Absorption spectra<br />

of !a) undoped and (b) electrochemically<br />

Bu4N +(n-type)-<br />

doped poly~,y-phenylene) films.<br />

Arrows indicate the positions of<br />

excitation wavelengths (488.0.<br />

514.5, 632.8, 720. and 1064 nm)<br />

used for Ranian measurements<br />

[66J.<br />

40<br />

.<br />

0<br />

30<br />

><br />

u<br />

(r<br />

w<br />

z<br />

;<br />

0<br />

20<br />

u<br />

0<br />

10<br />

2 3 4 5<br />

NUMBER OF RINGS<br />

Figure 4-7. Observed electronic transition energies<br />

of p-oligophenyls. (a) Neutral species: (b) band I of<br />

the radical anions; (c) band 11 of the radical anions;<br />

(d) band I' of the dianions. The data of neutral<br />

p-oligophenyls are taken from 1671. The data of the<br />

radical anions and dianions are taken from [69, 701.<br />

sition energy of poly(p-phenylene), 27400 cn-', is lower than that of PP6, 31 500<br />

cni-' , indicating that the conjugation length of poly( p-phenylene) is much longer<br />

than 6. Upon Bu4N+ doping) two broad absorptions appear at about 5600 cm-'<br />

(1 800 nm) and 19 400 cm-' (5 15 nm). The assignments of these bands will be discussed<br />

in the next section on the basis of the spectra of the radical anions (PPn-)


4.7 Pol)~(p-pheiz~~lene) 219<br />

ENERGY i 1 0 ~ ~ ~ - '<br />

ENERGY I oZcm<br />

Figure 4-8. Absorption spectra ofa THF solution ofp-terphenyl at (a) early and ib) later stages of<br />

reduction. In (a), the 35600 cni-' band decreases and the 20800 and 10900 cm-' bands increase in<br />

intensity. In (b), the 20800 and 10900 cm-' bands decrease and the 15400 cm-' band increases in<br />

intensity.<br />

and dianions (PPn-) of p-oligophenyls, which are models of negative polarons and<br />

bipolarons, respectively.<br />

4.7.1.2 Radical Anions and Dianions of p-Oligophenyls<br />

The electronic absorption spectra of PPn- and PPH- have been systeinatically<br />

studied [69, 701. When a tetrahydrofuran (THF) solution of PP3 is exposed to a<br />

sodium mirror, its absorption spectrum gives new bands (Figure 4-8). At an early<br />

stage, the band observed at 35700 cm-' (280 nm) arising from neutral PP3<br />

decreases and the bands observed at 20800 and 10900 cm-' (481 and 916 nm)<br />

increase in intensity. These new bands are attributed to the radical anion of PP3<br />

(PP3'-). In the course of this spectral change, an isosbestic point is observed at<br />

about 31 500 cm-', indicating that PP3 is quantitatively converted into PP3'-.<br />

When the solution is further exposed to the sodium mirror, the 20 800 and 10 900<br />

cm-' bands decrease and the band observed at 15400 cm-' (650 nm) increases in<br />

intensity. The 1.5400 cm-' band is ascribed to the dianion of PP3 IPP3'-). During<br />

this reaction, three isosbestic points are observed at about 25 800, 20400, and<br />

13400 cm-'. The THF solutions of PPiz (n = 4,.5, and 6) show similar reduction<br />

reactions which may be described as<br />

PPrz + Na 4 PPi7'- + Na'<br />

PPi7'- + Na i PPn2- + Na'<br />

Since PPn- is stable for a long time, we can conclude that the following disproportionation<br />

reaction does not occur.<br />

2PPn'- 4 PPiZ + PPd


220 4 Vibvofionnl Slwctvoscopy oj Intact and Doped Conjugated Po1ynier.s<br />

band I1<br />

-<br />

‘6 b?, @6 d hg<br />

4 -fs- 4 *-<br />

$4 ‘4<br />

tJ3 tJ3<br />

(a)<br />

(b)<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

Figure 4-9. Schematic energy level diagrams. (a) The<br />

radical anion of biphenyl IPP2.-); (b) the dianion of<br />

biphenyl (PP2’-). 0, electron; arrow. electronic transition.<br />

The inolecular orbital levels are taken from 1771.<br />

This reaction is similar to the process of the intramolecular formation of a bipolaron<br />

from two polarons.<br />

The electronic transition energies [69, 701 of the radical anions and dianions of<br />

p-oligophenyls are plotted in Figure 4-7. Two strong bands are observed for the<br />

radical anions and are called bands I and I1 (curves b and c in Figure 4-7, respectively).<br />

On the other hand, one intense band is observed for the dianions, which<br />

is called band I’ (curve d in Figure 4-7). In the electronic absorption spectra of<br />

the radical cations and dications of a-oligothiophenes [71-73], two strong bands<br />

and one strong band are commonly observed for the radical cations and dications,<br />

respectively. In the case of the radical cations of oligophenylenevinylenes, two<br />

bands are observed [74, 751, although different results are also reported [76].<br />

The absorption spectra of conducting polymers including poly( p-phenylene) have<br />

been discussed within the frame work of the Hiickel approxiination [7, 101. In order<br />

to correlate the spectra of the radical anions and dianions of y-oligophenyls with<br />

those of doped poly( p-phenylene), we will discuss the observed absorption spectra<br />

of the charged p-oligophenyls on a similar theoretical basis. The energy diagram of<br />

PP2’- is shown schematically in Figure 4-9a. Molecular orbital levels in this figure<br />

are taken from the results calculated by the Pariser-Parr-Pople-SCF-MO method<br />

for PP2 (021, symmetry) [77]. A Pariser-Parr-Pople-SCF-MO-CI calculation for<br />

PP2‘- [78] shows that band I is assigned to the transition mostly attributable to<br />

d6. These transitions are<br />

the electronic excitation of dl0 t d7 and band I1 to d7 +-<br />

polarized along the long molecular axis. The a, and d9 levels are nonbonding with<br />

respect to the inter-ring CC bond. It is reasonable to consider that these assignments<br />

hold also for PPn- (12 = 3-6). Moreover, band I’ of PP2’- is assigned to the transition<br />

mostly attributable to the electronic excitation bl,,


is worthwhile pointing out that the cu3 and (03' transitions are symmetry forbidden.<br />

According to a continuum electroii-phonon-coupled model reported by Fesser et<br />

al. [48], the (01 and 032 transitions are dominant among the three expected for a<br />

polaron, and the 031' transition is more intense between the two expected for a<br />

bipolaron. Finally, it is expected that a polaron hus two intense intrugup tramitions<br />

and a bipolavoir one intense tmnsition [79].<br />

The absorption spectrum of Bu4Nf-doped poly( p-phenylene) can now be<br />

explained as follows. Two broad bands centered at about 5600 and 19400 or1<br />

(Figure 4-6) are attributable to the 031 and cc)? transitions of negative yolarons,<br />

respectively. These assignments will be discussed again in Section 4.7.3 on the basis<br />

of the Raman results.<br />

4.7.2 Raman <strong>Spect</strong>ra<br />

4.7.2.1 Intact Poly(p-phenylene) and p-Oligophenyls<br />

The observed infrared and Raman spectra of intact poly( p-phenylene) [80-841 have<br />

been analyzed by normal coordinate calculations [43, 84-87]. The factor group of a<br />

coplanar polymer is isomorphous with the point group &. When the JZ plane is<br />

taken in the plienylene-ring plane (s axis along the polymer chain) and the x axis<br />

perpendicular to the phenylene-ring plane, the irreducible representation at the zone<br />

center (k = 0) is as follows:<br />

rvlb<br />

rvlb<br />

in-pldne = 5ug(R) + 4b1u(IR) + 4hu(IR) + 5hg(R)<br />

out-of-plane = 2 ~u + lbl,(R) + 3hg(R) + 2hu(IR)<br />

where R and IR denote Raman- and infrared-active vibrations, respectively. The<br />

1064-nm excited Raman spectrum and infrared absorption spectrum of intact<br />

poly( p-phenylene) prepared by dehalogenation polycoiidensation of y-dihalogenobenzene<br />

are shown in Figure 4-10a, b, respectively. The observed and calculated<br />

vibrational frequencies of poly( p-phenylene), "C-substituted analog, and perdeuterated<br />

analog are listed in Table 4-3, except for CH (CD) stretches [q(ug), i'g(hlu),<br />

v16(bzu), and v20(03~)]. Major Ranian and infrared bands have been assigned.<br />

In Figure 4-10b, the 809 cm-' infrared band is assigned to the CH out-of-plane<br />

bend of the phenylene rings, and the 765 and 694 cm-' bands are assigned to the<br />

CH out-of-plane bends of the terniinal phenyl rings.<br />

The Raman spectrum of poly( p-phenylene) has been compared with those of<br />

p-oligophenyls. The 1064-nm excited Rainan spectra of p-oligophenyls in the solid<br />

state [88] are shown in Figure 4-11. The bands at 1605-1593, 1278-1275, 1221-<br />

1220, and 795-774 cmrl correspond to those at 1595, 1282, 1222, and 798 ,nip' of<br />

intact poly( y-phenylene), respectively. These bands are assigned to the ug vibrations<br />

(v2-vg). The atomic displacements of these modes obtained by normal coordinate<br />

calculations based on the PM3 method [84] are depicted in Figure 4-12a-d. The<br />

1595 cm-' mode (Figure 4-12a) is contributed mainly by the CC stretch of the<br />

phenylene ring. The 1282 cm-' mode (Figure 4-12b) is mainly contributed by the


m<br />

0<br />

m<br />

I<br />

1500 1000 5<br />

WAVENUMBER icrn~l<br />

Figure 4-10. Vibrational spectra of intact poly(yphenylene).<br />

(a) Raman spectrum taken with the<br />

1064-nni line in powder sample; (b) infrared spectrum<br />

in a KBr disk [83].<br />

inter-ring CC stretch. The 1222 cm-' mode (Figure 4-12c) is a mixture of CC<br />

stretch and CH in-plane bend. The 798 cm-' mode (Figure 4-12d) is a mixture of<br />

CC stretch and CCC deformation. The intensity of the 1278-1275 cn-' band relative<br />

to that of the 1221-1220 c1n-l band decreases as the chain length becomes<br />

longer. Thus, this ratio is a marker of the length of conjugated segments [81, 82, 881.<br />

The wavenumbers of the Raman bands of poIy( p-phenylene) and y-oligophenyls<br />

are almost independent on the excitation wavelength [Sl]. It is well known that the<br />

wavenumbers of the two strong Rainan bands of tmm-polyacetyleiie depend greatly<br />

on the wavelength of the excitation laser. These large dispersions are explained by<br />

the existence of segments of various conjugation lengths that give rise to the Raman<br />

bands at different wavenumbers [13, 141. However, the Raniail spectra of other<br />

conducting polymers do not show such dependence on the excitation wavelength.<br />

The (resonance) Raman intensities of conjugated molecules have been analyzed<br />

by the classical vibronic theory due to Tang and Albrecht [89]. Inagaki et al. 1901<br />

have shown that resonance Ranian intensities of totally symmetric modes of p-<br />

carotene (an oligoene with eleven conjugated C=C bonds) arise from the A term<br />

(Franck-Condon term) of the Albrecht theory. Probably, the Franck-Condon<br />

factor plays an important role in determining the resonance Rainan intensities of<br />

almost all conjugated molecules. The resonance Ranian intensity of each totally<br />

symmetric mode is proportional to the square of the shift of the potential minimum<br />

in the resonant excited electronic state from that in the ground electronic state along<br />

the normal mode giving rise to the resonance Raman band, when the shift is small<br />

[91-941. The effective conjugation coordinate represents approximately the change<br />

of equilibrium geometry between the ground electronic state and the first dipoleallowed<br />

excited electronic state [15]. Thus, the resonance Raman intensity of a<br />

normal mode is determined by the contribution of the effective conjugation coordinate<br />

to the mode. Details are described in [15]. It has been demonstrated that the


~<br />

~<br />

~<br />

~<br />

~<br />

~<br />

~<br />

~<br />

~<br />

~<br />

~<br />

801<br />

Table 4-3. Vibrational frequencies (in cm-') of neutral polyi p-phenylenei and its "C-substituted<br />

and perdeuterated analogs.<br />

Sym- No. Normal species '3C-substituted Perdeuterated analog<br />

metric<br />

analog<br />

species<br />

Observed Cnlculated Observed Calculated Observed Calculated<br />

1831 [84] [43] [86] [87] [83] [841 ~ 4 1 1841 1871<br />

1595<br />

1282<br />

1222<br />

798<br />

-<br />

1482<br />

1048<br />

1000<br />

-<br />

1401<br />

I254<br />

1081<br />

-<br />

624<br />

809<br />

~<br />

~<br />

~<br />

~<br />

1599 1626 1661 1601 1541<br />

1295 1282 1289 1290 1239<br />

1226 1233 1127 1184 1209<br />

776 792 846 802 768<br />

969 951 961 ~<br />

427 367 401 - -<br />

831 823 834<br />

1477 1487 1510 1490 1452<br />

1060 1045 1051 1044 -<br />

991 984 968 1004 965<br />

945 968 945 ~<br />

768 754 760 -<br />

441 410 402<br />

1407 1399 1440 1412 1361<br />

1232 1275 1268 1343 ~<br />

1119 1162 1075 1118 ~<br />

1603 1601 1654 1652 -<br />

1290 1325 1328 1343<br />

633 625 602 616 602<br />

487 488 460 418 ~<br />

198 797 790 ~<br />

497 457 458 - -<br />

1547 1563<br />

1246 1255<br />

1216 892<br />

741 764<br />

960 ~<br />

415 ~<br />

825 ~<br />

1445 1354<br />

1035 817<br />

958 978<br />

936 ~<br />

742 ~<br />

426 ~<br />

1375 1338<br />

1 I93 -<br />

1092 ~<br />

1546 ~<br />

1277 -<br />

610 606<br />

470 -<br />

789 650<br />

48 1 ~<br />

1563 1572<br />

1281 1253<br />

881 863<br />

751 776<br />

775 ~<br />

378 -<br />

647 ~<br />

1353 1364<br />

812 811<br />

984 977<br />

824 ~<br />

663 -<br />

414 ~<br />

1334 1347<br />

1185 1268<br />

854 855<br />

1579 1620<br />

1016 1044<br />

600 593<br />

450 383<br />

661 ~<br />

435 -<br />

Rainan spectrum of neutral poly( p-phenylene) is well explained by the effectiveconjugation-coordinate<br />

model [95].<br />

4.7.2.2 Doped Poly(p-phenylene) and the Radical Anions and Dianions<br />

of p-Oligophenyls<br />

The Raman spectra of heavily Na-doped poly( y-phenylene) measured with the<br />

488.0, 514.5, 632.8, 720, and 1064 nm laser lines are shown in Figure 4-13 [70]. The<br />

Raman spectra of Na-doped poly( p-phenylene) are apparently different from that<br />

of intact poly( p-phenylene), indicating that these bands arise from charged domains<br />

(i.e., negative polarons and/or negative bipolarons) created by Na-doping. The<br />

assignments of these bands will be discussed on the basis of the data for the radical<br />

anions and dianions of p-oligophenyls.


224 4 Vibrcitioiznl <strong>Spect</strong>roscopy of Intact and Doped Conjugated <strong>Polymer</strong>s<br />

RAMAN SHIFT/cm-' Excitation wavelength is 1064 nin [88]<br />

Figure 4-12. Atomic displacements of inajor vibrational modes<br />

(k = 0) of poly(p-phenylene). (a) 1599 an-' [v2(ng)]; (b) 1295<br />

cm-' [q(ug)];(c) 1226 cm-I [114(a,)]: (d) 776 cm-I [v5(az)]:<br />

(d) (el (f) (e) 1477 c1n-l [1~,0(b,~)]; (f) 991 c1n-l [iy(hlu)][84]<br />

The Raman spectra of the radical anions of PP2 [92, 96, 971, PP3 [70, 981, PP4<br />

[70, 981, PP5 [70] and PP6 [70], and the diaiiions of PP3 [70, 981, PP4 [70, 981, PP5<br />

[70], and PP6 [70] have been reported by several authors. The Raman spectra of the<br />

radical anions (PPnP) and dianions (PPTz~~) in THF solutions are shown in Figures<br />

4-14 and 4-15, respectively. The excitation wavelengths used are rigorously or<br />

nearly resonant with band I1 of the radical anions and band I' of the dianions. The<br />

Raman spectra of the radical anions are similar to each other, and so are the<br />

Rainan spectra of the dianions. However, the two groups of the Raman spectra are<br />

different from each other and from those of neutral 12-oligopheiiyls. These results<br />

suggest that the negative polarons and bipolarons, corresponding respectively to the<br />

radical anions and dianions, can be identified by Raman spectroscopy.<br />

The Ranian spectra of the radical anions in Figure 4-14 show siiiall changes with


4.7 Poly(y-phenylene) 225<br />

Figure 4-13. Raman spectra of Na-doped poly( -<br />

p-phenylene) taken with various wavelengths of excitation<br />

lasers. (a) 488.0; (b) 514.5; (c) 632.8: (d) 720, (e)<br />

1064 nm [70].<br />

1500 1000<br />

RAMAN SHIFT / Cm-'<br />

Figure 4-14. Resonance Ranian spectra of the radical anions<br />

of p-oligophenyls in THF solutions. (a) p-Terphenyl; (b)<br />

p-quaterphenyl; (c) p-quinquephenyl; (d) p-sexiphenyl. Excitation<br />

wavelengths are 458.0 and 514.5 nm for (a) and (b-d),<br />

respectively. The bands of the solvent are subtracted [70].<br />

t<br />

t<br />

(I)<br />

2<br />

W<br />

b<br />

z<br />

2<br />

<<br />

2<br />

4<br />

LT<br />

1500 1000<br />

RAMAN SHlFT/cm-'<br />

the chain length. The intensities of the bands observed in the 1032-991 and 780-<br />

753 cm-' regions decrease as the chain length becomes longer, indicating that these<br />

bands are associated with the terminal rings. The bands observed in the 780-753<br />

cni-' region show an upshift in wavenumber, as the chain length becomes longer.<br />

This result indicates that the wavenumbers of these bands can be used as a measure<br />

of the localization length of polarons. The bands observed in the 1325-1305 c1n-l<br />

region are assigned to the inter-ring CC stretch [70, 92, 981. The corresponding<br />

bands in neutral PPIZ are observed in the range between 1278 and 1275 cm-'. About<br />

40-cm-' upshifts are explained by the increased TC-bond orders of the inter-ring CC<br />

bonds, i.e., shortening of the inter-ring CC bond lengths. In PP?, the calculated


226 4 k'ilmtioiiul <strong>Spect</strong>roscopy of Iiitact aid Doped Conjirgtrted Polyinem<br />

1500 7 000<br />

RAMAN SHIFTicm'<br />

Figure 4-15. Resonance Raman spectra of the dianions of<br />

p-oligophenyls in THF solutions. (a) p-Terphenyl; h) y-<br />

quaterphenyl; (c) p-quinquephenyl; (d) psexiphenyl. Excitation<br />

wavelengths are (a) 632.8, (b) 720, (cj 1064, and (d) 1064 nni,<br />

respectively, The bands with x are due to the solvent. Fluorescence<br />

background is subtracted in all the spectra 1701.<br />

length of the inter-ring CC bond (rq5') shrinks from 1.469 to 1.426 A in going from<br />

neutral PP3 to PP3'- (see Table 4-1).<br />

In the Raman spectra of the dianions in Figure 4-15, the bands in the 1479-1468<br />

cm-' and 1188-1 162 cm-' regions are strongly observed. Corresponding bands<br />

are weakly observed or are not observed in the spectra of the radical anions. This<br />

reflects structural differences between the electronic ground state and the excited<br />

state associated with band I1 of the radical anions and band I' of the dianions. The<br />

bands observed in the 1357-1320 cm-l region are assigned to the inter-ring CC<br />

stretch [70, 981. About 70 cm-l upshifts in going from neutral PPn to PPiz'- are also<br />

explained by the shortening of the inter-ring CC bond lengths. These upshifts of<br />

PPn- are larger than those of PPn'-. This observation suggests that the degrees of<br />

shrinkage of the inter-ring CC bonds in PPn'- are larger than those in PPn-. This<br />

is consistent with the results of calculation for PP3 given in Table 4-1 : 1.469 A (PP3)<br />

and 1.388 A (PP3'-). It is worth pointing out that the frequency of inter-ring CC<br />

stretch increases with the value of structural deformation coordinate R (column 8<br />

in Table 4-1) for PP3, PP3'-, and PP32-. In the 1000-940 cm-' region, there is a<br />

series of bands at 943 (PP3'-), 951 (PP4'-), 974 (PP5'-), and 992 cm-' (PP6'-).<br />

This band series may be useful in estimating the localization length of bipolarons.<br />

Next, we will analyze the Raman spectra of Na-doped poly( p-phenylene) (Figure<br />

4-13) on the basis of those of the radical anions and dianions ofp-oligophenyls. The<br />

Raman spectra of Na-doped poly( p-phenylene) taken with the 488.0- and 5 14.5-nm<br />

lines (Figure 4-13a, b) are different froin those with the other three laser lines<br />

(Figure 4-13c-e), and are similar to those of the radical anions of y-oligophenyls<br />

(Figure 4-14). This observation can be understood by considering that the Raman<br />

bands arising from charged domains formed by Na-doping in the polymer chains<br />

are quite similar to those of the radical anions. Since the radical anions are viewed<br />

as negative polarons in the polymer, negative polarons exist in Na-doped poly(pphenylene)<br />

[70]. These Raman spectra of Na-doped poly( p-phenylene) are observed<br />

under resonant conditions. Thus, the results described above indicate that the


Table 4-4. Observed Raman frequencies (in cn- ' ) of negative polarons and negative bipolarons in<br />

Na-doped poly( p-phenylene).<br />

No. Negative polaron" Negative bipolaron" Assignmentb<br />

A 1591-1590 1594-1586 syni. in-phase mode cousistirig of vz(~i0)<br />

B 1477-1371 sym.. VlO(h,)<br />

c 1315-1 314 1352- 1344 sym., in-phase, 1'; ((ig)<br />

D 1297-1190 Sylll., OUt-of-phdse, IQ[ag)<br />

E 1221-1219 1222-1219 sym., in-phase, 1'4(Ug)<br />

F 1179 1190-1 I79 sym., out-of-phase, 1'4 (ag)<br />

G 980-974 992-951 sym., 1'12(blu)<br />

H 783-777 syni., in-phase, 1'5 ((1,)<br />

Data are taken from [83].<br />

Numbering and symmetry correspond to those of neutral poly( pphenylene). Sym. denotes a<br />

symmetric mode.<br />

absorption around 488.0 and 514.5 nm (20500 and 19400 cm-') is attributed to<br />

polarons.<br />

The Raman spectra excited with laser lines in the 632.8-1064 iim region (Figure<br />

4-15-e) are quite similar to those of the dianions of p-oligophenyls (Figure 4-1 5),<br />

which correspond to negative bipolarons in the polymer. Accordingly, negative<br />

bipolarons also exist in Na-doped poly( y-phenylene), and the absorption in the<br />

632.8-1064 nm (15 800-9400 cni-I) region arises from bipolarons [70]. In the 1000-<br />

950 cm-' region of the Raman spectra, a few bands are observed at 979 and 957<br />

c11i-l (632.8-11111 excitation), at 977 and 952 Gin-I (720 nm), and at 992, 974, and<br />

956 cm-' (1064 nm). By comparing these bands with those of the Raman spectra of<br />

the dianions, it may be concluded that negative bipolarons localized over four, five,<br />

and six phenylene rings coexist in Na-doped poly( p-phenylene) .<br />

The Ranian bands arising from negative polarons and bipolarons, together with<br />

the tentative mode assignments based on the wavenumber shifts upon I3C-substitution<br />

[83], are listed in Table 4-4. The observed bands are called bands A-H as<br />

shown in Table 4-4. Bands A, C, E, and H correspond to the ag modes (vz, i = 2-5)<br />

of neutral poly( 13-phenylene), respectively. The 1282 cm-l Raman band of neutral<br />

poly(y-phenylene) upshifts to 13 15-13 14 cn-l in the spectra of polarons and 1352-<br />

1344 cm-' in the spectra of bipolarons. According to normal coordinate calculations<br />

[43, 84, 86, 871, the 1282-cmP' mode is mainly attributed to the inter-ring<br />

CC stretch. The observed upshifts upon Na-doping reflect the increasing z-bond<br />

order of the inter-ring CC bond, i.e., structural changes froin benzenoid to quinoid<br />

[70]. The upshifts for bipolarons (70-62 cm-') are larger than those for polarons<br />

(33-32 cm-l), indicating that the geometric changes in bipolarons are larger than<br />

those in polarons. On the other hand, bands A and E show little shifts in wavenumber<br />

and band H shows a moderate shift.<br />

In the Raman spectra of negative bipolaroils (Figure 4-13c-e), bands B, D, F,


228 4 Ii’Dratioizal <strong>Spect</strong>roscopy of Intact and Doped Coiijuyatecl Polytners<br />

and G are strongly observed. Band B is assignable to a symmetric mode consisting<br />

of the v,~(bl,,) mode (Figure 4-12e) of each benzene ring in the charged domain.<br />

Band G is also attributed to a symmetric mode consisting of the i112(61~,) mode<br />

(Figure 4-12f). Bands D and F are assignable to out-of-phase symmetric inodes<br />

consisting of the I I ~ ( L ~ and ~ ) i ~ ~ ( modes ~ i ~ )(Figure 4-12b, c) of each benzene ring,<br />

respectively. The appearance of bands B, D, F, and G indicates the loss of translational<br />

symmetry of neutral polyi p-phenylene) chain, wliich is consistent with the<br />

forniatioii of localized charged doniains upon doping. Bands B, D, F, and G are<br />

not observed or weakly observed for negative polarons, but strongly observed for<br />

negative bipolarons. These observations seem to reflect different geometric changes<br />

occurring between polarons and bipolarons 011 going from their respective ground<br />

electronic states to their respective excited electronic states.<br />

4.7.3 Assignments of Electronic Absorption <strong>Spect</strong>ra of Doped<br />

Poly( p-phen ylene)<br />

In Section 4.7.1, two broad bands centered at about 5600 and 19400 cm-l in the<br />

electronic absorption spectrum of Bu4N+-doped poly( p-phenylene) have been<br />

attributed to the QI and 02 transitions of negative polarons, respectively, on the<br />

basis of the data of the radical anions and dianions of p-oligophenyls. Raman<br />

spectroscopy provides information on the electronic levels associated with polarons<br />

and bipolarons as well as on their geometries. The Ranian results indicate the coexistence<br />

of polarons and bipolarons in Na-doped poly( p-phenylene). The Ranian<br />

bands due to polarons are resoiiantly enhanced by the use of the 488.0 and 514.5<br />

11111 laser lines, which are located within the electronic absorption around 19 400<br />

cin-l (51 5 nin). These observations confirm the proposed assignment that the<br />

19400 c1n-l band is due to the cuz transition of polarons. The Ranian bands due to<br />

bipolarons are resonantly enhanced by the use of the 632.8, 720, and 1064 nm laser<br />

lines. However, in the absorption spectruin of Bu4N+-doped poly( p-phenylene), no<br />

peaks are observed in the 15800-9400 cm-’ (632.8-1064 nm) region. This fact can<br />

be explained by considering that the content of the bipolaron is very small and the<br />

QI’ absorption due to the bipolaron is overlapped by the strong to1 absorption band<br />

due to polarons.<br />

4.8 Other <strong>Polymer</strong>s<br />

The Raman spectra of doped states of polyaniline [99], poly( p-phenylenevinylene)<br />

[75, 100-1021, and polythiophene [72, 731 have been also analyzed on the basis of<br />

the data of charged oligoiners (charged oligomer approach). The detected selflocalized<br />

excitations are suniniarized in Table 4-5. Some important conclusions can<br />

be drawn from these Rainaii studies.


