Sunteți pe pagina 1din 16

5 MULTIDIMENSIONAL

SOLID-STATE NMR OF
POLYMERS
H. W. SPIESS
Max-Planck-Institut fur Polymerforschung, Postfach 3148, D-55021 Mainz,
Germany

5.1 INTRODUCTION
One of the key goals of materials science is the establishment of structure-
property relationships in order to improve known properties and to permit the
design of new materials. This holds in particular for synthetic polymers, whose
properties depend on both the molecular structure and the organization of the
macromolecules in the solid state: their phase structure, morphology, molecular
order and their molecular dynamics [1, 2]. Both macroscopic and microscopic
parameters are influenced by the processing that follows the chemical synthesis.
This calls for powerful analytical tools that can probe these aspects in the
material as it is used predominantly in the solid state. The structural aspects are
studied mostly by scattering techniques or by microscopy. Information about
dynamic aspects is deduced mainly from scattering or relaxation experiments [3].
Among these nuclear magnetic resonance (NMR) [4,5] is well established for the
structural characterization of liquids or compounds in solution, but much less so
for solids [6, 7]. Indeed, NMR offers numerous ways to study dynamic aspects
over a large range of characteristic rates. The main advantage of NMR is its
unprecedented selectivity. It is thus desirable to develop this technique for
studying the structure and dynamics of solid polymers [8]. However, owing to
the presence of angular-dependent anisotropic interactions, the spectral resol-
ution of solid-state NMR spectra is orders of magnitude lower than that of high
resolution NMR in liquids. Important improvements were achieved in the 1970's
by combining high-speed mechanical rotation of the sample with ingenious
manipulations of the nuclear spins, such as multiple-pulse irradiation, high-
power decoupling and cross polarization [9]. Moreover, two-dimensional (and
higher) NMR techniques have been introduced that offer fundamental advan-
tages [10]. First, as in liquids, the introduction of a new frequency dimension
provides a means of increasing the spectral resolution. Even more important,
Polymer Spectroscopy. Edited by Allan H. Fawcett
© 19% John Wiley & Sons Ltd
multidimensional spectroscopy also provides routes to new information that is
unavailable from one-dimensional spectra even in the limit of high resolution.
Experiments can be designed which correlate different spin interactions provid-
ing different structural information, or relate various states taken up by the
molecular unit during different time periods by exchange, and in this way to
probe dynamic processes in real time.
Progress in multidimensional solid-state NMR has been hampered by experi-
mental and conceptual difficulties, but these are now overcome [11-14]. This
chapter briefly outlines some of the main concepts and illustrates the information
available from multidimensional solid-state NMR spectra concerning polymer
structure and dynamics through experimental examples selected from the
author's laboratory.
Polymer dynamics is of central interest in our studies. Of the great variety of
molecular motions possible in polymers (e.g., translations, rotation, vibrations),
rotations have the most pronounced effects of NMR lineshapes and relaxation
parameters.Thus, multidimensional NMR provides essentially unique information
about rotational motions. Their timescales may be followed in real time over
many orders of magnitude, covering in particular that regime of slow motions
which govern the mechanical properties of polymers. Moreover, the higher-order
correlation functions provided by multidimensional NMR yield previously
inaccessible model-independent information about the geometry of rotational
motions, the orientational memory of molecular units involved in complex
dynamics, and the nature of nonexponential relaxation in disordered systems.
When the information has been collected for a variety of polymers, it should
eventually lead to a better understanding of their mechanical and rheological
behavior, which is of interest not only for conventional but also for new advanced
polymeric materials.
Two other aspects are particularly important for establishing structure-
property relationships for polymer materials, namely chain alignment in partially
ordered systems and domain sizes in heterogeneous polymer materials. The
orientation of macromolecular units is used to improve the properties of poly-
mers for such diverse applications as high-tensile-strength fibres and nonlinear
optical materials for information technology. Advanced polymer materials al-
most inevitably consist of more than one component, which often leads to phase
separation. Careful design of the molar ratios as well as the size, composition and
morphology of the different phases offers a means to control the mechanical,
electrical and optical properties. Small domains that extend over only a few
nanometers, and also interfacial regions between the different phases, are particu-
larly difficult to characterize. Major advances have been achieved in these areas
by introducing the concepts of multidimensional spectroscopy. The new solid-
state NMR techniques nicely supplement well-established scattering and micro-
scopic methods, as we demonstrate by various experimental examples and by
explicit comparison.
5.2 MULTIDIMENSIONALSOLID-STATE
NMR SPECTRA
Solid-state NMR exploits anisotropic, angular-dependent interactions of nuclear
spins with their surroundings [4, 5], in particular magnetic dipole-dipole
coupling of nuclei among themselves. This leads to broad NMR lines covering
« 5 0 kHz for 1 H - 1 H homonuclear coupling and « 2 5 kHz for 1 H- 1 3 C hetero-
nuclear coupling. The anisotropy of the chemical shift results in powder patterns
ocvering « 1 5 kHz at a field strength of 7 T. In addition to these magnetic
interactions, nuclei with spin / > 1/2 can also have electric quadrupole moments,
and are subject to quadrupole coupling to the electric field gradient at the nuclear
site. For 2 H ( / = 1) in C- 2 H bonds this leads to spectral splittings of «250 kHz.
Since C-H bonds are common in polymers, 2 H labeling is particularly useful.
If one of the above mentioned couplings dominates, either because of its
strength, or because the others have been suppressed by decoupling, the angular
dependence of the NMR frequency in high magnetic fields is alike for all
couplings, and is given by:
a) = coL + ±A(3cos2O- I - rjsin20cos2(t)) (1)
Here coL is the Larmor frequency, A describes the strength of the anisotropic
coupling: i.e. the anisotropic chemical shift or x3 C - l H dipole-dipole coupling for
13
C and the quadrupole coupling for 2 H, and rj is the asymmetry parameter
describing the deviation of the anisotropic coupling from axial symmetry
(O ^ Y] ^ 1). The angles 0, <f> are the polar angles of the magnetic field B 0 in the
the principal axes system of the coupling tensor. This in turn is often simply
related to the molecular geometry; i.e. the unique axis being along a bond
direction, e.g. dipole-dipole coupling: 1 3 C- 1 H bond, quadrupole coupling:
C- 2 H bond, or perpendicular to an sp 2 plane as for 13 C chemical shift tensors in
aromatic rings etc. Depending on the total spin involved, signals described by
Equation (1) and their mirror images with respect to coL may be superimposed,
and in powder samples the spectra for all orientations are added to yield the
powder lineshape (e.g., the Pake pattern for 2 H with spin / = 1).
Details of the experiments designed to record multidimensional NMR spectra
are not given here, as ample literature exists on the subject [10-13] and an
extended monograph is available [14]. However, basis knowledge of solid-state
NMR, as in Chapter 4, and of the concept of two-dimensional (2D) Fourier
spectroscopy, is needed to read this chapter. A 2D NMR spectrum is generated by
recording a two-dimensional data set following pulsed irradiation as a function of
two time variables, as shown schematically in Figure 5.1, and subsequent double
Fourier transformation. The development of the nuclear spin system in the
evolution period with incremented time tx at the beginning of the pulse sequence
provides the basis for the first frequency dimension Co1. The NMR signal is
detected in the detection period with time 12 at the end of the pulse sequence,
preparation evolution mixing detection