4.8 Other Polwieys 229<br />

Table 4-5. Types of self-localized excitations detected by Raman spectroscopy.<br />

<strong>Polymer</strong> Dopant Species Reference<br />

Type Reagent Content<br />

poly( p-phenylene) 11<br />

poly ( p-phenylenevmylene) n<br />

polythiophene<br />

polyaniline (emeraldine base<br />

form, 2A)”<br />

polyaniline (leucoemeraldine, 1A)<br />

11<br />

n<br />

P<br />

P<br />

P<br />

P<br />

P<br />

p<br />

heavy<br />

heavy<br />

light<br />

heavy<br />

heavy<br />

heavy<br />

light<br />

heavy<br />

heavy<br />

heavy<br />

polaron, bipolaron<br />

polaron, bipolaron<br />

polaron, bipolaron<br />

polaron, bipolaron<br />

polaron<br />

polaron<br />

polaron<br />

polaron<br />

polaron<br />

polaron<br />

a HC1-doped emeraldine base form is called emeraldine acid form or 2s form<br />

1. Only polarons are detected for p-type doping. The Ranian spectra of the polymers<br />

doped with acceptors do not show large changes with various excitation<br />

wavelengths [72, 73, 751. The polymer-chain structures of p-type-doped polymers<br />

are more homogeneous, and more regular arrays of polarons are formed.<br />

2. Polarons and bipolarons coexist in n-type doped polymers in contrast to the<br />

results of p-type doping. The Raman spectra of the polymers doped with donors<br />

depend on the excitation wavelength [70, 1001. These observations are explained<br />

by the existence of various localization lengths of polarons and/or bipolarons<br />

[70, 1001. The polymer-chain structures of the n-type-doped polymers are inhomogeneous.<br />

Possibly, Na-doped polymers are easily hydrogenated, even in the<br />

presence of a small amount of water during the sample preparation, and negative<br />

polarons and/or bipolarons are formed on the short conjugated segments.<br />

The Raman spectra of doped polyacetylene have been reported with the excitation<br />

of visible laser lines [103-1101. There have been some discussions about the<br />

question of whether these bands arise from charged domains or undoped parts<br />

remaining in the doped films. Recently, the Ranian spectra of maximally Na-doped<br />

polyacetylene excited with NIR lasers have been reported [57, 60, 1111 and conipared<br />

with those of the radical anions of u, w-diphenyloligoenes and a, co-dithienyloligoenes<br />

[ 11 11. It has been concluded that the observed bands arise from charged<br />

domains (charged solitons and/or polarons), because the observed Raman spectra<br />

of the doped polymer are similar to those of the radical anions of x, oAiphenyloligoenes.<br />

However, further studies are required for characterizing doped polyacetylene<br />

in more detail.


230 4 Vilwntional Slwctvoscopjl of Intact mitl Doped Conjzrgutecl Poljwievs<br />

4.9 Electronic Absorption and ESR <strong>Spect</strong>roscopies and<br />

Theory<br />

It is widely accepted [7, 101 that the major species generated by chemical doping it?<br />

nondegenerate polymers is bipolarons, except for polyaniline. However, the result^<br />

obtained from Raman spectroscopy indicate the existence of polarons at heavily<br />

doped polymers. These results are inconsistent with the established view. The<br />

experimental basis for claiming the existence of bipolarons has been electronic<br />

absorption and ESR spectra [7, 9, lo]. We will comment on these experimental<br />

results .<br />

The electronic absorption spectra of p-type-doped polypyrrole have been interpreted<br />

in terms of polarons and bipolarons for the first time 11121. Electronic<br />

absorption spectra of doped polypyrrole in the range from visible to NIR show<br />

three bands due to intragap transitions at low dopant contents and two bands at<br />

high dopant contents [ 1131. The three bands are attributed to polarons and the two<br />

bands to bipolarons [112], because a polaron is expected to have three intragap<br />

transitions and a bipolaron is expected to have two intragap transitions (see Section<br />

4.4.2). A quantitative ESR study [114] has confirmed the interpretation of the electronic<br />

absorption experiments. This rule of thumb for band assignment has so fiiialso<br />

been applied to other conducting polymers.<br />

However, the two electronic absorption bands of doped poly( p-phenylene) are<br />

not explained by bipolarons, as demonstrated in Section 4.7.3. Furthermore,<br />

p-type-doped polythiophene shows two absorptions at about 12000 and 5200 cnir ’<br />

[72, 73, 115, 1161 and p-type-doped poly( p-phenylenevinylenej shows two broad<br />

absorptions at about 19000 and 7300 cmP1 [75, 117, 1181. In these two systems,<br />

only positive polarons are detected by Raman spectroscopy (see Table 4-5). These<br />

results of Rainan spectroscopy suggest that the two-band pattern in the electronic<br />

absorption spectrum is attributed to polarons. As discussed in Section 4.7.1.2, it is<br />

expected that a polaron has two intense intragap transitions and a bipolaron one<br />

intense transition. From these findings, we propose a new rule: a two-hmd pattern<br />

correspondy to yolavons mid one-barid puttern to hipolarons [79], when these species<br />

do not coexist. When polaroils and bipolarons coexist, and their electronic absorptions<br />

are overlapped, a more careful examination of the observed spectrum is of<br />

course necessary.<br />

Recently, Hill et al. [119] have proposed the existence of a singlet radical-cation<br />

dimer (i.e., polaron dimer) as an alternative to a spinless bipolaron to explain weak<br />

ESR signals from doped polymers. Their proposal is based on the observed dimerization<br />

of the radical cation of 2,5”-dimethylterthiophene. They have found by<br />

electronic absorption and ESR measurements that the above radical cation is<br />

dinierized at low temperature even at a low concentration (4 x lor4 mol/l). Singlet<br />

intrachain polaron pail-s and interchain polaron pairs give no ESR signals. Thus,<br />

the absence of ESR signals or observation of weak ESR signals does not directly<br />

lead to the conclusion that the doped polymer contains only bipolarons.<br />

The stability of polarons, bipolarons, and solitons has been studied theoretically


under various models. According to the SSH model and siinilar approximations<br />

with electron-lattice coupling for a single polymer chain [7, lo], the creation energy<br />

of one bipolaron in a nondegenerate system (or two charged solitons in a degenerate<br />

system) is lower than that of two polarons. Thus, it is considered that two separate<br />

polarons are always unstable, and changed into a bipolaron (or two charged solitons).<br />

For doped polyacetylene, the effect of the electron--electron interaction and<br />

the dopant potential, which are not taken into account in the SSH model, have been<br />

studied 11201. On the other hand, the existence of polarons has been theoretically<br />

proposed by several authors. Kivelson and Heeger [121] have proposed a polaron<br />

lattice (regular infinite array of polarons) for explaining the metallic properties of<br />

heavily doped polyacetylene. Mizes and Conwell [ 1221 have reported that polarons<br />

are stabilized in short conjugated segments for trmzs-polyacetylene and poly( p-<br />

phenylenevinylene) from the results obtained by a three-dimensional tight-binding<br />

calculation. Shimoi and Abe [123] have studied the stability of polarons and bipolarons<br />

in nondegenerate systems by using the Pariser-Parr-Pople method combined<br />

with the SSH model. A polaron lattice is stabilized by the electroii-electron<br />

interaction at dopant contents lower than a critical concentration, whereas a bipolaron<br />

lattice is stabilized at dopant contents higher than the critical concentration.<br />

Thus, importance of the electron-electron interaction and three-dimensional interaction<br />

as well as the electron-lattice coupling has been demonstrated. However,<br />

the nature of the metallic states of heavily doped polymers remains an unresolved<br />

theoretical problem.<br />

4.10 Mechanism of Charge Transport<br />

Overall electrical conduction in doped polymers is considered to have contributions<br />

from various processes such as intrachain, interchain, interdomain, interfibril<br />

transport, etc. The contributions of individual processes would depend on the solidstate<br />

structure and morphology [8]. Charge transport has been discussed in terms of<br />

polarons, bipolarons, charged solitons, and their lattices (6, 8, 101. Even when the<br />

dopant content is very small, electrical conductivity increases dramatically. At this<br />

doping level, discrete localized electronic levels associated with self-localized excitations<br />

are formed within the band gap, and the temperature dependencies of d.c.<br />

conductivity have been explained by the hopping between localized electronic states<br />

181. It is conceivable that the energy gap between the lower (or upper) localized state<br />

and the valence (or conduction) band for a positive (or a negative) polaron is<br />

smaller than that of a positive (or a negative) bipolaron. Then, it is expected that<br />

polarons play a more important role in hopping conduction than bipolarons.<br />

Heavily doped polymers are of special interest, because these polymers show<br />

metallic properties such as a Pauli susceptibility, a linear temperature dependence of<br />

the thermoelectric power, a high reflectivity in the infrared region, etc., as described<br />

in Section 4.2. The temperature dependence of d.c. conductivity in heavily doped<br />

polymers is not like a metal, whereas metallic properties are observed. These results


Figure 4-16. Schematic structures of polythiophene chains doped with electron acceptors (dopant<br />

content, 25 mole'% per thiophene ring) and bonding electronic levels of positive polarons and<br />

bipolarons. (a1 Polaron lattice: (bi bipolaron lattice. A. acceptor: +. positive charge; -, negative<br />

charge; 0, electron; -, electronic energy level.<br />

Ferrni level .-..<br />

-.... Ferrni level<br />

Figure 4-17. Schematic band structures of<br />

dn infinite polymer chain doped with<br />

acceptors (dopant content, 25 mole%) per<br />

thiophene ring). (a) Polaron lattice;<br />

(b) bipolaron lattice Shadowed areas are<br />

(a) (b) filled with electrons [ 1261.<br />

can be explained as follows. Although a single chain itself behaves like a metal,<br />

macroscopic electrical conduction is limited by interchain, interdomain, or interfibril<br />

charge transport. Here, we will discuss the intrachain metallic charge transport.<br />

Let us suppose an infinite nondegenerate polymer chain (e.g., polythiophene)<br />

doped heavily with electron acceptors. At a high dopant content, the polymer-chain<br />

structure and electronic structure of the doped polymer are radically different from<br />

those of the intact polymer. As typical cases, we will describe two kinds of lattice<br />

structures of doped polythiophene (dopant content, 25 mole'%, per thiophene ring):<br />

a polaron lattice and a bipolaron lattice. They are the regular infinite arrays of<br />

polarons and bipolarons. The schematic polymer-chain structures are shown in<br />

Figure 4-1 6. Band-structure calculations have been performed for polaron and/or<br />

bipolaron lattices of poly( p-phenylene) [ 1241, polypyrrole [ 1241, polyaniline [ 1251,<br />

polythiophene [ 124, 1261, and poly( p-phenylenevinylene) 11271, with the valenceeffective<br />

Hamiltonian pseudopotential method on the basis of geometries obtained<br />

by MO methods. The schematic electronic band structures shown in Figure 4-17


4. I0 Mechnriism oj Charge. Transport 233<br />

are based on the results calculated with this method for a polaron and a bipolaron<br />

lattice of p-type-doped polythiopheiie (dopant content, 25 inole%, per thiophene<br />

ring) [ 1261.<br />

A polaron has a localized electronic level occupied by one electron (i.e., SOMO).<br />

In a polaron lattice formed at a heavy doping level, electronic wave functions of<br />

neighboring polarons are overlapped and interact with each other. As a result, a<br />

half-filled metallic band is formed from the polaron bonding levels occupied by one<br />

electron (Figure 4-17a). The Fernii level is located at the center of the polaron<br />

bonding band. The electronic states of the polaron lattice are not localized, whereas<br />

the electronic states of a polaron are localized. The calculated bandwidths of the<br />

polaron bands are 8900, 7900, and 5600 cm-l (1.1. 0.98, and 0.69 eV) for heavily<br />

doped polyaniline [ 1351, polythiophene [ 1261, and poly( p-phenylenevinylene) [ 1271,<br />

respectively. If the conjugatioii length of the polymer chain is not sufficiently long,<br />

the metallic band is not formed. In this case, discrete levels are formed and hopping<br />

conduction is expected to take place. In contrast to the polaron lattice, no metallic<br />

band is formed for the bipolaron lattice, even if the conjugation length is long<br />

enough (Figure 4-17b). This is because a bipolaron has no singly occupied electronic<br />

level, whereas a polaron has a singly occupied level. In addition, the differences<br />

in the electronic band structure between the polaron and bipolaron lattices<br />

suggest that the bipolaron lattice is the result of a dinierization of the polaron lattice<br />

like the Peierls transition [126].<br />

In the case of rrnns-polyacetylene (a degenerate polymer), band-structure calculations<br />

[126] indicate that a polaron lattice has a metallic band whereas a charged<br />

soliton lattice has no metallic band. An alternating polaron-charged soliton lattice<br />

[ 1281 has been proposed for explaining the metallic properties of heavily doped<br />

states.<br />

We will discuss polymer-chain structures obtained by Raman spectroscopy and<br />

metallic properties. The Raman measurements indicate the existence of positive<br />

polaroiis for p-type-doped polythiophene, poly( p-phenylenevinylene), and polyaniline<br />

at heavy doping levels (Table 4-5). These results indicate that positive polarons<br />

are formed on the polymer chains upon acceptor doping. If conjugation length<br />

is sufficiently long and the dopant content sufficiently high, a large number of<br />

polarons are formed and interact with each other. Then, a polaron lattice is formed.<br />

Therefore, the results of p-type-doped polythiophene, poly( p-phenylenevinylene),<br />

and polyaniline are consistent with the formation of polaron lattices at heavy<br />

doping levels. Tlie rnetullic properties obs'eriled fbr cloped polpers probubly origincite<br />

from the poluroii lattice. From the results obtained so far, we can point out some<br />

important conditions for obtaining metallic polymers: (i) a long conjugation length;<br />

(ii) a high dopant content (large degree of charge transfer); and (iii) formation of<br />

polarons. In the case of n-type doping, both negative polarons and bipolarons are<br />

detected for doped poly( y-phenylene) and poly( p-phenylenevinylene) (Table 4-5).<br />

Polarons and bipolarons with various localization widths are detected, and electronic<br />

states are probably localized, even at heavy doping levels. It is considered<br />

that such an inhomogeneous structure does not lead to formation of a metallic<br />

band.<br />

It is well known that physical properties of conducting polymers depend on


234 4 I i’brationul <strong>Spect</strong>roscopy of Intact and Doped Conjugared Polynzers<br />

synthetic conditions. doping conditions, and other treatments of samples. Thus, in<br />

order to confirm tlie proposal that the metallic properties originate from the<br />

polaron lattice, it is requisite to use the same polymer film or the films from the<br />

same batch for measuring various physical properties and Raman spectra.<br />

4.11<br />

It has been demonstrated that Raman spectroscopy is a powerful tool for studying<br />

self-localized excitations such as polarons and bipolarons in doped conducting<br />

polymers. Resonance Rainan spectroscopy by using exciting laser lines in a wide<br />

range between visible and near-infrared gives valuable information on self-localized<br />

excitations; especially, near-infrared Raman spectroscopy developed recently is iniportant.<br />

<strong>Spect</strong>roscopic data on the radical ions and divalent ions of oligomers are<br />

useful for analyzing the electronic and vibrational spectra of doped polymers. New<br />

assignments of the electronic absorption spectra of polarons and bipolarons are<br />

proposed: a two-band pattern is attributed to polarons, and a one-band pattern<br />

to bipolarons. Polarons and/or bipolarons have been detected for doped poly( y-<br />

phenylene), polythiophene, poly( p-phenylenevinylene), and polyaniline by Raman<br />

spectroscopy. On the basis of tlie Rainan results, metallic properties at heavily<br />

doped polymers is attributed to formation of a polaron lattice with a half-filled<br />

metallic band.<br />

Postscript: After the manuscript of this Chapter was completed, a review book<br />

treating the polaron/bipolaron in conjugated polymers was published [ 1291.<br />

4.12 References<br />

[l] Shirakawa, H., Louis, E. J., MacDiarmid. A. G., Chiang, C. K., Hccger, A. J., J. Cheni. Soc.,<br />

Clieni. Coi72iiitiii. 1977, 578-580.<br />

[2] Chiang, C. K., Finchcr, Jr., C. R., Park, Y. W., Heeger, A. J.. Shirakawa, H.. Louis, E. J.,<br />

Gau, S. C., MacDiarmid, A.G., Phys. Reti. Lett. 1977, 39, 1098-1101.<br />

[3] Wegner, G., Angeiv. Clzetii. ht. Ed Engl. 1981, 20, 361-381.<br />

[4] Etcmad, S., Heegcr, A. J., MacDiarmid. A. G., Ann. Rev. Phys. Clzeni. 1982, 33, 443-469.<br />

[5] Baughman, R. H., Bredas, J. L.. Chance, R. R., Elsenbaumer, R. L., Shacklette, L. W.,<br />

Clzein. Rev. 1982, 82, 209-272.<br />

[6] Chien, J. C. W., Pol!~acetyletie-Clzerriistry, Physics, mzd hlatevial Science. New York:<br />

Academic Press, 1984.<br />

171 Brkdas, J. L., Street. G. B. Acc. Clzcrii. Rex 1985, 18, 309-315.<br />

[8] Roth, S., Blcier, H., Ah. Phys. 1987, 36, 385-462.<br />

[9] Patil, A. O., Hecger, A. J., Wudl, F., Cllen7. Rev. 1988, 88, 183-200.<br />

[ 101 Hecger, A. J., Kivclson, S., Schricffer, J. R., Su, W.-P., Reu. Mod Php. 1988, 60, 781L850.<br />

[ll] Kuzmaay, H., Mchring, M., Roth, S. (Eds.) Springer Ser. Solill-State Sci. 1987, 76; 1990, 91 ;<br />

1992, 107.<br />

[I 21 Lu, Y., Soktons arid Pokzr~or?~. Singapore: World Scientific, 1988.


4.12 Reftrences 235<br />

[ 131 Kuzmany, H.. Mokromol. Chern., iMacronzo/. svriip. 1990, 37, 8 1-97.<br />

[14] Harada, I., Furukawa; Y., in: l’ibrcitiorzul <strong>Spect</strong>iti atid Structure: Durig, J. R. (Ed.). Amsterdam:<br />

Elsevier, 1991; Vol. 19, pp. 369-469.<br />

[I 5) Gussoni, M., Castiglioni, C., Zerbi, G., in: Athwices in <strong>Spect</strong>ro.mp,v: Clark, R. J. H., Hester,<br />

R. E. (Eds.). Chichester: John Wiley & Sons, 1991; Vol. 19, pp. 251-353.<br />

[ 161 Bredas, J. L., Silbey. R. (Eds.), Conjirgured <strong>Polymer</strong>s. Netherlands: Kluwer Academic Publishers,<br />

1991.<br />

[I71 Tsukamoto, J., Ad;. Ph~,.s. 1992, 41, 509-546.<br />

1181 Su. W. P., Schrieffer, J. R., Heeger. A. J., Phys. Rev. Lett. 1979. 43, 1698-1701.<br />

(191 Su, W. P.. Schrieffer, J. R., Heeger, A. J., Phys. Rev. B 1980, 22, 2099-2111.<br />

[20] Su, W. P., Schrieffer, J. R., Proc Ntrtl. Actid Sci. 1ISA 1980, 77, 5626-5629.<br />

[211 Brazovskii, S. A., Kirova, N. N., Sozi. P/i]’s. JETP Lett. 1981, 33, 4-8.<br />

[22J Bishop, A. R., Campbell, D. K., Fesser, K., Mo/. Cr~ist. Liq. Crysl. 1981, 77, 253-264.<br />

[23] Bredas, J. L., Chance, R. R., Silbey, R., iMol. Crjvt. Liq. Cryst. 1981, 77, 319-332.<br />

[24] Ikehata, S., Kaufer, J., Woerner, T., Pron, A,, Druy, M. A., Sivak, A., Heeger, A. J., Mac-<br />

Diarmid, A. G., P/ij.s. Rev. Lett. 1980, 45, 1123-1126.<br />

[25] Chung, T.-C., Moraes, F., Flood, J. D., Heeger, A. J., Pliys. Rev. B 1984, 29, 2341-2343.<br />

1261 Park, Y. W., Denenstein, A,, Chiang, C. K., Heeger, A. J., MacDiarinid, A. G., Solid Sttrte<br />

CO/~IIIIUTI. 1979, 29, 747-75 1.<br />

[27] Tanaka, M., Watanabe, A,, Tanaka, J., Bid. Ckm. Soc. 4117. 1980, 53, 3430-3435.<br />

[28] Kume, K., Mizuno, K., Mizoguchi, K., Nomura, K., Maniwa, Y., Moo(. Cryst. Liq. Cryst.<br />

1982, 83, 285-290.<br />

[2Y] Moraes, F., Davidov, D., Kobayashi, M., Chung, T. C., Chen, J., Heeger, A. J., Wudl, F.,<br />

Synth. Met. 1985, 10. 169-179.<br />

[30] Kaneto, K., Hayashi, S., Ura, S., Yoshino, K., J. Pl7ys. So(,. Jpn. 1985, 54, 1146-1153.<br />

[31] Mizoguclii, K., Misoo, K., Kume, K., Kaneto, K., Shiraishi, T., Yoshino, K., Sjwfh. Met.<br />

1987, 18, 195-198.<br />

[32J Logdlund. M., Lazzaroni, R., Stafstrom, S., Salaneck, W. R., Bredas, J. L., Phys. Rev. Lett.<br />

1989, 17, 1841-1844.<br />

1331 Masubuchi, S., Kazania, S., Swzth. Met., 1995, 74, 151-158.<br />

[34] Ginder, J. M., Richter, A. F., MacDiarmid, A. G., Epstein, A. J., Solid Store Conimziri. 1987.<br />

63, 97-101.<br />

[35] Mizoguchi, K., Obana, T., Ueno, S., and Kume, K., Syiith. Met. 1993, 55, 601-606.<br />

[36] Salem, L., The Molecultu Orbitol Tlzcorp of” Co@pted Si1sterir.s. New York: Benjamin, 1966.<br />

[37] Kohler, B. E., in: Coi?jtigated Po1yiiier.s: Bredas, J. L. and Silbey, R. (Eds.). Dordrecht:<br />

Kluwer Academic Publishers, 1991; pp. 405-434.<br />

[38] Fmcher, Jr., C. R., Chen, C.-E., Heeger, A. J., MacDiarniid, A. G., Hasting, J. B., Plzys. Rev.<br />

Lett. 1982, 48, 100-104.<br />

[39] Yannoni, C. S., Clarke, T. C., Phys. Rev. Lett. 1983, 51, 1191-1193.<br />

[40] Villar, H. O., Dupuis, M., Watts, J. D., Hurst, G. J. B., Clementi. E., J. Clzenz. Phys. 1988,<br />

88, 1003-1009.<br />

[41] Hirata, S., Torii, H., Tasumi, M., J. Cl7enz. Pliys. 1995, 103, 8964-8979.<br />

1421 Castiglioni, C., Gussoni. M., Zerbi, G., Syntli. Met. 1989, 29, ElLE6.<br />

[43] Cuff, L., Kertesz, M., Mocroii7olecuke.s 1994, 27, 762-770.<br />

[44] Ohtsuka, H., Master’s Thesis, The University of Tokyo, 1993.<br />

[45] Longuet-Higgins, H. C., Salem, L., Proc. Roj~. Soc. A 1959, 251. 172-185.<br />

[46] Pople, J. A,, Walmsley. S. H., Mo/. Pliys. 1962, 5, 15-20.<br />

[47] Takayama, H., Lin-Liu, Y. R., Maki. K., Plzys. Reo. B 1980, 31, 2388-2393.<br />

1481 Fesser, K,, Bishop, A. R., Campbell, D. K.: Pkys. Reu. B 1983, 27, 4804-4825.<br />

1491 Onodera, Y., Pliys. Rex:. B 1984, 30, 775-785.<br />

[50] Zerbi, G., Gussoni, M., Castiglioni, C., in: Copjrgated Pdyrners: Bredas, J. L. and Silbey, R.<br />

(Eds.). Dordrecht: Kluwer Academic Publishers, 1991; pp. 435-507.<br />

[S 11 Horovitz, B., Solid Stntc. Contmi//7. 1982, 41. 729-734.<br />

1521 Vardeny, Z., Ehrenfreund, E., Brafman, 0.. Horovitz, B., PliI,s. Rea. Lert. 1983.51,23176-2329.<br />

[53] Ehrenfreund, E., Vardeny, Z., Brafman, O., Horovitz, B., Pliys. Rev. B 1987, 36, 1535-<br />

1553.


236 4 I’ihratioiznl <strong>Spect</strong>roscopy of Intact nizd Doped Conjirpted Polj~iiwrs<br />

[54] Castiglioni, C., Navarrete, J. T. L., Zerbi. G., Gussoni, M., Solid St~ite Coninirrn. 1988, 65.<br />

625-630.<br />

[55] Wilson, E. B., Decius, J. C., Cross, P. C., Molecular Vibra/ions. New York: MacGraw Hill.<br />

1955.<br />

[56] Furukawa, Y., Springer Ser. Solid-Stute Sci. 1992, 107, 137-143.<br />

[57] Furukawa, Y., Sakanioto, A., Ohta, H.. Tasumi, M., Svnrh. Ale/. 1992, 49, 335-340.<br />

[58] Hirschfeld, T., Chase, B., Appl, <strong>Spect</strong>rosc. 1986, 40, 133-137.<br />

(591 Fujiwara. M., Hamaguchi, H., Tasumi, M., Appl. <strong>Spect</strong>rosc. 1986, 40, 137-139.<br />

1601 Furukawa, Y., Ohta, H.. Sakamoto, A,, Tasumi, M., <strong>Spect</strong>rochirn. Actrr 1991,474 1367-1373.<br />

[hl] Trotter, J.. ilctri Cryst. 1961, 14. 1135-1340.<br />

[62] Hargreaves, A,, Rizvi, S. H., Act0 Cryst. 1962, 15, 365-373.<br />

[63] Baudour, J. L., Cailleau, H., Yelon, W. B., Acta Cryst. 1977, B33, 1773-1780.<br />

[64] Delugeard, Y., Desuche, J., Baudour, J. L., Actci Cribst. 1976, B32, 702-705.<br />

[65] Tabata, M., Satoh, M., Kaneto, K., Yoshino, K., J. Phj?,~. Soc. Jpn. 1986, 55, 1305-1310.<br />

[66] Satoh, M., Tabata, M., Uesugi, F., Kaneto, K., Yoshino, K., Sj’nth. Met. 1987. 17, 595-600.<br />

[67] Gillam, A. E., Hey, D. H., J. Chern. Soc. 1939, 1170-1177.<br />

1681 Dale, J., rlctu Clien?. Scnnd. 1957, 11, 650-659.<br />

[69] Balk, P., Hoijtink, G. J., Schreurs, J. W. H., Rec. Trrrv. Chiin. 1957. 76, 813-823.<br />

[70] Furukawa. Y., Ohtsuka, H., Tasumi, M., SJnth. Met. 1993, 55, 516-523.<br />

[71] Fichou, D., Horowitz, G., Xu, B., Garnier, F., Synth. Met. 1990, 39, 243-259.<br />

[72] Furukawa, Y., Yokonuma, N., Tasumi, M., Kuroda. M., Nakayama, J., Mol. C~J’.S/. Lfq.<br />

Crysf. 1994, 256, 113-120.<br />

[73] Yokonuma, N., Furukawa, Y., Tasumi, M., Kuroda, M., Nakayania, J., C/ie/n. Phys. Lett.<br />

Chein. Pliys. Lett. 1996, 255, 431-436.<br />

[74] Deussen, M., Bassler, H., Chenz. Phjx 1992, 164, 247-257.<br />

[75] Sakamoto, A,, Furukawa, Y., Tasunii, M., J. Phys. Chem. 1994, 98, 4635-4640; J. P/ij,x<br />

Chenz. 1997, 101, 1726-1732; Furukawa, Y., Sakamoto, A,, Tasunii, M. Mucrornol. Symp<br />

1996, 101, 95-102.<br />

[76] Spangler, C. W., Hall, T. J., Synth. Met. 1991, 44, 85-93.<br />

[77] Dewar, M. J. S.,TrinajstiC, N., Czech. Cheni. Cowviitm. 1970, 35. 3136-3189.<br />

[78] Zahradnik, R., Carsky, P., J. Phys. Chen7. 1970, 74, 1240-1248.<br />

[79] Furukawa, Y., SJint/7. Mct., 1995, 69, 629-632; J. Plrys. Chenz. 1997, 100, 15644-15653.<br />

[SO] Shacklette, L. W., Chance, R. R., Ivory, D. M., Miller, G. G., Bauglman, R. H., Spth. Met.<br />

1979. 1, 307-320.<br />

[81] Krichene, S., Lefrant, S., Froyer, G., Maurice, F., Pelous, Y., J. Plzys Colloq. 1983, 44, 733-<br />

736.<br />

[82] Krichene, S., Buisson, J. P., Lefrant, S., Slvth. Met. 1987, 17, 589-594.<br />

[83] Furukawa, Y., Ohtsuka, H., Tasumi. M., Wataru, I., Kanbara, T, Yaniamoto, T., J. Runicin<br />

<strong>Spect</strong>rosc. 1993, 24, 551-554.<br />

[84] Ohtsuka, H., Furukawa, Y., Tasumi, M., Wataru, I., Yamanioto, T., unpublished work.<br />

[85] Rakovic, D., Bozovic, I., Stepanyan, S. A,, Gribov, L. A,, Solid Stare Commun. 1982, 43,<br />

127-129.<br />

[86) Zannoni, G., Zerbi, G., J. Chern. Phys. 1985, 82, 31-38.<br />

1871 Buisson, J. P., Krichene, S., Lefrant, S., Synrh. Met. 1987, 21, 229-234.<br />

[88] Ohtsuka, H., Furukawa, Y., Tasumi, M., <strong>Spect</strong>rochinq. Act0 1993, 49A, 731-737.<br />

[89] Tang, J., Albrecht, A. C., in: Rnniaiz <strong>Spect</strong>roscopy: Szynianski, H. [Ed.). New York: Plenum,<br />

1970; Vol. 2, pp. 33-68.<br />

[90] Inagaki, F., Tasumi, M., Miyazawa, T., J. Mol. <strong>Spect</strong>rosc. 1974, 50, 286-303.<br />

[91] Warshel, A., Dauber, P., J. C/ierii. Ph?~s. 1977, 66, 547775488,<br />

1921 Yaniaguchi, S.? Yoshimizu. N., Maeda, S., J. Phys. Clzeni. 1978, 82, 1078-1080.<br />

1931 Peticolas, W. L., Blazej, D. C., Chein. Plzys. Le/t. 1979, 43, 604-608.<br />

[94] Kakitani, H., Clienz. Phys. Lett. 1979, 64, 344-347.<br />

[95] Cuff, L., Kertestz, M., J. Phys. Chern. 1994, 98, 12223-12231.<br />

[96] Takahashi, C., Maeda, S., Clieni. Phys. Lett. 1974, 24, 584-588.<br />

[97] Aleksandrov, I. V., Bobovich, Ya. S., Maslov, V. G., Sidorov, A. N., Opt. <strong>Spect</strong>rosc. 1975.<br />

38, 387-389.


4.12 References 237<br />

[98j Matsunuma, S., Ydmaguchi, S., Hirose, C., Maeda, S., J. Pliys. Chem. 1988, 92, 1777-1780.<br />

[99] Furukawa. Y., Ueda, F., Hyodo, Y., Harada, I., Nakajima. T., Kawagoe, T., Macro-<br />

7nolecule.s 1988, 21, 1297- 1305.<br />

[loo] Sakamoto, A,, Furukawa, Y., Tasumi, M., J. P/zJ~s. C/7eni. 1992, 96, 3870-3874.<br />

[lo1 J Sakamoto, A., Furukawa, Y., Tasunii, M., SJW//Z. Mct. 1993, 55, 593-598.<br />

[I021 Sakamoto. A,, Furukawa, Y., Tasumi, M., Noguchi, T., Ohnishi, T., Syrrtl7. Met. 1995. 69,<br />

439-440.<br />

[I031 Harada, I.. Furukawa, Y.. Tasunii, M., Shirakawa, H., Ikeda, S., J. Clzeni. Phys. 1980. 73,<br />

1746-4757.<br />

[I041 Kuzmany, H., Plzys. Stat. Sol. B 1980, 97, 521-531.<br />

[lo51 Furukawa. Y., Harada, I., Tasumi. M., Shirakawa, H., Ikeda, S., Clzetri. Lett. 1981, 1489-<br />

1492.<br />

[lo61 Schiigerl, B., Kuzniany, H., Phys. Stat. Sol. B 1982, 1//, 607-617.<br />

[I071 Faulques, E., Lefrant, S., J. P/i>i.r. Colloq. 1983, 44, 337-340.<br />

[lo81 Eckhardt. H., Shacklette, L. W., Szobota, J. S., Bauglunan, R. H., Mol. Cryst. Liq. Cryst.<br />

1985,1/7,401-409.<br />

[lo91 Tanaka, J., Saito, Y., Shimizu, M., Tanaka, C., Tanaka, M., Bzrll. Chem. Soc. Jpn. 1987, 60,<br />

1595-1605.<br />

[110] Mulazzi, E., Lefrant, S., Swtlr. Met. 1989, 28, D323-D329.<br />

11111 Furukawa, Y., Uchida. Y., Tasumi, M., Spangler, C. W., Mol. Cryst. Liq. Cryst. 1994, 556,<br />

721-726; Kim, J.-Y., Ando, S., Sakamoto, A., Furukawa, Y., Tasumi, M., Synth. Met. 1997,<br />

89, 149-152.<br />

[lly Brtdas, J. L.. Scott, J. C., Yakushi, K., Street, G. B.? Phy. Reu. B 1984, 30, 1023-1025.<br />

[I131 Yakushi, K., Lauchlan, L. J., Clarke, T. C., Street. G. B., J. Chm?. Plzys. 1983, 79, 4774-<br />

4778.<br />

11141 Nechtschein, M., Devreux, F., Genoud, F., Vieil, E., Pernaut, J. M., Genies, E., Synth. Met.<br />

1986, 15, S9-78.<br />

[ 1151 Chung, T.-C., Kaufman, J. H., Heeger, A. J., and Wudl, F., Phys. Rev. B 1984, 30, 702-7 10.<br />

[I161 Kaneto, K., Kohno. Y., and Yoshino, K., Mol. Cryst. Liq. Crj).~l. 1985, /IS, 217-220.<br />

[117] Bradley, D. D. C., Evans, G. P., and Friend, R. H., Syntlr. Met. 1987, 17, 651-656.<br />

[I181 Voss, K. F., Foster, C. M., Smilowitz, L., MihailoviC, D., Askari, S., Srdanov, G., Ni, Z.,<br />

Shi, S., Heeger, A. J., and Wudl, F., Phys. Rev. B 1991, 43, 5109-5118.<br />

[I191 Hill, M. G., Mann, K. R., Miller, L. L., Penneau, J.-F., J. Am. Cliem. Soc. 1992, 1/4, 2728-<br />

2730.<br />

[I201 Stafstrom. S., in: Corzjligated Polyrners: Brtdas, J. L. and Silbey, R. (Eds.). Dordrecht:<br />

Kluwer Academic Publishers, 1991; pp. 113-140.<br />

[I211 Kivelson, S., Heeger, A. J., Plzys. Rev. Lett. 1985, 55, 308-31 1.<br />

[122] Mizes, H. A., Conwell, E. M., Pl7y.s. Reu. Lett. 1993, 70, 1505-1508.<br />

[I231 Shimoi, Y., Abe, S., Phys. Re[). B 1994, 50, 14781-14784.<br />

[124] Bredas, J. L., Thtmans, B., Fripiat, J. G., Andrt. J. M., Chance, R. R., Plzys. Rev. B 1984,<br />

29, 6761-6773.<br />

[125] Stafstrom, S., Brtdas, J. L., Epstein, A. J., Woo. H. S., Tanner, D. B., Huang, W. S., Mac-<br />

Diarmid, A. G.. Phys. R ~P. Lett. 1987, 59, 1464-1467.<br />

11261 Stafstrom, S., Bredas, J. L., Phys. Reu. B 1988, 35, 4180-4191.<br />

11271 Brtdas, J. L., Beljonne, D., Shuai, Z., Toussaint, J. M., Syntlz. Met. 1991, 43, 3743-3746.<br />

[I281 Tanaka, C., Tanaka, J.. Syntlz. Met. 1993, 57, 4377-4384.<br />

[ 1291 Sariciftci, N. S. (Ed.) Primary Plzotoescitatioiis I/7 Conjugated <strong>Polymer</strong>s: Molecular Evciton<br />

I)P~SIISemicoiithrc~tor Bud Model. Singapore: World Scientific, 1997, Chapter 17.


This Page Intentionally Left Blank


5 Vibrational <strong>Spect</strong>roscopy of Polypeptides<br />

S. Kriinm<br />

5.1 Introduction<br />

The polypeptide chain, (NHCH(R)CO),l, is the basic backbone chemical unit of<br />

proteins, and knowledge of its possible structures and dynamics is crucial to an<br />

understanding of the functions of these important biological macromolecules. In<br />

synthetic polypeptides, all the side chains, R, are usually the same; in proteins, they<br />

can be any of 20 ainino acid residues whose specific sequence is determined by the<br />

base sequence in the DNA of the corresponding genome [I]. The unusual property<br />

of these chains is their ability to adopt a wide range of three-dimensional structures<br />

121, which accounts for the richness of biological functions that they can perforin [f].<br />

This presents a special challenge to mastering the spectroscopy of these polymeric<br />

molecules.<br />

While vibratioiial spectroscopy is not capable of the structural resolution of X-<br />

ray diffraction, it nevertheless has some important advantageous features. First, it<br />

is not generally limited by physical state: samples can be in the form of powders,<br />

crystals, films, solutions, membranous aggregates, etc. Second, a number of different<br />

experimental methods probe the structure-dependent vibrational modes of the<br />

system: infrared (IR), Raman (both visible and UV-excited resonance), vibrational<br />

circular dichroism, and Raman optical activity, many of these with time-resolution<br />

capabilities. Finally, in addition to providing structural information, vibrational<br />

spectra are sensitive to intra- and intermolecular interaction forces, and thus they<br />

also give information about these properties of the system.<br />

To gain such insights, however, requires that we achieve the deepest possible<br />

understanding of the vibrational spectrum. Correlations using characteristic band<br />

frequencies, particularly with resolution-enhanced spectra [3], can be usefd, but it is<br />

increasingly being recognized that this is a fundamentally limited approach [4]. For<br />

the kind of understanding we seek it is necessary to determine the detailed normal<br />

modes of the molecule. Only by achieving an accurate description of these modes<br />

can we expect to explain fully all features of the vibrational spectrum and thereby to<br />

validate the structure and force field inputs to the normal-mode calculation [S]. This<br />

is now beginning to be possible for the polypeptide chain.<br />

In this chapter we review the present state of the calculation of accurate normal<br />

modes of polypeptides. Since the burden of such calculations falls primarily on the<br />

quality of the force fields, we first review the state of present methods for obtain-


240 5 l i’hrrrtionnl <strong>Spect</strong>roscopy of Polypeptides<br />

ing such force fields, which include empirical, ~ i h initio, and molecular mechanics<br />

approaches. We then discuss applications to characterizing the aniide modes of<br />

the peptide group. Finally, we describe some of the results that have been obtained<br />

on characteristic polypeptide conformations, namely extended-chain and helical<br />

structures.<br />

5.2 Force Fields<br />

The normal modes of a molecule are determined by its three-dimensional distribution<br />

of atomic masses and its intra- and intermolecular force fields. The spectroscopic<br />

validation of the structure and these interactions derives from a satisfactory<br />

prediction of the frequencies and intensities of the observed IR and Raman bands,<br />

with the proviso that the bands are appropriately assigned to the calculated eigenvectors.<br />

Because of the complexity of the spectra this part of the process cannot be<br />

treated lightly, since more than one band match-up may be possible. A number of<br />

methods are available to verify band assignments [5] (isotopic substitution, dichroism,<br />

etc.), but since these are usually limited in practice, the success of the normal<br />

mode approach will ultimately depend on the reliability of the force fields. We<br />

therefore devote part of this section to a discussion of the development of such<br />

accurate force fields. This development has proceeded primarily through empirical<br />

refinement of force constants but is relying increasingly on inputs from ab initio<br />

calculations and will probably ultiinately be based on inolecular mechanics (MM)<br />

energy functions.<br />

The development of an empirical force field requires selecting a physical niodel<br />

for the potential, choosing which terms are important, and optimizing the force<br />

constant parameters by a least-squares fitting of calculated normal modes to<br />

observed bands. (Since there have been summary [5] as well as detailed [6-8]<br />

treatments of the methods for obtaining the frequencies (eigenvalues) and forms<br />

(eigenvectors) of the normal inodes for a given force field, we will not repeat this<br />

formalism here but will refer only to those aspects that are relevant to our discussion.)<br />

It is useful to recall that while the computational challenge becomes quite<br />

demanding for polypeptide structures lacking symmetry, such as globular proteins,<br />

for those structures that have helical or translational symmetry, such as the helical<br />

and extended-chain forms of synthetic polypeptides, the calculational problem is<br />

much simplified. This is because the only modes that can exhibit IR or Raman<br />

activity are those in which equivalent vibrations in each helix unit differ in phase by<br />

0, &[ , or ?2


5.2 FOYCC Fields 241<br />

such calculations can only be done for relatively small molecules, and the force<br />

constants generally need to be scaled so that calculated frequencies agree with<br />

experimental data. Nevertheless, valuable information can be obtained by this<br />

approach.<br />

The MM method will be seen to be the most useful one for obtaining force constants<br />

that can be used for different conformations of the molecule. In this approach<br />

the force constants are obtained from the second derivatives of an assumed potential<br />

energy function consisting of quadratic bonded terms and non-quadratic nonbonded<br />

terms. Although present MM functions are too crude to be spectroscopically<br />

reliable, a new method for deriving such functions holds the promise of<br />

providing a reliable vibrational force field for the polypeptide chain.<br />

5.2.1 Empirical Force Fields<br />

The most general force field of a molecule would include anharmonic as well as<br />

harmonic terms. However, with the limited experimental information generally<br />

available for refining an empirical force field for complex molecules, the harmonic<br />

approximation is the only feasible one at present. This means that, for the isolated<br />

molecule, we need to know the force constants, FQ, in the quadratic term of the<br />

Taylor series expansion of the potential energy, V:<br />

where r is, for example, an internal displacement coordinate and IZ = 3N - 6, N<br />

being the number of atoms in the molecule. If interactions with other molecules are<br />

involved, analogous intennolecular interaction energy terms must be included.<br />

Since the maximum number of observable frequencies is n, whereas the general<br />

number of F,s is n(n + 1)/2, it is necessary to find a physically meaningful model<br />

for T7 that brings the number of empirically determined F,Js into reasonable relation<br />

to the number of observable frequencies, even including those of isotopic derivatives.<br />

Two main models have been used to accomplish this: the Urey-Bradley<br />

force field (UBFF) and the general valence force field (GVFF). Both models incorporate<br />

the same diagonal (i.e., Fll) valence-type ternis, involving bond stretch, angle<br />

bend, and torsion coordinates, but differ in how they treat the off-diagonal (i.e., F,,)<br />

terms.<br />

In the unmodified UBFF, elaborated by Shimanouchi [lo], off-diagonal terms in<br />

T’ are represented by atom pair 1,3 non-bonded interactions. Because this introduces<br />

inherent redundancies in the coordinates, the equilibrium configuration of the<br />

molecule represents a state with internal tensions, and therefore linear terms must<br />

be included in the force field. By introducing the redundancy condition, the linear<br />

terms can be eliminated from V, although the coefficients of the quadratic terms<br />

then contain ‘intramolecular tension’ as well as the valence and nnn-bonded force<br />

constants. The important characteristic of such a UBFF is that only a limited


242 5 Vibrcitional Spcc.ii.o.scoyj~ of Po1jpt.ytitlt.s<br />

number of valence-type off-diagonal terms are in effect included in the potential.<br />

Such simple UBFFs can therefore be physically inaccurate, and as B result they<br />

have had to be modified by the explicit inclusion of some additional F;,s [lo].<br />

In the GVFF there is no inherent limit to the number of Fi,s that are included.<br />

There is only the restraint that the number of Fs should not exceed the number or<br />

observable frequencies; in fact, it should generally be much smaller so as to overdetermine<br />

the force field. Since an independent set of coordinates can be chosen,<br />

e.g.. local symmetry coordinates, and the redundancy conditions explicitly given.<br />

there is no need to include linear terms in V, and Eq. (51) is the most general<br />

representation of the force field. Our discussion will focus on applications of the<br />

GVFF.<br />

The refinement of an empirical force field for a niacroinolecule usually starts with<br />

the vibrational analysis of a smaller model system of known structure followed by a<br />

transfer of these force constants to the larger system, with possible subsequent force<br />

constant adjustments. In the case of the polypeptides, the preferred model system<br />

for the peptide group has been trans-N-methylacetamide (NMA), CH3CONHCHi.<br />

Synthetic polypeptides of known structure such as P-sheet polyglycine (PG), R = H,<br />

and a-helical and P-sheet poly(L-alanine) (PLA), R = CH3, have served as basic<br />

examples of the polypeptide systems.<br />

The force field refinement of the model system introduces important requirements<br />

and constraints [5, 7, 81.<br />

1. The types of F, to be included in V must be selected. In practice, this has been<br />

based on prior experiences with the GVFF, but nevertheless the choices are, in<br />

general, arbitrary.<br />

2. A starting set of force constant values, from which are calculated a set of normal<br />

coordinates, Pa, and normal frequencies, vx, needs to be chosen. These are<br />

refined to observed data by a least-squares optimization procedure. We note that<br />

a normal mode Q.* can be visualized in terms of the local, usually symmetry ($1,<br />

coordinates since<br />

S;=CL 1% p %<br />

1<br />

(5-7)<br />

and the normal mode calculation provides the elements L,, of the eigenvector<br />

matrix. Thus, for a given QLy the relative S, are given by the relative Lla. Visualization<br />

in Cartesian coordinates is also possible. An analogous description is given<br />

by another characteristic of the normal mode, namely the potential energy distribution<br />

[PED). This consists of the relative contributions of each S, to the total<br />

change in potential energy during QE, the fractional contribution of a particular<br />

S, being given by<br />

where I,, = 47c'c'v~. vu in cm-'. (Contributions from F,j elements are typically<br />

small and ai-e usually neglected. However, they can be large and negative, which


5.2 Force Fields 243<br />

accounts for occasional unusually large contributions in a PED. In such a case,<br />

the diagonal elements alone do not give a sufficiently complete description of the<br />

norinal mode.)<br />

3. Correct band assignments must be made in matching calculated to observed v,.<br />

This is probably the most important part of the procedure, since acceptable frequency<br />

agreement (usually 5-10 cm-' ) does not necessarily insure that correct<br />

eigenvectors have been obtained. There are a number of properties that help<br />

to judge the validity of band assignments: molecular symmetry, which specifies<br />

the activity of a mode, and in appropriate situations predicts the IR dichroisni<br />

(parallel, 11, or perpendicular, I, to the chain axis) and Raman polarization<br />

characteristics of the bands; isotopic substitution, which identifies the general<br />

nature of a mode, and by the reproduction of specific observed frequency shifts<br />

substantiates the correctness of the eigenvectors; overtone and combination<br />

bands, which should be consistent with the assignments of fundamentals; and<br />

predicted band intensities, which are very sensitive to the eigenvectors. The<br />

intensity agreement is particularly important since it provides a test of the<br />

eigenvectors that is independent of the frequencies. For example, the IR<br />

absorbance, A,, associated with normal mode Qa is given by [l 11<br />

A, = 42.2547 (g) (5-4)<br />

where dG/aQ, is the normal coordinate derivative of the dipole moment, viz., the<br />

transition dipole moment (Dk' amup'l2), and A, is in km mol-'. Since from<br />

Eq. (5-2)<br />

= Li, (g)<br />

(5-5)<br />

Qu<br />

we see that, through the eigenvector matrix, the IR intensities sample another<br />

physical property of the molecule, namely the dipole derivatives, which arise<br />

from the charge re-distributions during the vibrations. Such dipole derivatives<br />

can also provide a measure of the coordinate contributions to the IR intensity of<br />

a band through a dipole derivative distribution (DDD)[ 121. In addition, relative<br />

intensities of different symmetry species of a mode can provide information on<br />

structure. Thus, in a sample with a molecular chain axis aligned (say mechanically)<br />

in a known direction, since the orientation of a specified dc/aQ, with<br />

respect to the chain axis can be related mathematically to an experimentally<br />

observed dichroic ratio, DR-A,( ll)/Av(L), a DR measurement can help to<br />

establish the dG/L7Q, orientation and therefore provide a test of a structural<br />

hypothesis.<br />

The refinement of the macromolecular force field from that of the model system<br />

follows concepts similar to those described above, with three important additions.<br />

The first, and simplest, is that since the structures are different, new kinds of force