Figure 5.1 Time scheme of two-dimensional NMR [11]

providing the basis for the second frequency dimension co2. In the exchange
experiments described here, a variable mixing period of duration tm, during which
dynamic processes can take place, is inserted between evolution and detection.
The concept of 2D spectroscopy is readily extended to three and higher dimen-
sions by inserting additional evolution and mixing times.

5.3 EXAMPLES
5.3.1 INCREASE OF SPECTRAL RESOLUTION
Let us first consider experiments which increase the spectral resolution of solid
state NMR spectra in order to characterize polymer structure in terms of chemical
moieties on the basis of isotropic chemical shifts. This represents the most
important feature of standard NMR techniques, and makes up one or two
dimensions of many multidimensional experiments in which the chemical struc-
ture is correlated with other molecular properties such as mobility or order. In the
solid state, the tensorial nature of the chemical shift makes the NMR frequency
also depend on the orientation of the molecular unit under study with respect to
the applied magnetic field of the NMR spectrometer, cf. Equation (1). Since the
orientational dependence of the NMR frequency is comparable with or even
larger than the variation of the isotropic chemical shifts of different structural
units, the powder patterns resulting from the anisotropic chemical shifts overlap
severely in all but the simplest polymer structures. Thus, in many cases a quanti-
tative analysis is virtually impossible. The effect of the anisotropy can be removed
by rapidly spinning the sample about an axis inclined at the "magic angle" of
54.7° (magic angle spinning, MAS). However, this is accompanied by a loss of the
information about the molecular orientation, which is the basis of structural and
dynamic information typical of the solid state. Thus, experiments are desired that
retain this information without sacrificing spectral resolution.
For moderate spinning speeds in the range of a few kHz, the anisotropic
chemical shift is not "spun out", but leads to sideband patterns, from which the
anisotropies can be retrieved [15, 16]. However, in complex polymer structures
Figure 5.2 Separation of isotropic and anisotropic chemical shifts: (a) 13C MAS spec-
trum of an ether sulfone oligomer showing severe overlap of sidebands; (b) 2D 13C
sideband MAS spectrum of the same compound showing resolved sidebands [18]
the sideband patterns overlap heavily and hamper a quantitative analysis. By
ingenious spin manipulations through multiple-pulse sequences, these sideband
patterns can be removed from the spectrum in one dimension and retained in the
other [17, 18]. As shown in Figure 5.2, the crowded sideband patterns of
amorphous polymers can be nicely resolved in a second frequency dimension and
can then be exploited to provide structural and dynamic information.

5.3.2 SEPARATED LOCAL FIELD N M R

The dipole-dipole coupling, for instance between 1H and 13 C, or 1 H and 15 N,


provides valuable structural information. As the C-H bond lengths are known,
the measurements of dipolar splittings can be interpreted in terms of angles
between individual bonds and the applied magnetic fields. For proteins, measure-
ments of N - H bond lengths are of considerable interest, as they vary due to
hydrogen bonding and therefore contain important structural information. In
order to be useful, however, the dipolar patterns have to be separated according
to the chemically distinct sites in a molecule or monomer unit as identified by
their 13 C or 1 5 N chemical shits. As the dipolar couplings correspond to local
fields, such experiments are often named "separated local field" (SLF) experi-
ments [19]. Different schemes have been developed based on sample spinning
[20, 21]. If the polymer contains only carbons with a common chemical shift,
static techniques can be used [10]. As an example, Figure 5.3(a) displays the 13 C

Figure 53 Separated local field spectroscopy correlating heteronuclear dipole-


dipole coupling with chemical shifts: (a) ID * 3 Q 1 H spectrum of highly oriented poly(oxy-
methylene) (POM) with its order axis inclined at 30° with respect to the magnetic field
[22]; (b) orientation of chemical shift tensor deduced from this spectrum (principal axes
marked by <rn, <r22, <r33) relative to local POM structure
chemical ShIfV13C-1H dipolar powder pattern for oriented poly(oxymethylene)
(POM) [22]. From a quantitative analysis of such patterns, bond lengths and
bond angles can be determined, as well as the orientations of the principal axes of
the chemical-shift tensor in relation to structural units such as CH 2 groups,
Figure 5.3(b). Such information is needed for a quantitative analysis of other
experiments exploiting anisotropic chemical shifts.

5.3.3 WIDELINE SEPARATION EXPERIMENTS

Multidimensional NMR provides especially interesting information about poly-


mer dynamics. A long-standing method for qualitative characterization of mol-
ecular mobility is 1 H wideline NMR spectroscopy. There, large-amplitude
motions are detected through the reduction of the dipolar line width. However,
ID proton lineshapes leave many questions open, as they typically represent
superpositions of broad and narrow lines, and their relation to different struc-
tural units is often not obvious. In a straightforward combination of 1 H wideline
NMR, cross polarization (CP) and 13 C MAS spectroscopy in a 2D experiment
[23], it is possible to separate the dipolar patterns for the different structural units
(Wideline SEparation, WISE). This is demonstrated in Figure 5.4, where a WISE