244 5 Vibrational <strong>Spect</strong>roscopy oj Polypeptides<br />

constants must be introduced. This is not a major problem since initial values, for<br />

example for polypeptide side chains, are generally available. Second, the macromolecular<br />

system generally experiences inter- as well as intramolecular interactions,<br />

and these have to be incorporated into the force field. For polypeptides, these include<br />

hydrogen bonds and non-bonded interactions. Third, since for polypeptides in<br />

particular we are especially concerned with describing different molecular conformations,<br />

attention has to be given to incorporating the conformation dependence<br />

of the force constants. This is not easy to do in the fi-amework of an empirical force<br />

field.<br />

With respect to the intermolecular interactions, the hydrogen bonds can be<br />

treated like the other internal coordinates, i.e., in term of bond stretching, angle<br />

bending, and torsion force constants. However, since hydrogen bonds vary in their<br />

geometry, we should ideally know how these force constants depend on the structure<br />

of the bond. While preliminary studies have been done for the NH hydrogen<br />

bond of the peptide group [13], there is at present no comprehensive description of<br />

this relationship.<br />

The itnportant non-bonded interactions are of two kinds, dispersion and transition<br />

dipole moment interactions. The former correspond to the so-called van der<br />

Waals forces and are generally given by a Lennard-Jones type of potential, viz.<br />

where R,, is the distance between atoms i and j, the exponent of the repulsive term<br />

is traditionally 12 (although 9 has been found to be more effective in some cases),<br />

the exponent of the attractive term is for basic reasons 6, and A, and B, are respectively<br />

the repulsive and attractive van der Waals parameters of atom i. Such<br />

potentials have generally not been added to a GVFF for polypeptides since, in<br />

distinction to polyethylene [ 141, they do not affect midrange and high frequencies<br />

significantly. However, they probably are important in describing low-frequency<br />

modes.<br />

Transition dipole moment interactions between peptide groups can influence<br />

higher frequency modes, as was shown by their effects on the splittings of amide I<br />

and amide I1 modes of a-helix and P-sheet polypeptides [15, 161. Such transition<br />

dipole coupling (TDC) arises from the potential energy of interaction, V&, between<br />

transition moments, Aji, in different peptide groups, and is given by [17]<br />

where LY4, is a geometrical factor given by<br />

with E being the direction of the transition moment and YAB the distance between the


5.2 Force Fields 245<br />

centers of the moments. In terms of normal coordinate dipole derivatives,<br />

A@ = (4.1058/~:’~) ~<br />

(5-9)<br />

sg<br />

(3 P a<br />

where v, is the unperturbed frequency. In view of Eq. (5-51, a calculation of Vlfcl will<br />

require a summation over terms in 1 dcA/dSi I 1 ?yB/?5’, 1 . The force constant associated<br />

with F’(/(/ is<br />

in mdyn k’, and if we take the effect of TDC as a perturbation on va, the frequency<br />

shift is given by<br />

(5-1 1)<br />

Early applications of this theory [18, 191 parametrized d@/dQ from observed band<br />

splittings. Ah initio calculations of these dipole derivatives [20] showed that the<br />

empirical TDC parameters were in agreement with nb initio as well as IR intensity<br />

results [ 171, thus emphasizing that the TDC mechanism is an important intermolecular<br />

interaction in polypeptides, particularly for modes with large d@/aQ such as<br />

amide I and ainide 11.<br />

The matter of the conformation dependence of the force field of a polypeptide<br />

is one that has received little attention. We know that the differences in vibrational<br />

frequencies of different conformers derive mainly from the differences in their threedimensional<br />

structures, with force constant differences playing a secondary role.<br />

Nevertheless, certain spectral features cannot be understood without taking into<br />

account variations in force constants with conformation (a good example is the<br />

tvans versus gnzrche angle-angle interaction force constant [213). Independent<br />

empirical refinements of a-PLA and /I-PLA also demonstrate such force constant<br />

differences [5]. The problem is how to express such variations in a general explicit<br />

form in V. This is difficult to achieve simply from knowledge of independently refined<br />

force fields of different conformers. We return to this question below in our<br />

discussion of MM force fields.<br />

Empirical force fields have been developed for NMA, for PG, and for CY-PLA and<br />

/I-PLA. In the case of PLA, it has also been possible to refine an ‘approximate’ force<br />

field in which the CH3 side chain is replaced by a point mass (of 15 amu) [22]. This<br />

is useful in studying modes of the backbone chain that do not couple significantly<br />

with side-chain modes.<br />

5.2.2 Ab Iizitio Force Fields<br />

Ab initio molecular orbital calculations, by giving the potential energy surface of<br />

a molecule at its minimum energy conformations, provide a complete vibrational


246 5 Vibrationcil <strong>Spect</strong>roscopj.1 of Pol.vpeptides<br />

force field, i.e., all diagonal and off-diagonal force constants [23]. Since the force<br />

constants are given by the second derivatives of the potential in Cartesian, and<br />

therefore non-redundant, coordinates, the force field is unique. Transformation is<br />

then always possible into an internal or local symmetry coordinate basis. Dipole<br />

and polarizability derivatives are also obtained from the ah iiiitio calculation, and<br />

therefore IR and Ranian intensities can be predicted.<br />

Aside from the previously mentioned problem of doing such calculations on very<br />

large molecules, there are still some limitations at present in directly implementing<br />

ab iriitio force field calculations on model systems. At the Hartree-Fock (HF)<br />

level, force constants can be 10-30% too large because of basis set limitations and<br />

the neglect of electron correlation. This necessitates the empirical scaling of force<br />

constants so that calculated frequencies agree with observed. While systematic<br />

scaling procedures have been developed [24], there is still variability in the number<br />

and kind of scale factors that are used. In any case, a careful comparison of basis<br />

sets in terms of their geometry and frequency predictions is still very important. For<br />

example, the structure of NMA has non-planar symmetry at the HF/6-31G* level,<br />

but it exhibits planar symmetry at the HF/6-31 +G* level, i.e., with the addition<br />

of diffuse functions [25]; some well-assigned normal modes of malkanes cannot<br />

be reproduced in the proper order at the scaled HF/6-31G* level but are correctly<br />

calculated with a scaled HF/6-31G basis set [26]. With the inclusion of electron<br />

correlation, for example by the Mdler-Plesset (MP) perturbation method, the force<br />

constants are much closer to experiinentally compatible values, and such calculations,<br />

though very computer-intensive as the molecule gets larger, can provide<br />

usable force fields with minimal scaling.<br />

Despite these problems, ab irzitio force fields provide an improved route to<br />

avoiding the arbitrariness and incompleteness of empirical force fields. Although<br />

such calculations on model systems do not translate into force fields for macromolecules,<br />

they can provide important components of such force fields. For example:<br />

an ab initio force field for glycine hydrogen bonded to two HzO molecules [27]<br />

provided the necessary starting point for refining the carboxyl group part of the<br />

empirical force field for glutathione [28]; an ah iiiitio force field for diethyl disulfide<br />

[29] permitted detailed correlations to be made between SS and CS stretching frequencies<br />

and the conformations of disulfide bridges in proteins [30, 311; an ab iiiitio<br />

force field for ethanethiol, together with an empirical force field for the peptide<br />

group, provided detailed structure-spectra correlations between CS and SH stretch<br />

frequencies and the conformation of the cysteine side chains in proteins [32].<br />

Ab iizitio force fields also demonstrate that force constants vary with conforination,<br />

a factor that needs to be taken into account if we are to obtain an accurate<br />

description of the normal modes associated with the many possible conformations<br />

of the polypeptide chain. This is clearly illustrated by a study of the scaled ah irzitio<br />

force fields of four iion-hydrogen-bonded conformers of the alanine dipeptide,<br />

CH3CONHCaH(CPH3)CONHCH3 [33, 341. It was found that diagonal force constants<br />

such as the torsion (t) change by very large amounts, and even the stretch (s)<br />

and deformation (d) force constants are significantly dependent on the conforniation.<br />

Off-diagonal force constants can also change significantly with conformation.<br />

In principle this is not surprising since we must expect changes in electronic struc-


5.2 Force Fields 247<br />

ture with conformation, In terms of conveniently designated components of the<br />

energy of the molecule, which of course the ah initio calculation does not provide,<br />

we can conceive of these changes as arising from changes in the charge distribution<br />

with conformation, from changes in the dispersion energies as interatomic distances<br />

change, and from any iiitrinsic valence force constant variations with conformation.<br />

While we can show from ab initio calculations that the force field is not independent<br />

of molecular conformation, it is much more difficult to specify the explicit<br />

nature of this dependence. This is precisely because the method provides only a<br />

total energy for a given structure. If we wish to infer the varying contributions of<br />

different physical components, we must assume a model for the potential energy<br />

and strive to make its behavior mimic the ah initio results as closely as possible. It is<br />

toward this goal that some present efforts are directed.<br />

5.2.3 Molecular Mechanics Force Fields<br />

An MM energy function, whose second derivatives at a minimum are the spectroscopic<br />

force constants, is generally represented as a sum of quadratic and nonquadratic<br />

terms. The foimer encompass the valence-type deformations, such as<br />

bond stretching and angle bending, while the latter involve torsions, dispersion<br />

interactions, electrostatic interactions, and possible hydrogen-bond terms.<br />

In order to serve as a spectroscopic force field, such an MM function would at<br />

least have to reproduce vibrational frequencies to spectroscopic standards, viz., to<br />

root-mean-square (rms) errors of the order of 5-10 cnir'. Some early MM potentials<br />

came close to this mark [35], but since frequency agreement was not their primary<br />

goal (but rather reproduction of structures and energies), subsequent efforts<br />

resulted in functions with unacceptable spectroscopic predictability (rms errors of<br />

>50 cm-'). One of the reasons for this was the lack of proper attention to crossterms<br />

in the potential. Initial efforts to improve such potentials have led to only<br />

marginally better frequency agreement. For example, MM3 gives an rms error for<br />

frequencies below 1700 cnir' of 38 cm-' for butane [36a] (reduced to 22 cm-' in<br />

MM4 [36b]) and 47 cm-l for NMA [37]; a Hessian-biased force field [38] gives a<br />

similar rms error of 29 cn1-l for polyethylene [39]; for CFF93, the ims error for n-<br />

butane is 25 cmrl 1401 and is 34 cnrl for a group of four acyclic and three cyclic<br />

hydrocarbons; and a second generation version of AMBER gives 43 cm-' for<br />

NMA [41]. Some force fields have been specifically designed to give improved<br />

spectroscopic predictability and these [42, 431 indeed show better agreement, although<br />

improvement is still desirable: the rms error for the above NMA frequencies<br />

is 24 cn-l for SPASIBA [44].<br />

What is clearly needed is an approach that systematically incorporates spectroscopic<br />

agreement in the initial stages of the optimization of the MM parameters.<br />

This has been achieved by a procedure designed to produce a so-called spectroscopically<br />

determined force field (SDFF) [45, 461.<br />

The elements in the SDFF methodology are the following. First, a form is selected<br />

for the MM potential, one that is hopefully inclusive enough to incorporate all the


248 5 Vibrational <strong>Spect</strong>roscopy qf Polypeptides<br />

physically important contributions. The first iinplementation of this approach, viz.,<br />

for linear saturated hydrocarbon chains [47], used a potential of the form<br />

where the y,s are internal coordinates whose intrinsic (i.e., equilibrium) values are<br />

ql0 and the flJ are the intrinsic MM quadratic (i.e., valence-type) force constants. At<br />

this stage all possible cross-terms are included. The intrinsic torsion potential is<br />

represented by a Fourier cosine series,<br />

(5-13)<br />

where V, is the barrier to a rotation of periodicity m associated with torsion angle<br />

x. (In general, Vr,,. is a sum over terms in a Fourier series, but in the hydrocarbon<br />

case only the V3 term was found to be necessary.) The non-bonded potential consists<br />

of the dispersion term (an exponent of 9 was found to be more suitable for the<br />

repulsive term) plus a coulomb teim that is usually represented by the interactions<br />

between partial charges Qi and Q, on atoms i and j, i.e., in the case of the hydrocarbons<br />

(5-14)<br />

where K is the dielectric constant and EO the permittivity of free space. The summation<br />

of Vf& is over all atom pairs 1,4 and higher. The Q1 can be fixed charges or they<br />

can include charge fluxes, i.e., dQ1/dSJ [42]. (Inclusion of charge fluxes also permits<br />

the calculation of TR intensities.)<br />

The second element in the procedure is the choice of a spectroscopic force field,<br />

i.e., a set of F,J. This could be an empirical force field, but since we wish to provide<br />

as complete a description as possible at this stage, a scaled ab initio force field is the<br />

one of choice. The scaling process also connects the force constants to experimental<br />

frequencies and band assignments, and avoids the problems of having incorrect<br />

eigeiivectors because of using inappropriate basis sets [26].<br />

The most important part is the third element, viz., the ability to make a transformation<br />

from the spectroscopic to the MM force field [46]. Since this transformation<br />

is analytic, it preserves in the SDFF the frequency as well as structure<br />

agreement of the scaled ah initio calculation. Although the transformation is initiated<br />

by assuming a starting set of p:lt, parameters, these are subsequently optimized<br />

in the refinement procedure [48]. In addition to providing the flJ, the transformation<br />

procedure also gives the ql0 [46].<br />

The fourth, and also important, element is the application of this transformation


5.3 Amide Modes 249<br />

to the nh initio force fields of a set of conformers of the model molecules for the<br />

macromolecular system. In the case of the hydrocarbons, this consisted of the four<br />

stable conformers of M-pentane and the ten stable conformers of n-hexane [26, 471.<br />

At this stage, the non-bonded parameters are optimized under the standard MM<br />

assumption that the intrinsic force constants (fl,) and intrinsic geometry parameters<br />

(ql0) are the same for all conformers 1481. The relative (ah initio) conformer barriers<br />

and energies are incorporated in optiniizing the V, and Q1. At this point, specific<br />

forms of the conformation dependence of some fiJ can be determined and incorporated<br />

in V.<br />

Finally, since it is neither desirable nor necessary to include in I/ the huge number<br />

off,,, most of which are very small and do not significantly affect the vibrational<br />

frequencies, the set of force constants is systematically reduced based on a preassigned<br />

limit on the frequency error [46]. In this process, the remaining fl, are<br />

optimized, usually requiring minimal change, to compensate for the effects of the<br />

force constants that were dropped.<br />

There are a number of advantages to the SDFF approach to refining an MM<br />

force field. By utilizing a complete (i.e., ab iizitio) spectroscopic force field scaled<br />

to experimentally assigned bands, frequency agreement is assured at the start. This<br />

not only incorporates highly accurate information about the minima in the potential<br />

surface, but it avoids the possibility of biasing the non-bonded parameters to<br />

compensate for an incomplete formulation of the fl, terms in V. In addition, since<br />

the optimization is done in a nonredundant coordinate basis [49], the uniqueness of<br />

the force field is assured; subsequent transformation to a redundant basis of choice<br />

is always possible. Finally, since force fields for many different conformers are<br />

involved in the optimization, there is a better chance of revealing the nature of<br />

specific conformation dependences of the cross-terms.<br />

The application of this procedure to an SDFF for the hydrocarbon chain has<br />

given excellent results [47]. With a set of 50 force constant and geometry parameters<br />

the observed frequencies of tt-pentane and ttr-hexane below 1500 c1n-l could be<br />

reproduced with an rms error of N 11 cm-'. This force field has now been extended<br />

to include branched saturated hydrocarbons [50]. An SDFF for the polypeptide<br />

chain is now under development [51].<br />

5.3 Amide Modes<br />

Since the peptide group, -CONH-, is the most distinctive component of the backbone<br />

of the polypeptide chain, it is not surprising that the normal modes of this<br />

group, the so-called aniide modes, were first studied in the simplest representative<br />

molecule, viz., NMA. It is useful to examine the nature of these modes in NMA and<br />

in blocked dipeptides since they have served as important guideposts in the analysis<br />

of polypeptide spectra in terms of structure.


250 5 Vibuntiorid Si7ectvoscop.y of Polypeptides<br />

5.3.1 N-Methylacetamide<br />

Normal mode calculations on NMA have been done with both empirical and scaled<br />

ab iizitio force fields. In the empirical force field calculations, the CHj group has<br />

been treated as a point mass, using a UBFF [52] as well as a GVFF [51, and as an<br />

all-atom structure, again with a UBFF [53] and a GVFF [54, 551. In the ah iizitio<br />

calculations, a variety of basis sets and scaling methods have been used: 4-31G,<br />

with adjustment of mainly diagoiial force constants to liquid-state fi-equencies of<br />

NMA and its isotopomers [56]; 4-31G and 4-31G*, with no scaling [57]; 4-21, with<br />

four general scale factors and some approximated force constants [58]; and 4-31G*,<br />

with 10 scale factors [59], optimized to matrix-isolated frequencies [60].<br />

All of the above calculations have been on the isolated molecule, even though in<br />

many cases comparisons have been made with experimental results on hydrogenbonded<br />

systems (such as the neat liquid). The first noimal mode calculation on<br />

a hydrogen-bonded system was at 4-31G* on NMA-(H20)2, in which one H2O<br />

molecule was hydrogen bonded to the CO group and the other to the NH group,<br />

with scale factors optimized to IR and Raman data on the aqueous system [61]. In<br />

view of the absence of infomiation on the detailed water structure associated with<br />

the NMA molecule in aqueous solution, this was considered to be a satisfactory<br />

minimal model for representing the normal modes of hydrogen-bonded NMA.<br />

(A 4-31G calculation 011 NMA-(H20)3 has also been done [62].) Subsequently, a<br />

3-21G calculation was done on NMA-(fomiamide)z, NMA-(FA)?, using six<br />

general scale factors [63]. In view of the determination that diffhe functions are<br />

necessary to give a planar symmetric peptide group in the isolated molecule 1251,<br />

the NMA-(H20)2 calculations have been repeated at 6-31 + G* [64], and these are<br />

discussed below.<br />

Early studies [52, 65, 661 sought to describe the characteristic vibrational modes<br />

of the peptide group. In addition to the localized NH s mode, which is subject to<br />

Fermi resonance interactions [5], there are other relatively localized modes, labeled<br />

amide I-VII (although it was recognized [66] that some of these, because of their<br />

delocalization, would be more variable than others). These amide modes have been<br />

used as a basis for discussions of peptide spectra, and in Table 5-1 we describe them<br />

and compare the PEDs for isolated and aqueous hydrogen-bonded NMA. It should<br />

be noted that these results are on fully optimized structures (all frequencies real),<br />

in both cases the conformation being N,Ct. (The symbol indicates that the in-plane<br />

N-methyl hydrogen is cis to the (N)H and the in-plane C-methyl hydrogen is trans<br />

to the (C)O.)<br />

The amide I mode is primarily CO s, both in NMA and NMA-(H20)2. However,<br />

the additional contributions to the eigenvector depend very much on the<br />

specifics of the calculations (only contributions to the PED of 2 10 are given in<br />

Table 5-l), as well as, of course, on the absence or presence of hydrogen bonding.<br />

For NMA, the next largest contributor in a CH3 point-mass calculation is CN s<br />

[52, 51, whereas in an all-atom calculation it is CCN d [55, 591. The situation for<br />

NMA-( H20)2, where the lower frequency is an obvious consequence of hydrogen<br />

bonding, is complicated by the experimentally demonstrated coupling of CO s<br />

and HOH b 167, 681, which makes the latter the next largest contributor. In NMA-


5.3 Ainide Modes 251<br />

Table 5-1. Amide modes of isolated and aqueous hydrogen-bonded N-Methylacetamide"<br />

Mode v(obsib v(calc)c Potential energy distributiond<br />

I 1706<br />

1646<br />

1626<br />

I1 1511<br />

1584<br />

111 1265<br />

1313<br />

IV 658<br />

632<br />

V 391<br />

(7501"<br />

v1 626<br />

(600)'<br />

VII ~<br />

195g<br />

1708<br />

1632<br />

1626<br />

1512<br />

1575<br />

1264<br />

1305<br />

648<br />

640<br />

391<br />

744<br />

627<br />

605<br />

-<br />

20s<br />

CO ~(83) CCN d(l1)<br />

CO s(30j [O)HOH b(30) fH)HOH b(18)<br />

CO s(38) (0)HOH b(28) (H)HOH b(1S)<br />

NH ib(5l) CN s(28)<br />

CN s(48j NH ib(36)<br />

NH lb(21) CO lb(21) CN ~(18) CC S(10) CCH3 sb(l0)<br />

NH lb(26) CN ~(20) CCHi sb(20) CO lb(12)<br />

CC ~(36) CO ib(34)<br />

CO ib(35) CC s(32)<br />

CN t(118) NH ob(51) CO ob(23)<br />

CN t(43) H 0 b2128) NH ob(21)<br />

CO ob(67) CCH3 or(26) CN t( 11)<br />

CO ob(50) 0 . H b2(30) CCH3 OI(19) NH ob(l0)<br />

CN t(38) NH ob(21)<br />

First line: NMA; second line(s): NMA-(H20)2. 6-31+G* basis set.<br />

NMA: from [59]; NMA-(H20)2: from 1611 and 1681, except where noted.<br />

NMA: from [73]: NMA-(H20)2: from [64].<br />

s: stretch, d: deformation. b: bend. ib: in-plane bend, sb: synmetric bend, ob: out-of-plane bend,<br />

b2: essentially ob, or: out-of-plane rock, t: torsion. Group coordinates are defined in [59]. Contributions<br />

> 10.<br />

Possible value, from 1691.<br />

Observed in neat NMA [65].<br />

From 1621 (at 206 cm-' from [70]1.<br />

(FA)2 this contribution is NH in-plane bend (ib) [63], but such a result is uncertain<br />

because of the absence of experimental data on this complex and the resulting lack<br />

of optimization of the FA force constants. Nevertheless, we see that even for as<br />

localized a mode as CO s the specific nature of the eigenvector can depend on the<br />

details of the calculation.<br />

The amide I1 mode is a mixture of NH ib and CN s, with the former dominating<br />

in NMA and the latter, at a higher hydrogen-bonded frequency, dominating in<br />

NMA-(HzO)?, (Table 5-1). It is interesting that the CN s predominance does<br />

not appear at 3-21G for NMA-(FA)> [63], nor at 4-31G [62] or 4-31G" [61] for<br />

NMA-(H20)2, showing the evident influence of basis set on the predicted eigenvectors.<br />

Nor is CN s predominant in polypeptides [5].<br />

The amide I11 mode is also a mixture of NH ib and CN s, although other coordinates<br />

make significant contributions. Since CH3 symmetric bend (sb) is an important<br />

one, CH3 point-mass calculations [52, 51 distort the relative importance of<br />

CC s. As seen from Table 5-1, CO ib is a relatively large contributor in NMA and a


252 5 Vibvritional <strong>Spect</strong>roscopji of Polvpeptides<br />

small one in NMA-(HzO)?. (It is not clear why NH ib makes no contribution in<br />

NMA-(FA)z [63].) The position and nature of this mode is variable in polypeptides<br />

[5], since NH ib contributes significantly to a number of normal modes in the 1400-<br />

1200 cn-' region, with tlie mixing being dependent on tlie side-chain structure.<br />

The amide IV mode is a mixture of CO ib and CC s in NMA, although the latter<br />

contribution is often replaced by other coordinates in polypeptide structures [j]. This is<br />

an example of a peptide group mode that is not in general localized and well defined.<br />

The ainide V mode, on the other hand, seems to be a characteristic band of<br />

hydrogen-bonded peptide groups, although an unusual feature seems not to have<br />

been noted previously. This mode, which is generally a combination of CN t and<br />

NH out-of-plane bend (ob), has been assigned to a band near 725 cm-' in neat<br />

NMA [65, 661 and possibly -750 cni-' in aqueous NMA [69]. It is found to have a<br />

counterpart, the amide VII mode, calculated at -200 cm-' [54, 55, 51 and observed<br />

near this frequency [62, 701. However, ah iriitio calculations on an isolated molecule<br />

[57, 591 produce only a single CN t, NH ob mode at -400 cn-', consistent with an<br />

observed Ar-matrix band at 391 cnir' [71]. The increase in the NH ob force constant<br />

of an isolated molecule needed to give a hydrogen-bonded frequency in the<br />

700 cm-' region apparently divides the contributions of these coordinates between<br />

the two frequency regions. Isolated molecules also exclude the contributions from<br />

hydrogen-bond deformation modes (see Table 5-1).<br />

The amide VI mode is predominantly CO ob, with a significant CCH3 outof-plane<br />

rock (or) contribution. This, of course, cannot be represented in a CH3<br />

point-mass model [5]. The nature of this mode is also significantly dependent on<br />

conformation in polypeptide structures [5].<br />

5.3.2 Blocked Dipeptides<br />

Although the analysis of NMA serves to introduce the representative types of<br />

characteristic modes of the peptide group, it does not provide insights into how<br />

these modes might be influenced by varying conforniational interactions between<br />

adjacent groups, such as can occur in polypeptide chains. For this purpose it is<br />

useful to examine the modes of a blocked dipeptide, such as N-acetyl-L-alanine-Ninethylamide,<br />

the Ala dipeptide. This was first done for just the amide I and amide<br />

I11 modes using a fixed empirical force field and no TDC [72], so the conclusions are<br />

necessarily limited. Better insights can be obtained from an ab iriirio study of four<br />

non-hydrogen-bonded conformers of this molecule [33].<br />

In Table 5-2 we show the calculated aniide I, 11, 111, and V frequencies and PEDs<br />

of the four non-hydrogen-bonded conformers of the Ala dipeptide, which differ<br />

mainly in the values of the p(NCa), $(C"C) torsion angles (see Figure 5-1). (The<br />

amide IV and VI modes are no longer distinguishable as relatively pure modes, CO<br />

ib and CO ob contributing significantly and mixing with other coordinates in the<br />

region of -830-515 cm-' [33].) It should be noted that in this calculation the 4-21<br />

ah iizitio force constants were scaled with six general scale factors, so the frequencies<br />

are not necessarily representative of true experimental values. However, the changes<br />

in frequency should be indicative of the effects of changes in conformation.