Figure 5.4 2D WISE NMR specrating 1 H wideline spectra for different structural units
according to their 13C chemical shifts: (a) conventional 1 H wideline spectrum of a blend of
poly(styrene) (PS) and poly(vinylmethylether) (PVME); (b) 2D 1H 13C WISE NMR
spectrum indicating different mobilities of the two components [23]
NMR spectrum of a 50:50 wt% blend of poly(styrene) (PS) and poly(vinyl-
methylether) (PVME) is presented. The 1 H wideline spectrum, Figure 5.4(a),
consists of a rather featureless superposition of Components with different dipolar
linewidths, which are nicely separated in the second frequency dimension (Fig-
ure 5.4(b)) and related to structural units according to their 13 C chemical shifts.
Substantial motional heterogeneities, PVME being more mobile (narrower 1 H
lines) than PS, are detected despite the fact that the spectrum is recorded about
60 K above the caloric glass transition of this blend, which appears homogeneous
by most classical techniques. Such information about the mobility of the different
structural units is highly valuable for many practical applications.

5.3.4 2 D and 3D EXCHANGE N M R

For a thorough understanding of the chain motions in polymers, qualitative


information provided by 1H wideline spectra is not sufficient. In order to relate
chain motions to the structure of the polymer itself or to the packing of the
macromolecular chains, one requires knowledge about the geometry of the
motion, for example, the angles about which a molecular unit rotates during
individual motional steps. This information, which is hard to get otherwise, is
indeed provided by 2D exchange NMR as applied to rotational motions [12-14].
In simple cases this technique yields elliptical ridge patterns from which the angle
about which the molecules have rotated can be directly read off with a ruler [24].
As a specific example, Figure 5.5(a) displays such a 2 H 2D exchange spectrum for
poly(vinylidene fluoride) (PVF2), a polymer of considerable technological inter-
est because of its electrical properties. The geometric information from 2D NMR,
together with knowledge of the dipole moment change generated by the motion,
allowed us to identify the conformational change tgtg«—> gtgt of a chain defect as
being responsible for the mechanical and dielectric relaxation in the crystalline
regions of this polymer [25]. In a series of papers [26], 2D exchange NMR was
applied to study the chain motion of amorphous polymers above their glass
transition.
For complex motions, even the information accessible by 2D techniques is not
sufficient for an adequate description of the motional mechanisms involved in
chain dynamics. This is due to the fact that in 2D NMR the orientation of
molecules is measured only twice, in the evolution and in the detection periods.
Therefore, no information is provided about the trajectory a molecular unit
follows when rotating from one orientation to another. In order to distinguish
different mechanisms one has to determine the molecular orientation at least
three times. This is achieved in 3D exchange NMR. In a 3D exchange spectrum,
as displayed in Figure 5.5(b) for natural abundance 13 C in semicrystalline
poly(oxymethylene) (POM), different pathways pursued by the molecule lead to
different exchange signals, which can therefore be clearly distinguished. From
Figure 5.5 Multidimensional exchange NMR spectra elucidating slow molecular
motions: (a) 2D 2H NMR spectrum of crystalline poly(vinylidene fluoride) reflecting chain
motions through defect diffusion [25]; (b) 3D 13C exchange spectrum of oriented
poly(oxymethylene) reflecting helical jumps in the crystalline regions [14]
analysis of such 3D spectra for different semicrystalline polymers, which pack in
helical conformations, helical jump motions have been identified in which the
chain units rotate to neighboring positions and translate by one repeat unit [14].
This process can eventually lead to chain diffusion between crystalline and
amorphous regions with pronounced effects on the long-term mechanical prop-
erties [27].