Table 5-2. Some amide modes of four non-hydrogen-bonded conformers of the alanine dipeptidea.<br />

Mode<br />

Conformerb<br />

I 1719<br />

CO sI(78)"<br />

1694<br />

CO s2(81)<br />

1726<br />

CO sl(84)<br />

1679<br />

CO s2(84)<br />

1724<br />

CO sl(79)<br />

1703<br />

CO s2(83)<br />

1720<br />

CO s2(83)<br />

1710<br />

CO sl(82i<br />

I1 1550<br />

NH ib2(53)<br />

CN s2(27)<br />

1532<br />

NH ibl(61)<br />

CN sl(18)<br />

1552<br />

NH ib2(57)<br />

CN s2(27)<br />

1526<br />

NH ibl(61)<br />

CN sl(18j<br />

1545<br />

NH ib2(61)<br />

CN ~2128)<br />

1536<br />

NH ibl(68)<br />

CN sl(19)<br />

1540<br />

NH ib2(65)<br />

CN ~'(25)<br />

1529<br />

NH ibl(68)<br />

CN sl(22)<br />

I11 1265<br />

CN sl(28)<br />

CN s2(12)<br />

CO ib(12)<br />

1252<br />

H" b2(27)<br />

CN s2( 16)<br />

CN sl(14)<br />

1292<br />

CN s2(29)<br />

C'C s(22)<br />

NH ib2(22)<br />

CO ib2112)<br />

1253<br />

H" bl(33)<br />

CN sl(24)<br />

NH ibl( 11)<br />

1265<br />

CN sl(39)<br />

CO ibl(l7)<br />

CC s( 14)<br />

NH ibl(13)<br />

1247<br />

CN ~2126)<br />

H" b2172)<br />

H" bl(13)<br />

NC s(l1)<br />

1295<br />

C"C s(28)<br />

CN ~327)<br />

NH ib2( 12)<br />

CO ib2(11)<br />

1260<br />

Ha b2(37)<br />

CN sl(19)<br />

NH ibl(l2)<br />

v 672<br />

NH ob2(42)<br />

NH ob1135)<br />

CN t2(20)<br />

CN tl(17)<br />

NC" t(13j<br />

558<br />

NH ob2(48)<br />

NH obl(38)<br />

CN tl(12)<br />

CN t2(21)<br />

NC" t( 14)<br />

CO ibl(13)<br />

657<br />

NH ob2(28)<br />

NH ob1124)<br />

CN t2111)<br />

547<br />

NH obll61)<br />

CN tl(47)<br />

NH ob2(36)<br />

NC" t(28)<br />

CN tl(18)<br />

585<br />

CN t2(44)<br />

NH ob2(40)<br />

573 546<br />

NH ob2(40) NH obl(ll3)<br />

CN t2(13) CN tl(67)<br />

C"C s(l2) NC" ti65)<br />

NH obl(12)<br />

CO ob2( 10)<br />

605<br />

NH ob2(47)<br />

NH obl(451<br />

CN t2(29)<br />

CN tl(2l)<br />

NC' t(16)<br />

CO oblill)<br />

504<br />

NH obl(711<br />

NH ob2155j<br />

CN tl(S3)<br />

NC" t(43r<br />

CN t2(38)<br />

a From [33].<br />

q. li/ - p2: -134", 38"; C(R: -92", -6"; C(L: 61", 41"; a': -162", -55".<br />

s. stretch. ib: in-plane bend, b: bend, ob: out-of-plane bend, t: torsion. Group coordinates are defined in [34]. Contributions 2 10.


254 5 Vibrational <strong>Spect</strong>voscopj? of Polypeptides<br />

Keeping in mind that the peptide groups are not identical, because of the different<br />

chemical surroundings, we find that aniide I frequencies, because of their highly<br />

localized nature, generally follow the changes in CO s force constants with conformation<br />

[33]. These in turn are determined by the changes in CO bond length (the<br />

relationship is essentially linear [33]), which reflect the changing bonded and nonbonded<br />

interactions in the molecule with conformation. The larger variation in CO<br />

s2 (41 cm-') than in CO sl (16 cni-'1 is probably a result of its more variable conformational<br />

environinent adjacent to C", and therefore the varying (even if small)<br />

contributions of other coordinates. It is interesting that in a' the frequency order is<br />

reversed compared with the other conformers.<br />

The aiiiide I1 modes are less sensitive to conformation: the variation in either<br />

mode is of the order of 10 cm-', although the splittings do vary. As with amide I,<br />

the different frequencies of the two groups are mainly a reflection of their nonidentical<br />

chemical environments.<br />

The latter fact is demonstrated dramatically by the aniide I11 mode: the eigenvectors<br />

for the two peptide groups are significantly different, that of the higher fre-<br />

quency mode containing no PED contribution (210) of H" bend (b) although this<br />

coordinate is a major contributor to the lower frequency mode. Nor is NH ib a<br />

major contributor, as it was in NMA (in fact, it does not appear in the p2 modes).<br />

The large variations in frequency separation between the two amide I11 modes<br />

reflect the effect of conformation through H" b, since if H" is replaced by D" this<br />

variation disappears (the bands are now at 1294 & 1 and 1266 k 6 cin-'j [33]. This<br />

agrees with previous observations [5] that the amide I11 mode in polypeptides contains<br />

a significant conformation-dependent contribution from H" b. The interaction<br />

between H" b and NH ib is illustrated by the fact that in an (ND)? Ala dipeptide<br />

H" b occurs at 1300 and 1296 (p2), 1320 and 1272 ( a~),<br />

1305 and 1283 ( a~), and<br />

1324 and 1284 (a') cm-' [33]. The H" b coordinate also mixes with CN s to predict<br />

significant IR bands at 1277 cm-' in p2 and 1291 cm-' in a' [33]. Thus, with its<br />

added sensitivity to side-chain composition [5], amide I11 in polypeptides is no<br />

longer a simple characteristic NH ib, CN s mode.<br />

The amide V mode, which appears as a single band near 400 cn-' in NMA (see<br />

Table 5-l ), is split in the nonhydrogen-bonded conformers into at least two bands<br />

with predominant NH ob, CN t contributions. (These paired coordinates can make<br />

significant contributions to other modes, and each contributes individually to still<br />

others.) It might be thought that this splitting (and higher average frequency) is<br />

a result of the arbitrary scale factors that were used [33]. However, 4-31G" calculations<br />

with scale factors transferred from NMA [59] give similar results [73]. Thus,<br />

we must conclude that kinematic as well as potential coupling between the two<br />

peptide groups is responsible for the nonlocalized contribution of these coupled<br />

coordinates.<br />

The presence of hydrogen bonding obviously introduces a new element into the<br />

modes of the Ala dipeptide. This can be examined from the results on the three<br />

stable hydrogen-bonded conformers, C5, C7E, and C7A (Figure 5-1) [34]. In Table<br />

5-3 we show their calculated amide I, 11, 111, and V frequencies and PEDs.<br />

The relative amide I mode frequencies again generally follow the relative values<br />

of the CO s force constants [34], although mixing of the two CO s displacements can


5.3 Amidc Modes 255<br />

Figure 5-1. Optimized confonxatioiis<br />

of Ala dipeptide<br />

(p, ti/ in parentheses). Peptide<br />

groups-1: C6N7, 7: C12N17.<br />

Hydrogen bonds are indicated by<br />

broken lines. (From [33])<br />

alter this: for C5, even though the CO sl force constant (1 1.040) is larger than the<br />

CO s2 force constant (10.922), the predominantly CO sl mode is at a lower frequency<br />

(1 668 cm-') than CO s2 (1 698 cm-I). This is also counterintuitive in terms<br />

of the (admittedly weak) hydrogen bond to C02, whereas the relative frequencies<br />

are in the expected order for C7E and C7A. In distinction to the nonhydrogenbonded<br />

structures, NH ib is now a significant component of the amide I modes of<br />

the hydrogen-bonded structures (C7E and C7A). (To a lesser extent, this is also true<br />

for the Ala dipeptide in a crystal conformation similar to that of C5 but in which<br />

both peptide groups are fully hydrogen bonded [74].)<br />

The amide I1 mode frequencies generally reflect the influence of hydrogen bonding,<br />

which manifests itself primarily through the magnitudes of such force constants<br />

as NH ib and CN s. (It should be noted that explicit hydrogen bond coordinates are<br />

not included in the coordinate basis [33, 341, so the force constants implicitly reflect<br />

the influence of this interaction.) As expected, the higher frequency modes are<br />

associated with the peptide group having a hydrogen-bonded NH. In addition, the<br />

modes in C5 involve coupled NH ib displacements [34], which makes it difficult in<br />

general to infer the explicit dependence of amide I1 on conformation.<br />

The amide I11 modes demonstrate an increasing complexity compared with the


~<br />

Table 5-3. Some aniide modes of three hydrogen-bonded conformers of the alanine dipeptide".<br />

N<br />

ul<br />

Mode<br />

Conformer"<br />

I 1698<br />

CO s2(47)"<br />

CO sl(34j<br />

I1 1557<br />

NH ib2(49)<br />

CN s2(28j<br />

111 1276<br />

CN sl(23)<br />

NH ibl(20)<br />

CN s2(14)<br />

CO ibl(l1)<br />

c5<br />

1668<br />

CO sl(52)<br />

CO s2(32)<br />

1525<br />

NH ibl(49)<br />

CN sl(251<br />

NH ib2(12)<br />

1229<br />

H" b2(40)<br />

CN s2( 14)<br />

C7E<br />

1704 1676<br />

CO ~2164) CO sl(70)<br />

NH ib2(26) NH ibl(15)<br />

C7 A<br />

3<br />

1699 1682 2-<br />

CO s2(51) CO sl(55) E.<br />

NH ib2(231 CO ~2119) CO sl(14) NH ibl(l51 2<br />

$<br />

1587 1534<br />

1594 1551<br />

P"<br />

c1<br />

NH ib2(59) NH ibl(60) NH ib2(59) NH ibl(601 Y<br />

CN s2(27) CN sl(28) CN s2(28) CN sl(27)<br />

2<br />

c,<br />

co s2(21) co sl(14) CO s2( 18) CO sl(14)<br />

0<br />

?$<br />

1279 1258 1335 1325 1292<br />

H" bl(28) CN s2(31) H" b2(30) H' bl(26) CN ~1137)<br />

CN ~1123) H" b2120) CN s2( 13) CN s2(19j CO ib(16)<br />

NH ibl(16) H" b1116) H" bl(12) H' b2(18) MC s(12)<br />

NM s(13) C"C s(13j 1<br />

bl<br />

%<br />

2<br />

c<br />

V 693<br />

NH ob1132)<br />

CN tl(23)<br />

CO ib21 15 i<br />

NC" t( 14)<br />

683<br />

NH obl(39)<br />

CN tl(35)<br />

NC" t(24)<br />

CO ob2(14)<br />

585 754 577 809 58 1<br />

NH ob2(111) NH ob2(51) NH obl(102) NH ob2(51) NH oblill6)<br />

CN t2( 57) CN t2(40) CN tl(76) CN t2( 19) NC' t(951<br />

NC" t(67)<br />

CN tl(8ll<br />

r;<br />

a From [34].<br />

q, $ 4 5: -166", 167": C7E: -85", 73'; C7A: 75", -62".<br />

s: stretch, ib: in-plane bend, b: bend, ob: out-of-plane bend, t: torsion, d: deformation, M: methyl C. Group coordinates defined in [34]. Contributions<br />

2 10.


5.4 Polypeptides 257<br />

nonhydrogen-bonded conformers. While NH ib contributes (at the level of 2 10)<br />

to the higher frequency C5 and C7E modes, it is absent from the C7A modes.<br />

Furthermore, if we ascribe amide I11 character to H" b, CN s modes with minimal<br />

NH ib, then C7A has such modes at frequencies well over 1300 cnrl, again<br />

reflecting the conformation-dependent contribution of H" b. (The situation is<br />

somewhat reversed in the crystalline Ala dipeptide [74], with a calculated mode at<br />

1325 cni-' showing traditional CN s, NH ib character and modes at 1272 and 1258<br />

cm-' having H' b, CN s character.) Since NH ib can contribute throughout the<br />

-1400-1200 cn1-l region, and can mix variably with CN s and Ha bend, it is not<br />

surprising that the amide I11 modes are a complex function of conformation.<br />

The amide V modes are dominated by the effects of hydrogen bonding. Thus, the<br />

free amide V mode is at -580 cni-' in all three conformers, while the bonded<br />

modes reflect the increasing hydrogen bond strengths: -690 cm-l in C5 (the splitting<br />

probably due to an accidental resonance with another mode), 754 cm-' in C7E,<br />

and 809 cni-' in C7A. Hydrogen bonding thus seeins to mask conforinational<br />

differences in amide V.<br />

These studies on NMA and the Ala dipeptide illustrate that, although it is useful<br />

to characterize peptide spectra in terms of supposedly localized amide modes, it is<br />

the subtle differences in the detailed nature of these vibrations with varying conformation<br />

and hydrogen bonding that provide the spectral clues to differences in<br />

structure. This is demonstrated by analyses of tripeptides [28, 75-77], which also<br />

involve interactions between two peptide groups. Although done with an empirical<br />

force field [5], these analyses show the sensitivity of vibrational spectra to specific<br />

structural features. This emphasizes the importance of highly accurate force fields<br />

in enabling vibrational analysis to realize its potential for being a powerful tool to<br />

study the structures and forces in macromolecules.<br />

5.4 Polypeptides<br />

As will be obvious from our previous discussion, a comprehensive understanding of<br />

the vibrational spectra of polypeptides depends on having reliable normal mode<br />

analyses to combine with appropriate experimental studies. Such analyses in turn<br />

rest on the quality of the force field. At the present time, an SDFF for the polypeptide<br />

chain is not available, the most detailed normal mode analyses being based<br />

on extensive empirical force fields [5]. These force fields have been tested on tripeptides<br />

of known structure with excellent predictive results [28, 75-77], and have<br />

proven to be very satisfactory for polypeptide systems. For reasons given in Section<br />

5.2.1, however, we must be careful about a final acceptance of all of their detailed<br />

predictions. Acknowledging these reservations, the general features of polypeptide<br />

spectra provided by the present detailed empirical force fields can nevertheless be<br />

accepted with confidence.<br />

In this section, we describe results of some of the vibrational analyses that<br />

have been done on the main secondary conformations found in polypeptides, viz.,


258 5 Vibrationul <strong>Spect</strong>roscopy oj' Polypeptides<br />

extended-chain and helical structures. Not all known structures have been the subjects<br />

of such analyses, but this approach will certainly be extended to them in the<br />

future.<br />

5.4.1 Extended-Chain Polypeptide Structures<br />

5.4.1.1 General Features<br />

Stereochemically acceptable extended-chain (so-called p-sheet) structures were first<br />

described by Pauling and Corey from molecular model building (78, 791. In some<br />

cases, specific structures of synthetic polypeptides have been determined from X-ray<br />

diffraction studies. These have provided the bases for the vibrational analyses of this<br />

class of polypeptide chain conformations.<br />

In these analyses, some structural constraints have had to be imposed in order to<br />

achieve a reasonable degree of transferability in the force fields. These constraints<br />

have consisted of adopting a standard peptide group geometry, which is probably<br />

not a serious problem, and accepting a completely planar peptide group (i.e., the<br />

CN torsion angle, Q, equal to 180"), which may be more serious since departures<br />

from planarity of 5-10' require relatively little energy (and are not uncoininon in<br />

proteins) and may have an important effect on some of the modes. The standard<br />

geometry of the peptide group is given in Table 5-4, based [XO] on the refined X-ray<br />

structure of P-PLA [Sl]. (The dimensions for PGI are slightly different [19].) We<br />

therefore assume that chain conforinations differ only in their q, $ torsion angles.<br />

As Pauling and Corey first pointed out [78, 791, adjacent hydrogen-bonded chains<br />

in a P-sheet can occur in two arrangements, parallel and antiparallel (with respect to<br />

the direction of the chemical sequence along the chain). In addition, depending on<br />

the axial stagger between adjacent chains, the sheet can be 'pleated' or 'rippled', the<br />

latter being possible only if all-L and all-D chains alternate in the sheet. (In practice,<br />

only PG can satisfy this requirement, since it has not as yet been possible to<br />

Table 5-4. Standard geometry of the peptide group<br />

Bond length<br />

('4")<br />

Bond angle<br />

(degrees)<br />

C"C 1.53 C"CN 115.4<br />

co 1.24 C"C0 121.0<br />

CN 1.32 CNC" 120.9<br />

NH 1 .oo CNH 123.0<br />

NC ' 1.47 NCT<br />

1<br />

C"H 1.07 NC"H<br />

109.5<br />

CC"H<br />

OCNH 180.0<br />

a From 180, 811.


5.4 Polvpeptida 259<br />

Figure 5-2. Antiparallelchain<br />

pleated sheet of<br />

polyi L-alanine). The CH;<br />

group is represented by<br />

the largest circle. (From<br />

[51)<br />

Figure 5-3. Antiparallelchain<br />

rippled sheet of<br />

polyglycine. (From [5])<br />

produce such a regular alternation with other side-chain-residue polypeptides.)<br />

Thus, in principle we can have antiparallel-chain pleated sheet (APPS) (Figure 5-2),<br />

antiparallel-chain rippled sheet (APRS) (Figure 5-3), parallel-chain pleated sheet<br />

(PPS), and parallel-chain rippled sheet (PRS) extended-chain structures. Their<br />

structural parameters are given in Table 5-5 [82].<br />

In describing these structures, and in the vibrational analyses, we assume the<br />

sheets to be infinite in two dimensions and therefore calculate only the in-phase unit<br />

cell modes. Although modes of only a single sheet are treated, the TDC interactions<br />

have been evaluated over a number of sheets (7-11). The effect of finiteness of a<br />

single sheet on just the TDC interactions for the amide I mode has been examined<br />

[83, 841.<br />

Extended-chain regions are a major component of protein structures. They can<br />

occui- in mixed antiparallel and parallel arrangements even within a sheet, the sheets<br />

can be twisted, and complex ‘barrel’ arrangements are also found [2]. It is therefore


Table 5-5. Structural parameters of ,&sheet StrUCtUreS".<br />

Paraiiieterh APPS' APRSd PPSC PRS'<br />

4.73<br />

3.445<br />

138.4<br />

135.7<br />

2.73<br />

1.75<br />

9.8<br />

164.6<br />

4.77<br />

3.522<br />

149.9<br />

146.5<br />

2.91<br />

2.12<br />

31.4<br />

134.4<br />

4.85<br />

3.25<br />

119.0<br />

114.0<br />

2.74<br />

1.82<br />

5.5<br />

172.0<br />

4.80<br />

3.25<br />

- 119.0<br />

114.0<br />

2.82<br />

1.75<br />

4.9<br />

172.0<br />

" From [82].<br />

Lengths in A. angles in degrees. a (or a/2): lateral separation between adjacent chains. 17: axial<br />

separation between adjacent units in one chain.<br />

Antiparallel-chain pleated sheet of poly(L-alanine). From [80, 81 1.<br />

Antiparallel-chain rippled sheet of polyglycine. From [82].<br />

Parallel-chain pleated sheet of polyil-alanine). From [83].<br />

Parallel-chain rippled sheet of polyglycine. From [83].<br />

of major importance to understand the vibrational dynamics of this type of polypeptide<br />

chain conformation.<br />

5.4.1.2 Antiparallel-Chain Structures<br />

Antiparallel-Chain Pleated Sheet<br />

The APPS structure is the predominant one in synthetic polypeptides, and is a<br />

prevalent motif in proteins [2]. (Polyglycine is an exception, which xe treat below.)<br />

An X-ray diffraction study of the simplest representative polypeptide, /I-PLA [811,<br />

has provided the geometric parameters of the APPS. In addition to those given<br />

in Table 5-5, we note that the axial shift between adjacent chains, measured from<br />

the position where the hydrogen bond is linear, is 0.27A (as obtained from a TDC<br />

analysis of the amide I mode [80]). The vibrational analysis of /I-PLA thus provides<br />

the basic spectral characteristics of the APPS.<br />

A PLA sample can be easily oriented, and dichroic IR spectra are therefore<br />

readily obtainable. Such spectra [85], together with far-IR spectra [86], are shown in<br />

Figure 5-4. Together with spectra of the ND inolccule [87, 881, as well as Rainan<br />

spectra [89] (Figure 5-5), a large amount of spectral information is available, both<br />

to optimize an empirical force field and to provide a reasonable description of the<br />

normal modes.<br />

The symmetry of the p-PLA structure determines that the normal modes will<br />

be distributed among four symmetry species with optical activity and IR dichroism<br />

as follows [SO]: A-30 modes, Raman; B1-29 modes, Raman, IR (11); B2-29 modes,<br />

Raman, IR (I); B3-29 modes, Raman, IR (I). The results of the most recent nor-


I<br />

5.4 Polypeptides 261<br />

)I<br />

.- +<br />

$ 800 1000 1200 1400 1500 1600 1700<br />

0<br />

1 1 1<br />

2800 2900 3000 3100 3200 3300 3400<br />

Figure 5-4. Infrared spectra of /l-poly(Lalanine).<br />

Upper two panels: mid-IR<br />

region. (-) Electric vector perpendicular<br />

to the stretching direction. (- - -)<br />

Electric vector parallel to the Stretching<br />

direction. (From [SS]) Lower panel:<br />

far-IR region. (From [86])<br />

Frequency (ern-')<br />

2<br />

0) 1<br />

Figure 5-5. Ranian spectrum of /l-poly(L<br />

alanine). (From [SS])<br />

900 800 700 600 500 400 300 200 100<br />

FREQUENCY (cm-'<br />

ma1 mode analysis [22] are given in Table 5-6 (the NH s modes, which are influenced<br />

significantly by Fernii resonances 1901, and the CH s modes are discussed in<br />

an earlier paper [SO]).<br />

There are some general points to note about this normal mode analysis. First, the


Table 5-6. Observed and calculated frequencies (in cm-' ) of APPS poly(L-alanine).<br />

Observed" Calculated Potential energy distributionh<br />

Raman IR A Bi B2 Bi<br />

1669s<br />

1553VW<br />

1538W<br />

1451s<br />

1451s<br />

1399W<br />

1368W<br />

1335W<br />

1311W<br />

12433<br />

1694W 1 1<br />

1632VS I<br />

1555MW I<br />

15248 1)<br />

1454s / /<br />

1446s i<br />

1454s 11<br />

1446s _L<br />

1402MW 11<br />

1386W i<br />

1372MW / /<br />

1330W br I<br />

1309 sh<br />

1670<br />

1539<br />

1455<br />

1452<br />

1402<br />

1372<br />

c<br />

1236<br />

1695<br />

1528<br />

1455<br />

1452<br />

1399<br />

1372<br />

1630<br />

1562<br />

1452<br />

1451<br />

1383<br />

1317<br />

1299<br />

1698<br />

1592<br />

1453<br />

1451<br />

1385<br />

1337<br />

1305<br />

CO s(78) CN sf141<br />

CO sf76) CN si19)<br />

CO sf73) CN s(21 I<br />

CO sf70) CN ~(21)<br />

NH ib(571 CN sf211 '2°C s[l01<br />

NH ibi53) CN sf17) C"C ~(14)<br />

NH ibi48j CN sf22) CO ibi121 C*C $11)<br />

NH ibi41j CN si26) CO ib~141 C"C s(13)<br />

CH3 abl(34) CH7 ab2(39)<br />

CH3 ab1146) CH7 ab2137)<br />

CH-, ablf88) CH? rlfl0)<br />

CH; ab1186) CHI rlfl0)<br />

CH3 ab2145) CHi dblf43)<br />

CH3 ab2(48) CH? ablf41)<br />

CHI ab2i861 CH7 12(10)<br />

CH? ab2(89) CH7 12j10)<br />

H" b2(32) CH7 sb(17) C*C sfl4) NH lbi13)<br />

Ha b2f33) CHI sb120) NH ib(l5) C'C sill)<br />

CH7 sb(70i H' blil6I CYCD S! 111<br />

CH3 ~b174) H" bl (16, C^Ci'~i10)<br />

CH7 sbi64) H'b? (161<br />

CH1 sbi6lI H'b2 119)<br />

H" b2132) NH lb(25) C"C ~(12) CH3 sbfll)<br />

H" b2164) NH Ibf 15)<br />

H" b2130) CN ~(181 CO ibfl9<br />

NH M23) CO ibfl8) C'C si14) CN sf131<br />

H' b2134) NH ibil9) NC' ~(19) CN sil3i


1226M 1222s (1<br />

1165W<br />

1167s (1<br />

1120VW<br />

1092s<br />

10S4W 1<br />

1052M I<br />

1065M<br />

967M<br />

909VS<br />

837VW<br />

775M<br />

698VW<br />

966s 11<br />

925M I<br />

902 W 11<br />

778MW<br />

706SI / /<br />

1198<br />

1162<br />

1093<br />

c<br />

1055<br />

969<br />

913<br />

{<br />

704<br />

668<br />

1231<br />

1195<br />

1161<br />

1092<br />

1054<br />

970<br />

706<br />

662<br />

1195<br />

1125<br />

1085<br />

1064<br />

918<br />

844<br />

775<br />

708<br />

1196<br />

112s<br />

1086<br />

1065<br />

913<br />

846<br />

755<br />

705<br />

H' b2(?8j NC' ~(241 NH lb(l8) CN s( I I I<br />

NC" s(361 CH; rl(13) C'C $11) NH ib(10)<br />

C'CB s(331 H" bl(27) NC' s(22) CH3 sb(l1) CH; rl(10)<br />

NC" s(30) NH ib(13j C"C 412) H' b2(ll) CH3 rlill)<br />

C"CP s(33) H' bl(261 NC' ~(221 CH3 rl(10 CH; sbt 10,<br />

Hi bl(56) CH3 sb(19) C'CP ~(13)<br />

H" bl(541 CH3 sb(18j C"C1 ~(131 CN ~(10)<br />

CH3 r2(27] Ha bl(26) CH3 Sb(11)<br />

H" b1176) CHI 12/76) CHI sbflli<br />

CH3 12(55\ C"C1' ~(21)<br />

CH3 r2(541 C"CP $2)<br />

CH3 rl(271 CH3 r2(251 C'CD ~(101 CN s(l0)<br />

CH? rl(26) CHI r2126j C"Cp dl31 CN dl01<br />

C'C/'s(23! HI blL2l) CH3 rl(18) CH3 12(17)<br />

C"Cp s(22) H" bl(21) CH3 r2(19) CHI rl(17)<br />

C"Cp ~(501 CH3 rl(141 H" bl(l1)<br />

C"Cp s(50) CHI i-1113) HY blillj CHI i-2110,<br />

CH3 rl(50) NC' ~(251<br />

CH3 rl(501 NC" s(2Y<br />

NC" s(27) CH3 rl(23) CN s(15)<br />

C"C s(15) CN s(14) CH; r2(14) CNCd d(13j CO s(l1)<br />

NC" ~(30) CH? 11f23) CN ~(l4!<br />

'2°C s(15) CN s(14) CHI r2(14) CNC' d(13) CO s(12)<br />

C"C s(34) NC" s(13) CN s(12) CTp ~(12)<br />

C"C s(31) NC' s(19j C"CB s(14j<br />

CO lb(19) C'C s(l5) NC's(131 CI'b2i13j NC'Cdi12)<br />

CNC" 411)<br />

CO ib(181 C"C s(16) CB b2(14) NC' $12) NC"C d(l1)<br />

-+<br />

CNC" d( 11) 1<br />

CN t(74) NH ob(2Sj NH<br />

3<br />

0 lb(23)<br />

P<br />

CN t(41) NH ob(41) NH 0 lb(19~ CO ob1181 H" bl(l0)<br />

CN t(74) NH ob(28) NH 0 ib(72)<br />

CN t(48) NH ob(41) NH . 0 Ib(l9) CO ob(131 H" bl(10) N<br />

m<br />

w<br />

CO ob(50i CN ti351 Cb bliloj<br />

CO ob(47r CN t(41) CD bl(10)<br />

b<br />

a<br />

2<br />

e


Table 5-6. (coiztimied)<br />

Observed" Calculated Potential energy distributionh<br />

Raman IR A Bi B? B3<br />

628VW 624<br />

629W 11<br />

437w<br />

332W<br />

300M<br />

266VWsh<br />

615WI<br />

448M I<br />

432M )I<br />

326W //<br />

440<br />

327<br />

i<br />

27 1<br />

235Msh 240<br />

240M / /<br />

18SVW 177<br />

62 1<br />

440<br />

328<br />

279<br />

240<br />

211<br />

151<br />

626<br />

591<br />

447<br />

286<br />

286<br />

252<br />

238<br />

CO ib(46) (2°C si16) CO obil3)<br />

CO ib(49) C"C s(16) CO ob(12)<br />

CTN d(38) CO ob(l8) NCYC di16) NH ob(l6) CB b2(15)<br />

622 CTN d(31) CO ob(30) NH ob(22) Cb b2(161 NCT d(l4)<br />

592 CO ob(52) C"CN d(19)<br />

CO ob(62) C"CNd(l1) NH ob(1l) Ha b2(101<br />

CB bl(55) NH ob(15)<br />

447 Cp b1154) NH obil3)<br />

CP b2(75) NC"C d(13)<br />

CB b2174) NC"C d( 12)<br />

NC'C d(25) CB b2(11) CO ob(l1) CB bl(10)<br />

NC"Cd(24) Cp b2(11) COob(l1) CB blilO)<br />

CO ibf33) NC"C d(19) CNC" d(15j CB b2(15) CB bl(13)<br />

CO ob(l1j<br />

CO lb(33j NC"C d(19) CB b2(15) CNC" d(14) CB bl(13)<br />

CO ob( 10)<br />

CTN d(39) CaCP t(26r NCT d(1li<br />

C"CN d(38) C"CB t(25) NC"C d(17)<br />

CB b2( SO)<br />

252 Cp b2( 50)<br />

C'CP t(671 NC"C d(11)<br />

C'C'' t(49) NC"C d(16)<br />

C"C'7 t(91)<br />

238 C'CB t(90)<br />

CNC" di35) C'CD t123) H. 0 ~ (23)<br />

CNC' d(621 C"CN d(15)<br />

156 H 0 s(58) NC'C dl 101<br />

NH ob(79) CO obi131 NC" dl3) H" b2110j


135s<br />

9 1 Ms1i<br />

122Wbr<br />

147 NH ob(37) CO ob(16) NC' s(13) H" b2(101<br />

118<br />

CNC" d(21) NC" C d(18) CTN d(13) NH ob(l?i<br />

104<br />

NH...O ib(27) CN t(21) Cp b1115~<br />

103 NH.,.O ibi331 CN t(21) NH obi2O) C"C t(15) CO,..H t(l1)<br />

91<br />

Cp bl(26) NH ob( 19) H-.O s( 14) C"C t( 11)<br />

87 H...O ~(33) CNC" d(17) CN t(l1) CN'C d(10)<br />

74<br />

C'C t(24) NC" ti22)H.,.O ~(22) CN t(15)<br />

41 CO...H ib(21) NH ob(70) Cpblilhj NH...O ib(l5)<br />

NH...O ti 12)<br />

32<br />

NH.,.O ib(371 CO-.H ib(22) CN t(l5)<br />

31<br />

H...O s(15) CO...H t(l5) CO...H ib(14) NH...O ib(l4)<br />

NH...O t( 12) NH ob( 11 j CN t( 10)<br />

19<br />

CO,.,H ti35) NH...O t(27) CO.,.H ib(l4) H". . .H' sil3j<br />

a Data taken from [88]. S: strong, M: medium, W: weak, V: very, sh: shoulder, br: broad, 11: parallel dichroism, I: perpendicular dichroism.<br />

b s: stretch, as: antisymmetric stretch, ss: symmetric stretch, b: bend, ib: in-plane bend, ob: out-of-plane bend, ab: antisymmetric bend, sb: symmetric<br />

bend, r: rock, d: deforniation, t: torsion. Contributions 2 10.


966 5 Vibr-atioiial Slm-troscopj, of' Pol.vpeptides<br />

empirical force field is capable of giving very acceptable frequency agreement:<br />

the rms error between observed and calculated frequencies is 5.9 cm-'. Second, the<br />

character of the symmetry species is well predicted: A species modes exhibit only<br />

Rainan activity, B1 and B? species modes exhibit 1) and I IR dichroism. respectively.<br />

Third, while the agreement is somewhat poorer, band shifts on N-deuteration<br />

are reasonably well accounted for [88] (for amide I the observed downshift of<br />

4 cn-' is reproduced exactly by the calculation). Finally, TDC is crucial to<br />

explaining the large splittings in the amide I and amide I1 modes: in its absence the<br />

ainide I modes are calculated at 1670 (A), 1673 (B,), 1665 (B2), and 1670 (B3j cin-.',<br />

and the amide I1 modes are predicted at 1545 (A), 1549 (BI), 1550 (B?), and 1554<br />

(B3) c111-l [5].<br />

With respect to the amide modes, a comparison with the results of the ah initio<br />

calculation on the hydrogen-bonded Ala dipeptide is of interest, even though the<br />

force fields are different. Although localized primarily in CO s, the amide I modes<br />

of /I-PLA are constrained by symmetry to be coupled modes of the four peptide<br />

groups in the unit cell. Interestingly, CN s is now the second major component<br />

compared with NH ib foi- the Ala dipeptide. The coupled amide I1 modes contain,<br />

in addition to NH ib and CN s, C"C s and CO ib contributions not present (at the<br />

level of 2 10) in the Ala dipeptide. With respect to the strong amide 111 modes, at<br />

1243 and 1224 cni-', H" b dominates and NC" s also contributes in addition to NH<br />

ib and CN s. As noted above, NH ib is a contributor to many other modes in<br />

the 1400-1200 cm-l region. The amide V modes are well accounted for, and in<br />

particular the observed absence of any significant dichroism in this band 1911 is<br />

explained by the near coincidence in frequency of the predicted B1 and B3 species<br />

modes. Besides the aniide modes, local as well as other skeletal modes are very well<br />

accounted for.<br />

The excellent agreement obtained for /I-PLA would indicate that this force field<br />

should be transferable to other APPS polypeptides, of course taking due account of<br />

side-chain differences. This has been done for Ca-poly( L-glutamate), Ca-PLG (921,<br />

which X-ray and electron diffraction data had indicated to have an APPS structure<br />

but which were not extensive enough to provide a definitive conclusion. A proposed<br />

model [93] was used as the basis for the vibrational analysis, and the good prediction<br />

of the observed bands [92] supports both the model as well as the viability<br />

of the force field. The subtle differences between the P-PLA and /I-Ca-PLG spectra<br />

are accounted for by the normal mode calculation, and emphasize the influence<br />

of the side chain on the main-chain modes, particularly on amide 111. A similar<br />

successful vibrational analysis has been done on an alternating copolymer, APPS<br />

poly( L-alanylglycine) [ 191.<br />

Antiparallel-Chain Rippled Sheet<br />

Although early studies assumed that PG also adopts an APPS structure, electron<br />

diffraction studies of 'single crystals' and oriented thin films [94] suggested that<br />

extended-chain PG, PGI, has an APRS structure. This was supported by conforniational<br />

energy calculations [95], and was the basis for a detailed vibrational<br />

analysis of the PGI spectra [19, 961, strongly confirming this proposal. (Since the


5.4 Polypeptides 267<br />

glycine residue is achiral, it can adopt the D or L configuration required of alternate<br />

chains in the APRS.) This was aided by extensive IR spectra of isotopic derivatives<br />

(NH, CD2; and ND, CD2) [97] and by Raman spectra [9S], which were important<br />

because macroscopic samples of oriented PGI could not be prepared.<br />

The conclusion from vibrational analysis that PGI has an APRS structure was<br />

based on a detailed TDC study of the aniide I mode splittings for the APPS as<br />

compared with the APRS as a function of the axial shift between adjacent chain<br />

[19]. No reasonable shift values for the APPS were in agreement with experiment,<br />

whereas the APRS gave excellent agreement for a shift of 0 A, exactly that indicated<br />

by the electron diffraction analysis [94]. This structure (Figure 5-3) has a<br />

center of inversion and therefore the mutual exclusion rule applies, giving the<br />

following symmetry species and activity [ 191: As-21 modes, Raman; A,-20 modes,<br />

IR( 11); Bg-21 modes, Rainan; Bu-19 modes, IR(I). The results of the normal mode<br />

analysis [96] give an overall rnis error of 6.1 cm-', comparable with that for P-PLA.<br />

(The force field for p-PLA was in fact refined from that for PGI [80].) A description<br />

of the amide modes of PGI is given in Table 5-7.<br />

The splitting of the amide I modes, somewhat smaller than for the APPS structure,<br />

is again due to TDC: in its absence the calculated frequencies are 1684 (Ag),<br />

1677 (Au), 1676 (Bg), and 1674 (B,) cm-I. Compared with APPS PLA, CTN d<br />

now makes a noticeable contribution to the eigenvector.<br />

The amide I1 modes, which are probably at a lower frequency because of the<br />

weaker hydrogen bonds, are also split by the TDC interactions: in their absence<br />

the calculated modes are at 1534 (Ag), 1535 (A"), 1559 (Bg), and 1559 (B,]) cni-'.<br />

The absence of clearly observed high-frequency modes, particularly an IR band<br />

near the expected frequency of 1572 cm-I, raises a very important point: it is necessary<br />

to be able to predict intensities as well as frequencies. Intensities require<br />

knowledge of dipole derivatives (Eq. (5-4)), but since they also depend on correct<br />

eigenvectors (Eq. (5-5)) they provide an independent test of the force field. Such an<br />

1R intensity calculation has been done on PGI [20], and the results show that the B,<br />

amide I1 mode is indeed predicted to be very weak, in agreement with observation.<br />

The amide I11 mode region is quite complex, with mainly NH ib, CN s contributing<br />

near 1300 and 1200 ciii-', where only weak bands are observed. In the<br />

'expected' region (1300-1200 cm-'), the strong bands are due to CH? twist (tw)-<br />

dominated vibrations. The NH ib contributions now extend to bands about 1400<br />

and below 1200 cm-' .<br />

The amide V mode is at a similar frequency to that in ,!I-PLA, even though the<br />

PGI hydrogen bond is weaker. This may be because of the large CN t contribution<br />

and/or the somewhat different nature of the modes.<br />

5.4.1.3 Parallel-Chain Structures<br />

Parallel-Chain Pleated Sheet<br />

Although there are no known PPS polypeptides, this general structure is found<br />

in proteins [2] and in some tripeptides. in two of which it has been the subject of<br />

detailed vibrational analyses [75, 771. The excellent prediction of observed I R and


Table 5-7. Amide mode frequencies (in cm-') of APRS polyglycine I.<br />

Observeda Calculated Potential energy distributionb<br />

Raman IR A, B" B, B u<br />

1695 CO s(77) CN si15) CTN d(l1)<br />

1685M 1689 CO 475) CN si2Oi C'CN d( 11)<br />

1674s 1677 CO s(74) CNs(211 C"CNd(ll1<br />

1636s 1643 CO s(69) CN si22) C"CN d(ll'1<br />

1602<br />

NH ib(56) CN s(19i C T ~(12)<br />

1572 NH ib(51) C"C s(16) CN s!14!<br />

1517s 1515 NH ibi35) CN s(28i C T s(171 CO ib(14)<br />

1515W 1514 NH ib(35) CN s(271 C'C s(17) CO ibil4)<br />

1410M 1415 CH? w(44) CH? b(3l) NH ib(l4)<br />

1408W 1415 CH? w(40) CH? b(33) NH ib(13)<br />

1304<br />

NH ib(30) CO ib(19) CN si18j C"C sil6)<br />

1295W 1286 NH ibi39) (2°C s(17) CO ib(1hi CN ~(12)<br />

1220w 1213 NC" s(29) NH ib(23) CH2 w(18) CH2 tw(l6) CN s(13j<br />

1214W 1212. NC' s(29) NH ib(23) CH: w(18) CH2 tw(l5j CN ~(13)<br />

1162M 1153 NC" s(50) C"C ~(13) NH ib(l2.)<br />

1152<br />

736<br />

NC' s(50) C'C s(14) NH ib(l2.i<br />

CN t(63) NH...O ib(15) NH ob(l1) H...O s(l1)<br />

708s<br />

718 CN t(75) NH...O ib(19) NH ob(16) H...O s(l0)<br />

{<br />

718<br />

CN t(791 NH ob(36) NH...O ib(23) H...O s(10)<br />

701<br />

CN t(79) NH ob(29) NH...O ib(25r H...O ~(15)<br />

Data taken from [95]. S: strong, M: medium, W: weak.<br />

s: stretch, b: bend, ib: in-plane bend, oh: out-of-plane bend, d: deformation, w: wag, tw: twist. t: torsion. Contributions 2 10.


5.4 Polypeptides 269<br />

Raman bands in these cases, using the same empirical force field as for APPS PLA<br />

[5], led to the belief that it would be useful to obtain the frequencies of PPS PLA in<br />

order to characterize the vibrational spectrum of this structure. The unperturbed<br />

normal modes have been obtained, and the influence of TDC interactions on the<br />

aniide I and ainide I1 modes, as determined by the intersbeet separation, has been<br />

assessed [82]. While unperturbed aniide I modes occur at 1678 and 1675 cm-I, TDC<br />

interactions within one sheet shift these to 1663 and 1642 cm-' and a multisheet<br />

array ( 11 x 11 x 11 residues) with a separation of 5.3 A (comparable with that in<br />

APPS PLA) leads to modes at 1669 and 1677 cm-'. A similar effect is predicted for<br />

ainide 11: single sheet modes are at 1553 and 1586 c~ii-', with the above niultislieet<br />

separation giving 1549 and 1583 cin-' .<br />

Parallel-Chain Rippled Sheet<br />

A PRS structure is only possible for PG, and is ruled out by the evidence for an<br />

APRS structure for this polypeptide. However, the norinal, modes of PRS PG have<br />

also been calculated (821. It is interesting that, when multisheet TDC coupling is<br />

included, the strong Raman amide I mode is predicted to be at a much lower frequency<br />

than the strong IR amide I mode, just the opposite of what is observed and<br />

predicted for the APRS structure. This demonstrates the power of detailed vibrational<br />

spectroscopic analysis in revealing fine details of polypeptide structure.<br />

5.4.2 Helical Polypeptide Structures<br />

5.4.2.1 General Features<br />

The concept of a helical structure as the most general one for a long polymeric<br />

chain was developed quite early [99, 1001, and possible polypeptide helices having<br />

an integral number of residues per turn were described [loll. However, it was only<br />

when this constraint was relaxed and well-defined stereochemical criteria invoked<br />

[lo21 that a specific conforination, the or-helix, was discovered [102, 1031 that is in<br />

fact found in polypeptides and proteins [2]. A number of other helical conforniations<br />

have been found since then, and their structural parameters, are given in Table<br />

5-8. Vibrational analyses of some of these are described in this section.<br />

Polypeptides are found in both single-stranded intramolecularly hydrogenbonded<br />

helices as well as in conformations in which hydrogen bonds are fornied<br />

between helices. In the former case, the structure is conveniently described by<br />

several parameters: the number of residues per turn, n (positive if right-handed,<br />

negative if left-handed); the axial translation per residue, h; and the backbone torsion<br />

angles, p, $, and o. Sometimes the number of atoms in the 'ring' formed by the<br />

hydrogen bond, in, is specified [loll (and the residue connectivity via the hydrogen<br />

bond can be given; thus, 5 - 1 iiieans the NH group on residue 5 is bonded to the<br />

CO group on residue 1). In such a case, the helix will be designated PI,?, [loll.<br />

Among such helices. the a-helix (3.6213) structure has been obtained froin a<br />

careful analysis of X-ray fiber diffraction patterns of PLA [104]. As is general for<br />

helical structures [ 1051, this helix, designated, a ~, has a counterpart, designated all,


~~ ~<br />

270 5 Vibrational Sl~ectroscop~~ oj Polypeptides<br />

Table 5-8. Parameters of some observed helical polypeptide structures<br />

Structure<br />

hlrvrlnloleL 1rltrr<br />

~1-Hel1x<br />

q1-Helix<br />

3lo-Hel1~<br />

to-tieli~<br />

R-HKIIX<br />

n” nib hL Yd id cod H-bonde Ref.<br />

3.62 13 1.496 -57.4 -47.5 180 5-1 104<br />

3.62 13 1.496 -70.5 -35.8 180 5-1 106<br />

2.99 10 2.01 -45 -30 176 4- 1 111<br />

-4.00 13 1.325 62.8 54.9 175 541 1 I ?<br />

-4.25 16 1.17 51 74 177 6-1 115<br />

Intevniolec iiliw<br />

Polyglycine I1<br />

Polypioline 11<br />

Polypioline 1’<br />

-3.00 3.10 -76.9 145.3 180<br />

-3.00 3.10 -78.3 148.9 180<br />

3.33 1.90 -83.1 158.0 0<br />

122<br />

128<br />

135<br />

Number of residues/turn, + for right-handed and - for left-handed helices<br />

Number of atoms in hydrogen-bonded ‘ring’.<br />

Unit axial translation, in A.<br />

‘’ Dihedral angle, in degrees.<br />

Hydrogen bonding pattern, N-H-O=C.<br />

Cis peptide groups.<br />

(j<br />

b<br />

Figure 5-6. Structure of a-helix of poly(L-alanine). (a) al-helix; ib) q1-helix. Methyl groups are<br />

represented by large circles. (From 151)<br />

which, though having different p, $, has the same IZ and 11, and is comparably<br />

allowed energetically [ 1061 (and may be present in bacteriorhodopsin [ 1071). These<br />

structures are shown in Figure 5-6, where it can be seen that the plane of the peptide


5.4 Polypeptides 271<br />

Figure 5-7. Structure of 3lo-helix of poly(a-amiiioisobutyric acid).<br />

Methyl groups are represented by large circles. (From [S])<br />

group in a11 is inclined to the helix axis whereas it is almost parallel to the axis in MI.<br />

The a-helix conformation is not only coniinoii in polypeptides, and is a basic component<br />

of proteins [2], but its topology forins the basis for two-stranded coiled-coil<br />

structures such as tropomyosin [108]. The 3lo-helix is similar to the a-helix, but has<br />

a 4 i 1 hydrogen-bonding pattern. Although segments are found in proteins [2],<br />

the only polypeptide in which it has been found, as suggested by electron diffraction<br />

[ 1091 and supported by vibrational analysis [ 1101, is poly(a-aminoisobutyric acid),<br />

PAIB (Figure 5-7). Its structure, as indicated by conformational energy calculations<br />

[111], has a nonplanar peptide group. The o-helix is a left-handed 5 - 1<br />

helix, -4.013, proposed from X-ray analysis for the structure of poly(j?-benzyl-Laspartate)<br />

[ 1121. Coiiformational energy calculations have indicated somewhat different<br />

parameters [ 1131, with other calculatioiis suggesting a further variant [114].<br />

The n-helix, said to be present in poly(j?-phenethyl-L-aspartate [115], is also a lefthanded<br />

helix, -4.2516, although a right-handed form with a 4" distorted NC"C<br />

angle has been suggested [116]. The 0- and n-helices also have nonplanar peptide<br />

groups. Other helical conformations have been proposed, a y-helix [lo?, 1031 and a<br />

2.27-helix [117], but neither of these has as yet been observed in polypeptides. When<br />

L and D residues alternate along a chain, which occurs in some transinembrane<br />

peptides, a new class of helices results, the so-called /?-helices[I 18, 1191. Such chains<br />

can also form double-stranded structures with interchain hydrogen bonds.<br />

The simplest polypeptide structure that has intermolecular hydrogen bonds<br />

throughout its lattice is that of PGII, which basically has threefold helical symmetry<br />

[120]. (Because of its achiral C", PGII can be a right- or left-handed helix.) Although<br />

the originally proposed structure was modeled with parallel chains, the ease<br />

of mechanical conversion of PGII to PGI and the evidence for antiparallel chains<br />

in single crystals of PGI led to the suggestion [ 1211 that an antiparallel-chain structure<br />

should exist. Such a model was subsequently proposed 11221, and a detailed


272 5 ViDrutiomil <strong>Spect</strong>roscopy oj Polvpepticles<br />

Figure 5-8. Antiparallel-chain structure of polyglycine 11. (From [ 1231)<br />

vibrational analysis of the two structures 11231 showed that the antiparallel-chain<br />

structure (Figure 5-8) was indeed preferred. In these studies, the suggestion that<br />

C"-H ... O=C hydrogen bonds were present [ 1241, which was supported by spectroscopic<br />

studies [ 125, 1261, was convincingly confirmed by normal-mode analysis<br />

[123]. The same chain symmetry, in a required left-handed helix, is found in polyproline<br />

II, which has trans peptide groups 1127, 1281, a short contact again suggesting<br />

the possibility of a C"-H...O=C hydrogen bond [129]. It is interesting that,<br />

from circular dichroisrn studies [130], a polyproline I1 type of structure, with 2.5-<br />

3.0 residues/turn [131], was indicated to be a preferred local conformation in many<br />

'unordered' polypeptides in solution. This proposal has since received strong support<br />

from Rainan and normal-mode studies [ 1321 as well as from a more detailed<br />

analysis of the circular dichroism results [133]. A triplet of left-handed PGII or<br />

polyproline 11 helices can be thought of as the topological parent of the triplestranded<br />

right-handed coiled-coil structure of collagen [ 134, 1351. When the peptide<br />

groups in polyproline take a cis configuration, i.e., polyproline I, the chain adopts a<br />

right-handed 10 residue/3 turn helical structure [ 1361.<br />

The variety of helical as well as extended-chain and turn [5] structures emphasizes<br />

the flexibility of the polypeptide chain, and thereby the importance of achieving a<br />

full understanding of its vibrational dynamics as a function of conformation.


Frequency (cm-'1<br />

100%<br />

b<br />

700 500 300 100 (ern-')<br />

Figure 5-9. Infrared spectra of a-poly(L-alanine). (a) Mid-IR region. (From [85]).(b) Far-IR<br />

region. (From [I 371). (-) Electric vector perpendicular to stretching direction. (- - -) Electric vector<br />

parallel to stretchin2 direction.<br />

5.4.2.2 a-Helix<br />

The a1-PLA structure is well determined from X-ray diffraction [ 1041, and therefore<br />

permits a secure refinement of an empirical force field. The helical symmetry results<br />

in three optically active modes, defined by the phase difference between similar<br />

vibrations in adjacent uiiits of the helix: A(0") - 28 modes, Raman, IR( 11); El (99.6')<br />

- 29 doubly degenerate modes, Raman, IR (1); E2 (199.1")-30 doubly degenerate<br />

modes, Raman. Extensive IR and Raman studies on a-PLA and a-PLA-ND [5]<br />

have provided good experimental data. Polarized IR spectra [85, 1371 are shown in<br />

Figure 5-9 and Raman spectra [138] in Figure 5-10. The force field for the normalinode<br />

calculation [lo61 was refined using that of /?-PLA [22] as a starting point.<br />

The frequencies and PEDs [lo61 are given in Table 5-9; the iins frequency error is<br />

1.6 cm-'.