5.3.5 CHAIN A L I G N M E N T FROM 2D A N D 3D NMR

Structural characterization of polymers often requires the determination of the


alignment of macromolecular chains. Orientation in polymers is often induced by
the production process and has strong effects on product properties. Through the
angular-dependent NMR interactions, orientation is amenable to measurement
in NMR experiments. In contrast to most classical techniques, the order of both
crystalline and amorphous components can be studied. For 2 H- or 13C-enriched
samples, one-dimensional experiments are sufficient to obtain orientation dis-
tributions in terms of a single angle [28]. In order to reconstruct two-dimensional
orientation distributions, or to resolve overlapping patterns in natural-abun-
dance 13 C spectra, multidimensional spectra are required. Two examples are
presented in Figure 5.6. Both involve 3D spectra of 13 C in natural abundance.
In the case of a biaxially drawn film of poly(ethylene terephthalate) (PET), the
orientational distribution was mapped out by flipping the sample between
different orientations with respect to the magnetic field (Direction Exchange with
Correlation for Orientation Distribution Evaluation and Reconstruction, DE-
CODER) [29]. The NMR signals of the different structural units of PET are
completely resolved in the 3D cube and their orientation distributions are thus
separately determined [30]. The chain axes are confined to the film plane
(full-width-at-half-maximum,fwhm, of 15°), whereas the in-plane distribution is
much broader (fwhm approx. 90°). The planes containing the phenylene rings and
the carboxyl group are oriented preferentially parallel to the plane of the film
(fwhm of 55°). In the case of a liquid-crystalline side-group polymer (Fig-
ure 5.6(b)), the extension to a third dimension is performed in order to achieve the
necessary spectral resolution [31]. Rotor-synchronized MAS is applied to map
out the degree of molecular order along Ox. The sidebands are separated by their
order along o2 (see Figure 5.2), and the chemical structure is probed via the
isotropic chemical shift along o3. Each dot in the 3D spectrum represents a single
sideband of a carbon position in the repeat unit with resolved 13 C chemical shift
in the MAS spectrum. This allows the molecular order of relatively complex
polymers to be analyzed quantitatively even for liquid-crystalline polymers,
where different moieties of the repeat unit exhibit substantially different degrees
of alignment [32]. The specific example studied here is a frozen smectic polyac-
rylate with a phenylbenzoate mesogenic side group and a spacer of six methylene
Figure 5.6 3D NMR spectra for determining molecular order: (a) 3D 13C DECODER
NMR spectrum of biaxially drawn polyethylene terephthalate) [30]; (b) 3D 13C MAS
spectrum of a liquid-crystalline side-group polymer [31]
units. The 3D MAS NMR spectrum not only reveals a pronounced order
gradient from the aligned mesogen to the disordered polymer chain, it also shows
that the acrylic carbonyl group exhibits a much higher order than the hydrocar-
bon units of the main chain. Because of its selectivity, high resolution solid-state
NMR is able to reveal that the carbon-carbon bond which links the acrylic
carbonyl to the main chain plays a crucial role in the decoupling of the ordered
side groups from the polymer chain [31].

5.3.6 D O M A I N SIZES FROM SPIN D I F F U S I O N


EXPERIMENTS

In heterogeneous polymers, domain sizes, or structures on the scale of up to a few


hundred nanometers, can be investigated. In NMR the proximity of molecular
units is probed by spin diffusion [4,14], which is most effective among protons.
Thus, NMR is particularly suited for characterizing small domains, nano-
heterogeneities or concentration fluctuations on length scales of a few nm, where
other methods often fail owing to liminations in resolution or contrast. An
advanced approach exploiting 1 H spin diffusion with highly sensitive 13 C
detection has been introduced recently [33, 34].
As a particularly clear-cut case, this technique has been applied to a series of
symmetric diblock copolymers of PS and poly(methyl methacrylate), PMMA
[34]. In Figure 5.7(a) the increase of the carbon signals of the phenyl ring in the
PS block after selection of the PMMA block is plotted against the spin diffusion
time for various block lengths. The equal lengths of both blocks in the symmetric
copolymers ensures a lamellar structure which makes the quantitative analysis of
the data easy. PMMA and PS are known to be immiscible. The spin diffusion
data yield domain sizes which are consistent with the scaling law M n 0 66 , where
Mn denotes the molar mass of the blocks, as predicted theoretically [35]. For
comparison, a statistical copolymer, in which no phase separation is possible, and
a blend of both homopolymers were included in the data set. The close agreement
between experimental intensities and the time dependence calculated from the
diffusion equation is apparent in Figure 5.7(b) and demonstrates that domain
sizes between 0.5 and 100 nm can be determined quantitatively. This technique
has already been applied to a number of homogeneous and heterogeneous
polymer systems, in particular to block copolymers containing both mobile and
rigid components affording detection of heterogeneities on a scale as small as
2 nm in systems that exhibit a single Tg in differential scanning calorimetry [36].