I I I I I I I I I I I I I<br />

1600 1400 I200 loo0 800 Mxl 400 (crn-'l<br />

Figure 5-10. Ratnan spectra of a-poly (L-alanine) (-)<br />

381)<br />

Undeuterated. ( ..' ) N-deuterated. (From<br />

The observed and calculated amide I mode splittings agree quite well, and are<br />

much smaller than for p-PLA. (The specific A-Ez splitting, due to TDC, remains to<br />

be verified; this will depend on having good polarized Raman spectra.) The mode<br />

now contains less CN s and more C"CN d than P-PLA, and a change in the small<br />

(< 10) NH ib contribution is largely responsible for the observed 8 cni-' downshift<br />

(calculated 10 cm-' ) on N-deuteration [ 1061. The calculated unperturbed amide I1<br />

modes at 1529 (A), 1532 (El), and 1536 (Ez) cn1-l again show how important TDC<br />

is in explaining the observed 29 cn1-l A-El splitting. The NH ib contribution to<br />

amide I11 now dominates over Ha b as compared with p-PLA, and CN s does not<br />

appear. The increase in these frequencies, from the 1240-1220 cm-I region in p-<br />

PLA to the 1280-1260 cm-' region in a-PLA, must be due to the difference in<br />

conformation (and the resulting change in interaction between H" b and NH ib),<br />

since the relevant force constants are in fact smaller in the a-helix than in the p-sheet<br />

[106]. The lower amide V frequencies in a-PLA, 658 and 618 cm-l, than in /I-PLA,<br />

706 cin-', are undoubtedly a result of the weaker hydrogen bond in the helical<br />

structure (N...O = 2.86 A in a-PLA versus 2.73 A in P-PLA).<br />

Other synthetic polypeptides adopt the al-helix structure, and it is of interest to<br />

see how their normal modes are influenced by a change in the side chain. A polypeptide<br />

that has been analyzed with the same main chain force field as a-PLA is<br />

a-poly(L-glutamic acid), a-PLGH [139]. Some of its modes are compared with those<br />

of q-PLA in Table 5-10. The amide I mode is hardly affected. The observed amide<br />

I1 mode A-El splitting increases somewhat, from 29 cm-l in a-PLA to 40 cn-I in<br />

a-PLGH; since the eigenvectors are identical [139], this may reflect a very subtle<br />

change in structure that alters the TDC interactions. The amide I11 modes are<br />

somewhat more affected, and this is a result of the side chain: CitH? tw now contributes<br />

to the E2 and El modes. Analogous changes influence skeletal modes in the<br />

1200-900 cm-' region, in the case of the CN s, CNC" d modes even altering the<br />

frequency order of the species. The aniide V modes are also shifted by such changes<br />

in eigenvectors, and these have an even greater impact on the low-frequency skeletal<br />

modes. The constellation of such changes may well be characteristic of the side<br />

chain.


5.4 Polypptidt~s 215<br />

Table 5-9. Observed and calculated frequencies !in cin-') of ccr-poly(L-alanine).<br />

Observed" Calculated Potential energy distributionb<br />

Ramaii IR A EI E2<br />

1655s<br />

1658VS 1 1<br />

1543VW 1545VS I<br />

15 16Msh<br />

1458s 1458VS 1 ) 1<br />

1377W<br />

1 3 38Msh<br />

13268 1328sh / /<br />

1308M 1307 /I 1<br />

1278W<br />

1271 W l270M I<br />

1261 W 1265Msh<br />

1167M<br />

1105s<br />

1050W<br />

1017W 1016W 1 1<br />

910W 968M 11<br />

940VW<br />

908VS 909M 1)<br />

8933 I<br />

882W<br />

173VW<br />

156W<br />

1657 CO $32) CN s( 10) CTN d(10)<br />

1655 CO ~(82) CN s( 11) C"CN di 10)<br />

1519<br />

1452<br />

1538<br />

1645 CO s(83) CN s!12) CTN d(10)<br />

1540 NH ib(46) CN s(31) CO ib(12) C"C ~(11)<br />

NH ib(46) CN s(33) CO Ib(l1) C"C silo)<br />

NH ib(45) CN si34) CO ib( 11)<br />

CHI abl(47) CH; ab2(41) CH3 11(10)<br />

1452 CHI ab2(54) CH3 abI(31)<br />

1452 CH3 db2(59) CHI abl(26)<br />

1451 CH3 ab2j45) CH3 abl(41) CH3 r2(10)<br />

1451 CH3 abl(57) CH3 ab2(31)<br />

1451 CH3 abl(62) CHI ab2(26)<br />

1381SA 1379 1379 1379 CH; sb(100)<br />

1349 Ha bl(34) NH ib(17) (2°C s(17) C"Cp s(10)<br />

1345 H" bl(25) H" b2(20) C"C s(17) NH ib(12)<br />

1334 H" b2158) C"C s(16)<br />

1314 H" b2(81)<br />

H" b2(56)H" bl(23)<br />

{ 1301 I3O8 H" bl(62)<br />

1287 NH ib(23) NC" s(18) H" bl(l8)<br />

1278 NH ib(28) H" b2(15) NC' s(14) H" bl(l1)<br />

1262 NH ib(40) Ha b2(28)<br />

NC" ~(32) CH: 11(20) C"C" ~(6)<br />

1170s 11 I<br />

{1178 1167 C"CD s(32) NC" 421) CH3 rl(15) H" bl(10)<br />

1162 C"CD s(41) NC" ~(16) H' bl(12) CH3 rl(12)<br />

1115 CTL' s(64) CH; r2(15)<br />

1108S1<br />

1103 C"Cfl s(39) CH? r2(18)<br />

1094 C'CD s(26) CHI r2(21)<br />

1043 CH3 rl(47) Ha bl(20)<br />

1051VS i 1037 CH3 rl(39) H" bl(22) CH3 i2(15)<br />

1026<br />

CH; r2(28) H" bl(25) CH: rl(24)<br />

982 CHI il(34) NC" ~(29) CH3 12(17) C"C ~(15)<br />

955 CH; r2(32) NC" 420) CH; rl(15)<br />

951 CH; r2(41) NC" s(16)<br />

910<br />

CNs (23) CNC" d( 16) CO 1b( 12) CH3 r2( 1 I ) CO s( 10)<br />

901<br />

CN s(18) C"C s(12) CO ibi12) C'CD $10)<br />

896 C'"'s(l9)C"C a(l4) CN s(14) CO ib(12) NC" s(l1)<br />

714M<br />

780 CO ob(42) Cp bl( 10) CN t( 10)<br />

767 CO ob(57) Cp bl(10)<br />

754 CO obi38) CN t(30)


276 5 Vibrational SiJectroscopy of Polypeptides<br />

Table 5-9. (contiriued)<br />

Observed" Calculated Potential energy distributionh<br />

Raman IR A El El<br />

-<br />

693M 691Wsh 11 700<br />

675<br />

662W 658s L 660<br />

637<br />

618s I 608<br />

589<br />

I537<br />

530VS<br />

375s<br />

328W<br />

310s<br />

294M<br />

260M<br />

209VW<br />

189M<br />

165M<br />

159s<br />

87W<br />

375s i 374<br />

-366Msh 367<br />

3248 I 326<br />

290M 11 307<br />

259Wsh 264<br />

240W 245<br />

223VW 230<br />

188M i 197<br />

l63M I 155<br />

120s 1 1 136<br />

113MshI 96<br />

84W 94<br />

40<br />

492<br />

366<br />

310<br />

244<br />

205<br />

151<br />

87<br />

49<br />

38<br />

"2°C d(36) C"CN d(16)<br />

CN t(32) CO ib( 18) C"C s( 14)<br />

CN t(37) NH ob(21) NC"C d(12)<br />

CN t(59) NH ob(43) NH...O ib(13)<br />

CN t(47) NH ob(23) CO ob( 15) CO ibi 12)<br />

CN t(68) NH ob(36) CO ob(26) NH.,.O ibill)<br />

CO ib(29) C"CN d(21) c"C s(17) CB b2(17)<br />

NH ob( 11)<br />

NC'C d(31) C'T s(14) CTN d(1l) CO ib(1l)<br />

NCT d(37) CO ib(17) (2°C s(16)<br />

CO ob(16) NH ob(16) Cp b2(15) C"CN d(15)<br />

CO ib( 15) CNC" d( 10)<br />

CO ob(21) Cp bl(17) NC"C d(14) NH ob(13)<br />

d b2(11)<br />

Cp b2(49) C"CN d(22) Cp bl(16)<br />

Cp b2(42) Cp bl(19) CO ib(16)<br />

CNC" d(30) CO ib(20) Cp bl(20) CO ob(17)<br />

Cp b2(34) CO ib(29) CNC" d(15)<br />

C"C0 t(35) Cfi b2(14)<br />

caCD t(91 j<br />

co t(95)<br />

C'CB t(63)<br />

CTN d(27) Cp b2(25) Cp bl(12) CO ob(l0)<br />

CaCN d( 15) Cp b2( 12) CO ob( 12)<br />

CNC"d(33)C'CNd(19) Cp b1(14)NHob(14)<br />

NC"C d( 12)<br />

NH ob(43) CNC" d(20) NC'C d( 11)<br />

CNC" d(34) CTN d(21) NC"C d(16) Co bl(12)<br />

NH ob(24) C'C t(20) NC" t(l8) CN t(18) H...O ~(10)<br />

CN t(27) NC" t(23) C"C t(22) H...O ~(15) NH ob(l0)<br />

NH ob(33) NH-.O ib(20j Cp bl(12) NCT d(10)<br />

CNC" d(l0)<br />

NH ob(38j NC" t(24) H-.O s(23) CN t(15) Cp bl(13)<br />

NH ob(58) Cp bl(20) H...O ~ (ll)<br />

(2°C t(53) H...O ~(18)<br />

a S: strong, M: medium, W: weak, V: very, sh: shoulder, 1): parallel dichroism, 1: perpendicular<br />

dichroism.<br />

s: stretch, as: antisyrnnietric stretch, ss: symmetric stretch, b: bend, ib: in-plane bend, r: rock,<br />

d: deformation, t: torsion. Constributions 2 10.


~~ ~~ ~~~ ~ ~ ~~<br />

5.4 Polypeptides 277<br />

Table 5-10. Comparison of some modes of ccl-poly( L-glutaniic acid) and q1-poly(L-alanine) with<br />

those of ccl-poly(L-alaniiie).<br />

Moded u1-PLAb ~I-PLGH' MII-PLA<br />

Observed Calculated Observed Calculated Observedd Calculatedb<br />

I<br />

1658 1657iA)<br />

1655 1655(El)<br />

1653<br />

1652<br />

1657(A)<br />

1655(El)<br />

1667 1667(A)<br />

1661 166O( El)<br />

I1<br />

1545 1538(El)<br />

1516 1519(A)<br />

1 550<br />

1510<br />

1537(Ei)<br />

1517(A)<br />

1547 1540( Ei )<br />

1515(A)<br />

I11<br />

1338 1345(El)<br />

1278 1287iE2)<br />

1270 1278(E1)<br />

1265 1262(A)<br />

1340<br />

1296<br />

1283<br />

1 326(El)<br />

12991 E2)<br />

1287(Ei)<br />

1263(A)<br />

1346(E1<br />

)<br />

1281 (E2)<br />

1272(El)<br />

1260( A)<br />

NCn s, C"CP s<br />

1178(A)<br />

1 167(El)<br />

1118<br />

1 l29(A)<br />

1129(El)<br />

1175(A)<br />

1 169(EL)<br />

CN s. CNC" d<br />

V<br />

909 910(A)<br />

893 901 (€5<br />

658 660(E1)<br />

618 608(E1)<br />

928<br />

924<br />

670<br />

618<br />

929( El)<br />

922(A)<br />

678(Ei)<br />

626iE1)<br />

909(A)<br />

900(E1)<br />

1<br />

615(Ei)<br />

CO ib, C"CN d<br />

528 537(A)<br />

565<br />

549(A)<br />

536(A)<br />

NC'C d<br />

528 522(El )<br />

515(E1)<br />

518(E1)<br />

CO ob, Cb b<br />

Cb b<br />

CTP t<br />

375 374(Ei)<br />

366 367(A)<br />

292 307(A)<br />

260 264(A)<br />

409<br />

318<br />

280<br />

402(E1)<br />

368(A)<br />

334(A)<br />

265(A)<br />

376(A)<br />

369(E1)<br />

302jA)<br />

274(A)<br />

a I, 11, 111. V: amide modes, in cm-'; s: stretch, b: bend, ib: in-plane bend, ob: out-of-plane bend,<br />

d: deformation, t: torsion.<br />

From [106].<br />

From [139].<br />

From [140].<br />

Small changes in backbone structure, such as are represented by the aII-helix, also<br />

lead to predictable spectral changes, as seen from Table 5-10. The amide I mode,<br />

found about 10 cm-' higher in (putative [107]) qr-helices [140], is expected to be<br />

particularly affected because of the weaker hydrogen bond resulting from the tilted<br />

peptide group. The other higher-frequency modes do not change very much, but<br />

there are some possible characteristic changes in the low frequency region: the frequency<br />

order of the symmetry species of the CO ob, Cb b modes inverts, and the<br />

frequency separation between Cb b and C*Cb t is expected to decrease significantly<br />

in an. Thus, vibrational spectra can be sensitive to even very small changes in<br />

structure.


218 5 Vibratioid <strong>Spect</strong>roscopy oj Pollyeprides<br />

In addition to frequencies, dichroic properties can be altered by small structural<br />

changes. Since the peptide group in the uII-helix is more inclined to the helix axis<br />

than it is in the a,-helix, the DRs of amide modes are expected to differ significantly<br />

between the two structures. To calculate the DR of, for example, ainide I of an<br />

a-helix we need to know (?g/dQ~),, and (l@/?Q,),, which involves knowing the L,,<br />

and dc/dS, (see Eq. (5-5)). Since the L,r are known from the normal mode calculation<br />

and the d$/ciSl have been obtained from ab azirio calculations [63], it becomes<br />

possible to synthesize a band intensity profile for a given (even finite) ol-helix structure.<br />

Such a calculation has been done for the oll-helix and the a11-helix for amide I<br />

and aniide I1 by computing the IR intensity profiles for the two polarization components<br />

of these two bands [141]. The results show that the DRs are significantly<br />

different enough to pennit identification of these structures from this property.<br />

The ability to predict a band intensity profile opens up an important additional<br />

dimension in vibrational analysis. It means that we will be able to relate subtle<br />

spectral differences to small structural changes with a greater degree of confidence.<br />

Thus, it has been possible to confirm that an observed three-component contour of<br />

the amide I band of tropomyosin is indeed expected for a coiled-coil a-helix [142].<br />

Extensions to understanding the normal modes of proteins become possible [ 1431.<br />

The systematic incorporation in an SDFF of dipoles and dipole fluxes to calculate<br />

IR intensities [ 1441 will finally bring to the vibrational analysis of polymeric niolecules<br />

the completeness and flexibility needed to make it a much more powerful<br />

structural tool.<br />

5.4.2.3 3lo-Helix<br />

The 3lo-helix represents a different topology than the a-helix, being 4 - I rather<br />

than 5 - 1, and is therefore of importance in studying the vibrational dynamics<br />

of polypeptide helices. At present, its only clear identification has been in PAIB, so<br />

its characterization from this polypeptide may lack soine generality. On the other<br />

hand, the excellence of the experimental data [I 101 makes its vibrational analysis<br />

secure.<br />

Since the structure of PAIB has not been determined in detail by diffraction<br />

methods, the normal-mode studies [110] were based on a 310 structure obtained<br />

from conforniational analysis [l 1 11 (Figure 5-7). The vibrational studies compared<br />

experimental data with predictions for two helices, and the results clearly favored<br />

the 310- over the a-helix. The threefold screw symmetry of this structure results in<br />

El and E2 species modes reducing to doubly degenerate E species modes. The main<br />

chain force field was the same as that for oll-PLA, with additional force constants<br />

refined for the (CH3)2 group. Some inodes from the full analysis [110] are compared<br />

with those of q-PLA in Table 5-1 1.<br />

The observed amide I frequencies are slightly lower for the 3lo-helix than for<br />

the cq-helix because of the slightly stronger hydrogen bond (N...O = 2.83 A versus<br />

2.86 A, respectively). The increased splitting is undoubtedly due to the different<br />

TDC interactions as a function of conformation. This probably also accounts for<br />

the large change in the A-E splitting of the aniide 11 modes. The nominal aniide 111<br />

modes (they contain no CN s and NH ib dominates only in the 1312 cn-' mode!


5.4 Poliyeptides 219<br />

Table 5-11. Comparison of some modes of 3 ~~,-poly(cc-arninoisobutyri~ acid) and ccc-poly( L-<br />

alanine j .<br />

Mode" 3 !"-PA 1 B" XI-PLA<br />

Observed Calculated Calculated<br />

I<br />

II<br />

III<br />

Raman<br />

1647<br />

1531<br />

1339<br />

1313<br />

1280<br />

CN s, CNC' d 908<br />

V<br />

NC"C d 594<br />

C"CN d, CO ib 568<br />

IR<br />

1656 11 1665(A)<br />

1661iE)<br />

1545 i 1547(E)<br />

1533(Aj<br />

1346(E)<br />

1312(A)<br />

1280 )I 1287iA)<br />

905 I/ 905(A)<br />

694 1 701(E)<br />

680 I1 676(A)<br />

595 II 594(A)<br />

557(A)<br />

1657(A)<br />

1655iEl)<br />

1538( El)<br />

15 19(A)<br />

1345(El)<br />

1287(E1)<br />

1278( E2)<br />

1262(A)<br />

910(A)<br />

660(E1)<br />

608(E)<br />

589(A)<br />

522(E1)<br />

537iA)<br />

a I, 11. 111, V: ainide modes in cm-l; s: stretch, d: deformation, ib: in-plane bend.<br />

From [110]. 11:<br />

parallel dichroism, I: perpendicular dichroism.<br />

From [106].<br />

are well accounted for, and exhibit a quite different pattern from that of the MIhelix.<br />

This is undoubtedly due in part to the different involvement of side-chain<br />

structures in the two polypeptides. Since no such major difference is seen in the<br />

aniide V eigenvectors, the large frequency differences between the two helices must<br />

be due to differences in hydrogen-bonding geometry and/or main-chain conformation.<br />

(The disappearance of the 694 and 680 cm-l bands on N-deuteration [110]<br />

unambiguously confirms their assignments to ainide V.) The skeletal N CT d and<br />

CTN d, CO ib modes have reversed frequency order (and in one case changed<br />

species) in 3l"-PAIB, and again, this is probably partly due to the involvement of<br />

side-chain coordinates in the eigenvectors. Future studies of, for example, a 3 10-<br />

helix of PLA should help to establish the possible generality of these results.<br />

5.4.2.4 Polyglycine I1<br />

The PGII structure is the simplest one that is representative of polypeptide helices<br />

without intramolecular hydrogen bonding. In this case, the antiparallel chains of


280 5 Vibrational <strong>Spect</strong>roscopy of Polvpeytides<br />

essentially threefold helices [ 122, 1231 are completely intermolecularly N-H . .. O=C<br />

hydrogen bonded (Figure 5-8), with strong spectroscopic evidence [ 1231 for additional<br />

C"-H...O=C hydrogen bonds. The basis for this conclusion is of interest<br />

since it demonstrates the sensitivity of the spectroscopic technique to subtle features<br />

of the structure.<br />

The vibrational analysis [ 1231 was based on comparing the normal modes of<br />

a parallel-chain and an antiparallel-chain structure of PGII. In both structures all<br />

N-H... O=C hydrogen bonds can be made. However, whereas in the parallel-chain<br />

arrangement all CH2 groups participate equivalently in C"-H ... O=C hydrogen<br />

bonds (implying that all C=O groups have bifurcated hydrogen bonds), in the<br />

antiparallel-chain arrangement only every third CH2 group can be involved in such<br />

interactions, and only between like-directed chains. The result is that the chain no<br />

longer has strict threefold symmetry, and as a result the degeneracy of the E species<br />

modes is broken. This leads to the expected, and observed, presence of a larger<br />

number of bands than predicted by a strictly threefold-symmetric chain structure,<br />

and also of bands that occur in regions where no modes are predicted for the<br />

parallel-chain structure. Such symmetry-based conclusions are powerful constraints<br />

on structural possibilities.<br />

The antiparallel-chain structure (Figure 5-8) has only a screw axis of symmetry,<br />

and its modes are distributed as follows: A-62 modes, Raman, IR; B-61 modes,<br />

Raman, IR. As for PGI, extensive IR spectra on isotopic derivatives (NH, CD2;<br />

ND, CH?; and ND, CD2) were available [97] as well as Raman spectra on the N-<br />

deuterated molecule [98]. The PGI force field [96] was used as a starting point, and<br />

refinement required small adjustments in 10 of -70 intramolecular force constants<br />

[123]. Amide mode frequencies are given in Table 5-12; the overall rnis frequency<br />

error is 5.4 cm-'.<br />

The hydrogen bond in PGII is stronger than that in PGI (N...O=2.69 b, vs<br />

2.91 A), so it is not surprising that the average amide I mode frequency is lower in<br />

PGII (the unperturbed modes are at 1651.0 f 1.7 cm-I). The very different splittings<br />

are due to the different TDC interactions for the two structures. The somewhat<br />

higher amide I1 modes are also consistent with stronger hydrogen bonds. The<br />

amide I11 modes with significant NH ib, CN s contributions are found near 1280<br />

cm-', strongly mixed with CH2 tw, quite different from the situation in PGI. The<br />

isolated NH ib contribution is now found in the 1400-1300 cm-' region. The aniide<br />

V modes in PGII are divided between a higher frequency group near 750 cmP1 and<br />

a lower frequency group near 670 cm-'. This is a much larger separation than<br />

found for other helical structures, and may be a consequence of the interchain<br />

hydrogen bonding.<br />

5.5 Summary<br />

The main message conveyed by the body of present vibrational spectroscopic<br />

studies of peptides and polypeptides is that detailed information on their structures


Table 5-12. Aniide mode frequencies (in cm-' ) of antiparallel-chain polyglycine 11.<br />

Observed' Calculated Potential energy distribution<br />

Rainan IR A B<br />

(1656 CO s(71) CN ~(20) C"CN d(l0)<br />

1654 CO s(74) CN s(20) C"CN d( 10)<br />

1654VS 1655W<br />

1653 CO s(72) CN ~(20) CTN dil0)<br />

( 1651<br />

CO s(73) CN ~(19) C"CN d(10)<br />

1649<br />

CO 473) CN s(20) CTN d(10)<br />

1640VS 1645 CO 474) CN s(19) CTN d(l0j<br />

1560W 156OW 1565 1565 NH ib(59) CN s(18)<br />

1550s<br />

1555 1552 NH ib(52) CN s(20) C"C s(10)<br />

1548 1548 NH ib(54) CN s(21) '2°C s(12)<br />

1383MS 1377M 1353 1380 CH? ~(58) C"C ~(15) NH ib(13)<br />

1350 1350 CH2 w(46) CH2 b(16) NH ib(16) C"C s(14)<br />

1334VW 1332VW 1344 1345 CH2 w(50) NH ib(16) CH2 b(14) C" C ~(13)<br />

1303 1304 CH1 tw(31) NH ib(14) CH2 w(13) CN s(12)<br />

1283M 1283M 1290 1290 CH? tw(29) CH2 w(23) NH ib(13) CN s(12)<br />

752VW 751W 760 759 CN t(20) NH ob(19) NH-.O ib(14) NC"C d(13)<br />

CO ib( 12)<br />

742VW 740 CN t(53) NH.-.O ib(22) NH ob(l8)<br />

673M<br />

740M 738<br />

678<br />

661<br />

CN t(49) NH...O ib(20) NH ob( 16) CO ib( 11)<br />

CN t(77) NH ob(27j NH t(l1) CO ob( 10)<br />

CN t(83) NH ob(28) NH t( 11) NH.-O ib( 10)<br />

CN t(88) NH ob(30) NH t(14)<br />

{<br />

a Data taken from [123]. S: strong, M: medium, W: weak, V: very.<br />

b s: stretch, b: bend, ib: in-plane bend, ob: out-of-plane bend, d: deformation, w: wag, tw: twist,<br />

t: torsion. Contributions 2 10.<br />

and interactions can be obtained if experimental data are interpreted by careful<br />

normal-mode analysis. Although the systematic application of extensive empirical<br />

force fields has revealed the validity of this approach [5], the full power of this<br />

technique remains to be exploited. Our discussion has concentrated, because of<br />

space limitations, on the amide modes, but of course the sensitivity to conformation<br />

and forces resides in the entire spectrum. The key to extracting this information will<br />

be the physical reliability of the force fields used in the normal-mode calculations.<br />

We have come a substantial distance toward this goal with existing force fields [5],<br />

but improvements can be made and some are already in sight. These presently involve<br />

empirical force fields, such as the development of an ab initio-based MI-PLA<br />

force field [145] utilizing better IR [145] and polarized Raman [146] data as well as<br />

an understanding of helix vibrations [ 1471 that avoids the weak-coupling and perturbation<br />

approximations 191. However, the ultimate goal will be achieved when<br />

we have a conformation-dependent force field as represented by an SDFF for the.<br />

polypeptide chain [51].