5.3.7 SPATIALLY RESOLVED SOLID STATE NMR

Eventually one would like to obtain the information about molecular structure
and dynamics accessible by solid-state NMR not just for the sample as a whole
(PMMA)

(PMMA)

ppm

toilms]
signal intensity PS

t m 1/2 [ms"*]

Figure 5.7 Spin diffusion as a tool for determining domain sizes in heterogeneous
polymers [34]: (a) 13C MAS spectra of the symmetrical diblock copolymer PS-b-PMMA
as a function of the diffusion time tm after selection of proton magnetization of the methoxy
group in PMMA; (b) signal intensity of the phenyl carbons in PS as function of tm for
different molecular weights. The numbers indicate domain sizes obtained from the fit
Figure 5.8 (a) Spatially resolved 2D 13 C spectrum of an injection-molded drawn tensile
bar of syndiotactic poly(propylene); (b) geometry of sample [37]

but with spatial resolution. For instance, one would like to distinguish the
molecular parameters in regions close to the surface from those in the bulk. In
fact, by applying pulsed magnetic field gradients, the concepts of multidimen-
sional NMR can also be used to generate a spatial dimension in 2D spectra
through Fourier imaging [10]; see also Chapter 6. As a first example, a spatially
resolved 13 C NMR spectrum of a drawn poly(propylene) sample is displayed in
Figure 5.8 [37]. It reflects differences of density and chain alignment between skin
and core due to the processing of the material, and demonstrates that spectro-
scopic solid-state NMR imaging is indeed possible. This exciting field is still in its
infancy and considerable progress is expected in the near future, as indicated in
a recent book on the subject [38] which is based on lectures at an international
conference.

5.4 CONCLUSION
As these examples indicate, a wealth of information about polymer structure and
dynamics is available through advanced multidimensional solid-state NMR
techniques. Profound correlations between macroscopic and molecular behavior
are emerging from such studies. Examples are helical jumps or defect diffusion in
crystallites, specific local motions in amorphous polymers and conformational
transitions coupled to relaxations of larger chain units in highly viscous polymer
melts. Such information has already been used successfully in the design of
polymers with improved properties. Thus, it can be anticipated that the import-
ance of multidimensional solid-state NMR as an advanced tool of polymer
spectroscopy will increase substantially in the future.

5.5 ACKNOWLEDGEMENTS
It is a pleasure to thank my coworkers engaged in the work described here: Drs.
B. Bliimich, B. Chmelka, J. Clauss, S. Feaux de Lacroix, E. Gunther, D. Schaefer,
JJ. Titmann, M. Wilhelm and, in particular, Dr. K. Schmidt-Rohr.