282 5 Vibrational <strong>Spect</strong>roscopy of Polypt>ptiu'c..c<br />

Acknowledgments<br />

The author's contributions to the development of this field have been supported by<br />

the National Science Foundation through grants from its Biophysics and <strong>Polymer</strong>s<br />

Programs. Helpful discussions with Noemi G. Mirkin and Kim Palmo are gratefdly<br />

acknowledged.<br />

5.6 References<br />

Stryer, L., Biocheniistry. Fourth Edition. New York: W. H. Freeman & Co., 1995.<br />

Richardson, J., Ad[). Protein Cliem. 1981, 34, 167-339.<br />

Surewicz, W. K., Mantsch, H. H.. Bioihini. Biopliys. Acto 1988, 9-72, 115-130.<br />

Surewicz, W. K., Mantsch, H. H., Chapman, D., Biochemistry 1993, 32, 389-394.<br />

Krinim, S. and Bandekar, J., Ado. Protein Clieni. 1986, 38, 181-364.<br />

Herzberg, G., Molecular <strong>Spect</strong>ra and Struelure. I1 Injkved cind Rninan <strong>Spect</strong>in oj'Pulyatornic<br />

Molectiles. New Jersey: Van Nostrand-Reinhold, 1945.<br />

Wilson, E. B., Decius, J. C., Cross, P. C., Molecular. Vibrations. New York: McGraw-Hill,<br />

1955.<br />

Califano, S., Vibrational States. New York: Wiley, 1976.<br />

Higgs, P. W., Proc. Roy. Soc. London Ser. A 1953,220, 472-485.<br />

Shimanouchi, T., in: Physical Chemistry: An Aducmceti Treatise: Eyring, H., Henderson, D.,<br />

Jast, W. (Eds.). New York: Academic Press, 1970; Vol. 4, pp. 233-306.<br />

Person, W., Zerbi, G., Eds., Vibrationcil Intensities in IrIfitiretl aIid Rtirnnn Siiectroscopy.<br />

Amsterdam: Elsevier, 1982.<br />

Qian, W., Krimni, S., J. Plzys. Chem 1993, 97, 11578-1 1579.<br />

Cheam, T. C., Krinim, S., J. Mol. Struct. 1986, 146, 175-189.<br />

Tasumi, M., Krimm, S., J. Chem. Plzys. 1967, 46, 755-766.<br />

Krimm, S., Abe, Y., Proc. Natl. Act7d. Sci. U.S.A. 1972, 69, 2788-2792.<br />

Moore. W. H., Krimm, S., Proc. Nrrtl. Acntl. Sci. U.S.A. 197.5, 7-7, 4933-4935.<br />

Cheam, T. C.! Krimm, S.. Chem. Plijx Lett. 1984, 107, 613-616.<br />

Abe, Y., Krimm, S., Biopo/ytners 1972, 11, 1817-1839.<br />

Moore, W. H., Krimm, S., Biopolymers 1976, 15, 2439-2464.<br />

Cheam, T. C., Krimm, S., J. Cliern. Phys. 198.5, S2, 1631-1641.<br />

Schachtschneider, J. H., Snyder, R. G., <strong>Spect</strong>rochiin. Actu 1963, 19, 117-168.<br />

Dwivedi, A. M., Krimm, S., J. Phy.~. Clzein. 1984, 88, 620-627.<br />

Hehre, W. J., Radoni, L., v. R. Schleyer, P., Pople, J. A,, Ab Iizitio Moleciilrir Orbited Theory.<br />

New York: John Wiley & Sons, 1986.<br />

Fogarasi, G., Pulay, P., in Vibrational Siiectm mid Structure: Durig, J. R. (Ed.). Amsterdam:<br />

Elsevier, 1985; Vol. 14, pp. 125-219.<br />

Mirkin, N. G., Krimm, S., J. Mol. Strtrct. (Theockern) 1995, 334, 1-6.<br />

Mirkin, N. G., Krimm, S., J. Phys. Cheni. 1993, 97, 13887-13895.<br />

Qian, W., Krinim, S., J. Phj>s. Chem. 1994, 98, 9992-10000.<br />

Qian, W., Krimm, S., Biopolyiiievs 1994, 34, 1377-1394.<br />

Zhao, W., Bandekar, J., Krimm, S., J. Mol. Struct. 1990, 238, 43-54.<br />

Qian, W., Zhao, W., Krimm, S., J. Mol. Struct. 1991, 250, 89-102.<br />

Qian, W., Krirnm, S., Bio&niers 1992, 32, 1025-1033.<br />

Qian, W., Krimm, S., Biopolyniers 1992, 317, 1503-1518.<br />

Cheam, T. C., Krimm, S.. J. Mol. Struct. ( Theoclzeni) 1990.206, 173-203.<br />

Cheam. T. C., Kriinm, S., J. Md Struct. (Tlzeoclieni) 1989, IS8, 15-43.


1351 Lifson, S., Stern, P. S., J. Cliew. Phys. 1982. 77. 4542-4550.<br />

[36] (a) Lii, J.-H., Allinger, N. L., J. Am. Cliem. Soc. 1989, 111, 8566-8575. (b) Allinger, N. L.,<br />

Chen, K., Lii, J.-H.. Corny. Chi. 1996, 17, 642-668.<br />

[37] Lii, J.-H., Allinger, N. L., J. Cony. Cl~ct~i. 1991, 13, 186-199.<br />

[38J Dasgupta, S., Goddard 111, W. A., J. Cliem. Plij~s. 1989, 90, 7207-7215.<br />

1391 Karasawa, N., Dasgupta, S., Goddard 111, W. A., J. Phys. Chem 1991, 95, 2260-2272.<br />

1401 Hwang, M. J., Stockfisch, T. P., Hagler, A. T., J. iln.1. Clieni. Soc. 1994, 116, 2515-2525.<br />

1411 Cornell, W. D., Cieplak, P., Bayly, C. I., Gould, 1. R., Merz, Jr., K. M., Ferguson, D. M.,<br />

Spellineyer, D. C., Fox, T., Caldwell, J. W., Kollman, P. A., J. Atit. Clieui. SOC. 1995, 117,<br />

5179-51 97.<br />

1421 Miwa, Y., Machida, K.? J. Am. Cl7e117. Soc. 1988, 110, 5183-5189.<br />

[43] Derreumaux, P., Dauchez, M., Vergoten, G., J. Mol. Str~rct. 1993,295, 203-221.<br />

1441 Derreumaux, P., Vergoten. G., J. Clwrn. Plz.vs. 1995, 10-1, 8586-8605.<br />

1451 Palmo, K., Pietila, L.-O., Kriinm, S., Coniputer.~ Clieiv 1991, 15, 249-250.<br />

1461 Palmo, K., Pietila. L.-O., Krimm, S., J. Cornp. Chem. 1991, 12, 385-390.<br />

1471 Palmo, K., Mirkin, N. G., Pietila, L.-O., Krimm, S. Macromolecules 1993, 26, 6831-6840.<br />

[48] Palmo, K., Pietila, L.-O., Krimm, S., Cowipicters Clzeriz. 1993, 17, 67-72.<br />

[49] Palnio, K., Pietila, L.-O., Krimm, S. J. Con7p. Clieni. 1992, 13, 1142-1150.<br />

[SO] Pahno, K., Mirkin, N. G., Krimm. S., J. Phys. Cliem. A 1998, 102, 6448-6456.<br />

1511 Palmo. K., Mirkin, N. G., Krimm, S., to be published.<br />

1521 Miyazawa, T., Shimanouchi, T., Mizushima, S., J. Clzeni. Phys. 1958, 29, 611-616.<br />

[53] JakeS, J., Schneider, B., Coil. Czech. Clieni. Cornrn~iri. 1968, 33, 643-655.<br />

1541 JakeS, J., Krimm, S., S/~ecrrochitn. Actn 1971, 27A, 19-34.<br />

[55] Rey-Lafon, M.. Forel, M. T., Garrigou-Lagrange, C., <strong>Spect</strong>rochitii. Act0 1973, 29A, 471-<br />

486.<br />

[56] Sugawara, Y., Hirakawa, A. Y., Tsuboi. M., J. Mol. Spcctrosc. 1984, 108, 206-214.<br />

1571 Sugawara, Y., Hirakawa, A. Y., Tsuboi, M., Kato, K., Morokuma, K., J. Mol. Spxmxc.<br />

1986,115, 21-33.<br />

[58] Balazs, A., J. Mol. Stvuct. (Tlleoch7i) 1987, 153, 103-120.<br />

[59] Mirkin, N. G., Krimm, S., J. Mol. Struct. 1991, 242, 143-160.<br />

[60] Ataka, S., Takeuchi, H., Tasumi, M., J. Mol. Stixct. 1984, 113, 147-160.<br />

[61] Mirkin, N. G., Krinim, S., J. Am. Clieni. Soc. 1991, 113, 9742-9747.<br />

1621 Williams, R. W., Biol~olyiiiers 1992, 32, 829-847.<br />

[63] Cheam. T. C., J. Mol. Struct. (Tlieoclieni) 1992, 257, 57-73.<br />

[64] Mirkin, N. G., Krimm, S., J. Mol. Struct., 1996, 377, 219-234.<br />

[65] Miyazawa, T., Shimanouchi, T., Mizushima, S., J. Clien?. Plzys. 1956, 24, 408-418.<br />

[66] Miyazawa, T., in: Polyaniiiio Acids, Po/.ypeptides, and Proteins: M. A. Stahmann (Ed.).<br />

Madison: University of Wisconsin Press, 1962; pp. 201-217.<br />

[67] Chen, X. G., Schweitzer-Stenner, R., Krimni, S., Mirkin, N. G., Asher, S. A., J. Ani. Che’hrnz.<br />

SOC. 1994,116, 11111-11142.<br />

1681 Chen, X. G., Schweitzer-Stenner, R.? Asher, S. A., Mirkin, N. G., Krimm, S., J. Plz~vs. Clzeni.<br />

1995, 99, 3074-3083.<br />

[69j Song, S., Asher, S. A., Krimm, S., Bandekar, J., J. Am. Chem Soc. 1988, 110, 8547-8548.<br />

[70] Miyazawa, T., Bull. Clieni. Soc. Japan 1961, 34, 691-696.<br />

[71] Fillaux, F., de Loze, C., Clzern. Phys. Lett. 1976, 39, 547-551.<br />

1721 Hsu, S. L., Moore, W. H., Krimm, S., Biopolyniecr 1976, 15, 1513-1528.<br />

[73] Mirkin, N. G., Krimm, S., private communication.<br />

[74] Cheam, T. C.! J. Mol. Sttwt. 1993, 295, 259-271.<br />

1751 Bandekar, J., Krimm, S., Biopolyiners 1988, 27, 885-908.<br />

[76] Sundius, T., Bandekar. J., Krimm, S., J. Add. Sfnrct. 1989, 214, 119-142.<br />

(771 Qian, W., Bandekar, J., Krimm, S., Biopo/.viiienr 1991, 31, 193-210.<br />

1781 Pauling, L., Corey, R.B.. Proc. Natl. Acnd. Sci U.S.A. 1951. 37, 729-740.<br />

[79] Pauling, L., Corey, R. B., Proc. Natl. Acad. Sci. US. A. 1953, 39, 253-256.<br />

[80] Moore, W. H., Krimm, S., Biopolyrners 1976, 15, 2465-2483.<br />

1811 Arnott, S., Dover, S. D., Elliott, A,, J. Add Biol. 1967, 30, 201-208.<br />

[S2] Bandekar, J., Krimm. S.. Biopo1ynier.r 1988, 27, 909-921,


284 5 Vibi~itionnl <strong>Spect</strong>roscopy qf Polypeptides<br />

[83] Chirgadze, Yu. N., Nevskaya, N. A.. Biopo/,vnwrs 1976, 15, 607-625.<br />

[84] Chirgadze, Yu. N., Nevskaya, N. A,, Biopolynevs 1976, 15, 627-636.<br />

[85] Elliott, A., Puoc. ROY. Soc. Loridoil Ser. A 1954, 226, 408-421.<br />

[86] Itoh, K., Katabuchi, H., Biopolwwrs 1972, 11, 1593-1605.<br />

[87] Masuda, Y., Fukushima, K., Fujii, T., Miyazawa. T., Biopo/wners 1969, 8, 9 1-99,<br />

[88] Dwivedi, A. M., Krimm, S., Macrornoleciiles 1982. 15, 186-193.<br />

[89] Fanconi, B., Biupo&r?ws 1973, 12, 2759-2776.<br />

[90] Krimm, S., Dwivedi, A. M., J. Rmm7n Spcctrosc. 1982, 12, 133-137.<br />

1911 Itoh, K., Nakahara? T., Shimanouchi, T., Oya, M., LJno, J., Iwaknra, Y., Biopo/jwwr.y 1968,<br />

6, 1759-1766.<br />

[92J Sengupta, P. K., Krimm, S., Hsu, S. L., Bioyolpiners 1984, 23, 1565-1594.<br />

[93] Keith, H. D., Padden, F. J.. Giannoni, G., J. h4o/. Biol. 1969, 43, 423-438.<br />

[94] Lotz, B., J. Mol. Bid. 1974, S7, 169-180.<br />

[95] Colonna-Cesari, F., Premilat, S., Lotz, B., J. Mol. Bid. 1974, 87, 181-191.<br />

[96] Dwivedi, A. M., Krimm, S., Mucroniolecules 1982. 15, 177-155.<br />

[97] Suzuki, S., Iwashita, Y., Shimanouchi, T., Tsuboi, M., Bioyolyrners 1966, 4, 337-350.<br />

[98] Small, E., Fanconi, B.. Peticolas. W. L., J. Clzeiri. Pliys. 1970, 52, 4369-4379.<br />

[99] Huggins, M. L., Chem Reos. 1943. 3-7, 195-218.<br />

[loo] Crane, H. R., Sci. Mon. 1950, 70, 376-389.<br />

[loll Bragg, L., Kendrew. J. C., Perutz, M. F., Proc. Roy. Soc. London Ser.. '4. 1950, 203, 321-357.<br />

[102] Pauling, L.. Corey, R. B., Banson, H. R., Proc. Nutl. Acud, Sci. U.S.A. 1951, 37, 205-211.<br />

[I031 Pauling, L., Corey, R. B., Proc. Nutl. Acurl. Sci. U.S.A. 1951, 37, 235-240.<br />

[I041 Amott, S., Dover, S. D., J. Mol. Bid. 1967, 30, 209-212.<br />

I1051 Miyazawa, T., J. Polynz. Sci. 1961, 55, 215-231.<br />

[lo61 Dwivedi. A. M., Krimm, S., Biopo/,ymers 1984, 23> 923-943.<br />

[lo71 Krimm, S., Dwivedi, A. M., Science 1982,216, 407-408.<br />

[lo81 Whitby, F. G., Kent, H., Stewart, F., Stewart, M., Xie, X., Hatch, V., Cohen, C., Phillips,<br />

Jr., G. N., J. Mol. Bid 1992, 227, 441-452.<br />

[lo91 Malcolm, B. R., Biopo/ymers 1983, 22, 319-321.<br />

[110] Dwivedi. A. M., Krimm, S., Malcolm, B. R., Biopol-vn7er.c 1984, 23, 2025-2065.<br />

[ill] Prasad, B. V. V., Sasisekharan, V., Macroniolecu/es 1979, 12, 1107-11 10.<br />

[112] Bradbury, E. M., Brown, L., Downie, A. R., Elliott, A,, Fraser, R. D. B., Hanby, W. E., J.<br />

Mol. Bid. 1962, 5, 230-247.<br />

[I 131 McGuire, R. E., Vanderkooi, G., Momany, F. A,, Ingwall, R. T.. Crippen, G. M., Lotan, N.,<br />

Tuttle, R. W.. Kashuba, K. L., Scheraga, H. A,, M~/croniolecules 1971, 4, 112-124.<br />

[114] Nambudripad, R., Bansal, M., Sasisekharan, V., Int. J. Peptide Protein Re.7. 1981, 18, 374-<br />

382.<br />

[115] Saski, S.: Yasumoto, Y.. Uematsu, I., Mucrornolecult~s 1981, 14, 1797-1801.<br />

[I 161 Low, B., Grenville-Wells, H. J., Proc. Narl. Acud Sci. US. A. 1953, 39, 785-801.<br />

11171 Donohue, J., Froc. Nutl. Acud Sci. U.S.A. 1953, 39, 470-478.<br />

[I181 Urry, D. W., Pvoc. Nutl. Acad. Sci. U.S.A. 1971, 68, 672-676.<br />

[I191 Lotz, B., Colonna-Cesari, F.? Heitz, F.. Spach, G., J. Mol. Biol. 1976, 106, 915-942.<br />

[I201 Crick, F. H. C., Rich, A,, Nuture 1955. 176, 780-781.<br />

[121] Krimm, S., Nature 1966,212, 1482-1483.<br />

[ 1221 Ramachandran, G. N., Ramakrishnan, C., Venkatachalam, C. M., in: Corzforinution of Biopolymers:<br />

Ramachandran, G. N. (Ed.) New York: Academic Press, 1967; Vol. 2, pp. 429-<br />

438.<br />

[123] Dwivedi, A. M., Krinini, S., Biopolymeus 1982, 21, 2377-2397.<br />

11241 Ramachandran, G. N., Sasisekharan, V., Ramakrishnan, C., Biochirn. Biophys. Actu 1966,<br />

112, 168-170.<br />

[125] Krimm. S., Kuroiwa, K.. Rebane. T., in: Conformi~tion of' Biopoljwiers: Ramachandran, G.<br />

N. (Ed.) New York: Academic Press, 1967; Vol. 2, pp. 439-447.<br />

[126] Krimm. S., Kuroiwa, K., Biopol-viners 1968, 6, 401-407.<br />

[I271 Cowan, P. M., McGavin, S., Ni7ture 1955, 176, 501-503.<br />

[IZSJ Arnott, S., Dover, S. D., Actu Cr~~mllogr. 1968, B24, 599-601.<br />

11291 Sasisekharan. V., Acfu Cry~t. 1959, 12, 897-903.


5.6 Refeverices 285<br />

[130] Tiffany, M. L., Krimm, S., Biopo/ynzer.s 1968, 6, 1379-1382.<br />

[131] Krimin, S., Mark. J. E., Proc. Nof/. Acnd Sci. U.S.A. 1968, 60, 1122-1129.<br />

[I321 Sengupta. P. K., Krimm, S., Biopo/i~rnt~rs 1987, 26, S99-Sl07.<br />

[133] Woody, R. W., Ado. Biop/ivs. Chetn. 1992. -7, 37-79.<br />

[134] Ramachandran, G. N., Kartha, G., Norm 1955, 176, 593-595.<br />

[135] Rich, A., Crick, F. H. C., J. Mol. Biol. 1961, 3, 483-506.<br />

[136] Traub, W., Shmueli, U., in: Aspecfs qf Protciii Structzirr: Ramachandran. G. N. (Ed.) New<br />

York: Academic Press. 1963; pp. 8 1-92.<br />

[ 1371 Itoh, K.. Shirnanouchi, T.. Bioyoljwwrs 1970, 9, 383-399.<br />

[138] Frushour, B. G., Painter, P. C., Koenig, J. L., J. Macrorml. Sci. Reo. hIacr.onzol. Cheni. 1976,<br />

C15, 29-1 15.<br />

[I391 Sengupta, P. K., Krimm, S., Biopo/.yn?er.r 1985, 24, 1479-1491.<br />

[140] Rothschild, K. J., Clark, N. A,, Biophys. J. 1979, 35, 473-488.<br />

[ 1411 Reisdorf, Jr., W. C., Krimin, S., Bioyh-vs. J. 1995, 69, 271-273.<br />

[142J Reisdorf, Jr., W. C., Krimm, S. Biocken~istr,~~, 1996, 35, 1383-1386.<br />

[143] Krimm, S., Reisdorf, Jr., W. C., Faraday Discuss. 1994, 99, 181-197.<br />

[144] Palmo, K., Krimm, S., J. Coinp. Chenz. 1998, 19, 754-768.<br />

[145] Lee, S.-H., Krimm, S., Biopolyniers 1998, 46. 283-317.<br />

[146] Lee, S.-H., Krimm, S., .I. Rarnnrz Spwrosc. 1998, 29, 73-80.<br />

[147] Lee, S.-H., Krimm, S., Clzern. Phys. 1998, 230, 277-295.


This Page Intentionally Left Blank


Index<br />

Amide modes 249<br />

Amorphous polymers 16<br />

Anisotropic surfaces 37<br />

Assignments 243<br />

Asynchronous spectrum 9ff<br />

Autopeaks 11<br />

Band-widths 177, 180<br />

Bilayers systems 157<br />

Biopolymers 27<br />

Biphenyl cyano-alkyl vibrational<br />

spectra 146<br />

Bipolaron 21 1, 213, 229<br />

- lattice 232<br />

Block copolymers 26<br />

Broadening of vibrational bands 172<br />

Chain folding 150<br />

Charge transport 231<br />

Chemical regularity of polymer<br />

chains 99<br />

Conducting polymer 207<br />

Conformational periodicity 99, 111<br />

Conjugated polymer 207<br />

Coordinates<br />

- local symmetry 93<br />

- localized 97<br />

- vibrational internal 91<br />

Cross peaks 11<br />

Cross-cosl-elation function 9<br />

p-Cyanophenyl p-n-hexylbenzoate<br />

(6CPB) 40<br />

- bandassignment 45<br />

- 2D-correlation analysis 48<br />

2DIR 1<br />

Defects<br />

- chemical, stereochemical 126<br />

- conformational 143<br />

- mass 142<br />

- vibrations of 129f, 142<br />

Density of states 117,127<br />

Dichroic difference 3, 6<br />

Dichroic ratio 6<br />

Dichroism 243<br />

- infrared 3<br />

Dipeptides 246,252<br />

Dipole transition moment 3<br />

Disordered polymers 122<br />

Doping 208,210<br />

Dynamic IR linear dichroism (DIRLD) 3<br />

Dynamic spectrum 2<br />

Electric field-induced orientation 38, 42<br />

- variable-temperature cell 43<br />

Electronic measurement set-up 44<br />

Elongation-recovery 77<br />

Fatty acids conformation 144, 157<br />

Ferini resonances 163.168<br />

Ferroelectric side-chain liquidcrystalline<br />

polymer (FLCP) 55<br />

- 2D-correlation analysis 60<br />

- polarization geometry and mesogen<br />

alignment 57<br />

- poling experiment 58<br />

- schematic representation of the<br />

segmental motions 61<br />

Force constants 89, 91<br />

- group 92<br />

Force fields 240<br />

- ab iiiitio 24.5<br />

- empirical 241<br />

- molecular mechanics 247<br />

- spectroscopically determined 237<br />

Fourier-transform infrared (FTIR) 33


288 Index<br />

Glass transition 18, 26<br />

HCI dynamics 13 1<br />

In-phase spectrum 5<br />

Intensities, infrared and Raman 91<br />

Inteiphase 26<br />

Laser-irradiation 63<br />

- apparatus for on-line measurements 63f<br />

Lattice dynamics 88<br />

Longitudinal Accordion Motion 159<br />

Methylene group coordinates 95<br />

Microdomains 26<br />

Na-doped poly@-phenylene) 217, 223<br />

Near-infrared spectroscopy 75<br />

Nematic liquid-crystalline guest-host<br />

system 52<br />

Nematic liquid-crystalline side-chain<br />

polymer (NLCP) 40f, 5 1<br />

- 2D-cori-elation analysis 5 I<br />

- switching process 52<br />

Neutron Scattering 119f<br />

NIR Raman spectrometry 2 16<br />

Nonndecane 180<br />

Non-bonded interactions 244, 248<br />

Normal modes 90<br />

- alanine dipeptide 253, 256<br />

- N-methylacetamide 251<br />

- 3,,-poly(a-aminoisobutyric acid) 278<br />

- polyglycine I 266<br />

- polyglycine I1 279<br />

- a-poly(L-alanine) 273<br />

- P-poly(L-alanine) 260<br />

- a-poly(L-glutamic acid) 277<br />

Normalized intensity 47<br />

p-Oligophenyl 218,221<br />

- dianion 219, 223<br />

- radical anion 219, 223<br />

Order parameter 36<br />

Orientation 3, 5ff<br />

Orientation factor 6<br />

Phonon dispersion curves lOlf<br />

Polaron 211,213,229<br />

- lattice 232<br />

Polyacetylene 208<br />

Poly(acetylene), ~ IZIIZS 105,197<br />

Polyaniline 229<br />

Polyesters 147, 139<br />

Polyethylene 18, 20<br />

- films 152<br />

- reflection spectrum 155<br />

- vibrations 102, 110, 112, 122, 124, 150,<br />

172<br />

<strong>Polymer</strong> blends 20, 22<br />

<strong>Polymer</strong>isation of ethylene, mechanism 138<br />

Poly(methy1 vinylether) 23<br />

Polypeptide structures<br />

- helical 269<br />

- sheet 258<br />

Poly@-phenylene 217, 221<br />

Poly(pphenyleneviny1ene) 229<br />

Poly(propy1ene) , isotactic 113<br />

Polystyrene 16. 20, 22<br />

Poly(tetrafluoroethy1ene) 102<br />

Polythiophene 229<br />

Poly(vinylch1oride) 118,158<br />

Potential energy distribution 95, 242<br />

Proteins 27<br />

Quadrature spectrum 5<br />

Rapid-scan FTIR spectroscopy 34<br />

Regularity bands 113<br />

Relative absorbance 46<br />

Reversible optical data-storage 67<br />

Rheo-optical elongation-recovery cycle 76<br />

Rheo-optical FTIR spectrocopy 75<br />

- variable temperature stretching machine 75<br />

Secular equation<br />

- for molecules 88<br />

- forpolymers 101<br />

Selection rules 109, 125<br />

Semicrystalline polymers 18<br />

Shear-induced orientation 69<br />

Side-chain liquid-crystalline<br />

copolysiloxanes 69, 72<br />

Side-chain liquid-crystalline<br />

polyacrylates 37<br />

Side-chain liquid-crystalline polyester 65<br />

- band assignment 65<br />

- irradiation-induced segmental alignment 66<br />

- specifically deuterated 65f<br />

Soliton 211, 213<br />

- conformational 197, 199


Step-scan FTIR spectroscopy 14,34<br />

- data collection sequence 44<br />

Stereoregularity 99<br />

Strain, dynamic 4<br />

Stress-strain diagram 77<br />

Structural absorbance 7. 36<br />

Synchronous spectrum 9f<br />

Tacticity bands 113<br />

Thermal erasure of laser-induced<br />

alignment 68<br />

fmris-cis-from Isomerization 76<br />

Transition dipole 3<br />

- coupling 244<br />

Twistons 197<br />

Two-dimensional correlation analysis 9<br />

Two-dimensional infrared spectroscopy 1<br />

Vibrational coupling 95, 97<br />

Vibrations<br />

- of chain polymers 100<br />

- of finite chains 124<br />

- of one-dimensional lattices 197<br />

- of tridimensional lattices 11 1


This Page Intentionally Left Blank

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!