5.6 REFERENCES
[1] PJ. Flory, Principles of Polymer Chemistry, Cornell University Press, Ithaca,
1953.
[2] J.I. Kroschwitz (Ed.), Concise Encyclopedia of Polymer Science and Technology, John
Wiley & Sons, New York, 1990.
[3] N.G. McCrum, B.E. Read and G. Williams, Anelastic and Dielectric Effects in
Polymeric Solids, Dover Publishing, New York, 1967.
[4] A. Abragam, The Principles of Nuclear Magnetism, Oxford University Press, Oxford,
1961.
[5] CP. Slichter, Principles of Magnetic Resonance, Springer-Verlag, Berlin, 1980.
[6] F.A. Bovey, Nuclear Magnetic Resonance Spectroscopy, Academic Press, San Diego,
1988.
[7] CA. Fyfe, Solid State NMR for Chemists, C F . C Press, Guelph, 1983.
[8] R. A. Komoroski (Ed.), High Resolution NMR Spectroscopy of Synthetic Polymers in
Bulk, VCH, Deerfield Beach, FL, 1986.
[9] M. Mehring, Principles of High Resolution NMR in Solids, Springer-Verlag, Berlin,
1983.
[10] R.R. Ernst, G. Bodenhausen and A. Wokaun, Principles of Nuclear Magnetic
Resonance in One and Two Dimensions, Clarendon Press, Oxford, 1987.
[11] B. Bliimich and H.W. Spiess, Angew. Chem. Int. Ed. EngL, 1988, 27,1655.
[12] K. Schmidt-Rohr, A. Hagemeyer and H.W. Spiess, Adv. Magn. Reson., 1989,13, 85.
[13] H.W. Spiess, Chem. Rev., 1991,91, 1321.
[14] K. Schmidt-Rohr and H.W. Spiess, Multidimensional Solid-State NMR and Poly-
mers, Academic Press, London, 1994.
[15] M.M. Maricq and J.S. Waugh, J. Chem. Phys., 1979, 70, 3300.
[16] J. Herzfeld and A.H. Berger, J. Chem. Phys., 1980,73, 6021.
[17] A.C. Kolbert and R.G. Griffin, Chem. Phys. Lett., 1990,166, 87.
[18] S. Feaux de Lacroix, JJ. Titman, A. Hagemeyer and H.W. Spiess, J. Magn. Reson.,
1992,97,435.
[19] J.S. Waugh, Proc. Natl. Acad. Sci. USA, 1976, 73, 1394.
[20] T. Nakai, J. Ashida and T. Terao, J. Chem. Phys., 1988,88, 6049.
[21] J.E. Roberts, G.S. Harbison, M.G. Munowitz, J. Herzfeld and R G . Griffin, J. Am.
Chem. Soc, 1987,109, 4163.
[22] K. Schmidt-Rohr, M. Wilhelm, A. Johansson and W. Spiess, Magn. Reson. Chem.,
1993,31, 352.
[23] K. Schmidt-Rohr, J. Clauss and H.W. Spiess, Macromolecules, 1992,25, 3273.
[24] C. Schmidt, S. Wefing, B. Blumich and H.W. Spiess, Chem. Phys. Lett., 1986,130,84.
[25] J. Hirschinger, D. Schaefer, H.W. Spiess and AJ. Lovinger, Macromolecules, 1991,
24, 2428.
[26] S. Wefing and H.W. Spiess, J. Chem. Phys., 1988,89,1219; S. Wefing, S. Kaufmann
and H.W. Spiess, Ibid., 1988,89,1234; S. Kaufmann, S. Wefing, D. Schaefer and H.
W. Spiess, Ibid., 1990, 93, 197; D. Schaefer and H.W. Spiess, Ibid., 1992,97, 7944.
[27] K. Schmidt-Rohr and H.W. Spiess, Macromolecules, 1991,24, 5288.
[28] H.W. Spiess, in LM. Ward (Ed.), Developments in Oriented Polymers, 1, Applied
Science, London, 1982, p. 44.
[29] K. Schmidt-Rohr, M. Hehn, D. Schaefer and H.W. Spiess, J. Chem. Phys., 1992,97,
2247.
[30] B.F. Chmelka, K. Schmidt-Rohr and H.W. Spiess, Macromolecules, 1993,26, 2282.
[31] J J . Titman, S. Feaux de Lacroix and H.W. Spiess, J. Chem. Phys., 1993, 98, 3816.
[32] C B . McArdle (Ed.), Side Chain Liquid Crystal Polymers, Blackie, Glasgow, 1989.
[33] K. Schmidt-Rohr, J. Clauss, B. Blumich and H. W. Spiess, Magn. Reson. Chem., 1990,
28, S3.
[34] J. Clauss, K. Schmidt-Rohr and H.W. Spiess, Acta Polym., 1993,44, 1.
[35] E. Helfand and Z.R. Wassermann, in I. Goodman (Ed.), Developments in Block
Copolymers, 1, Applied Science, London, 1982, Chapter 4.
[36] W.Z. Cai, K. Schmidt-Rohr, N. Egger, B. Gerharz and H.W. Spiess, Polymer, 1993,
34, 267.
[37] E. Gunther, B. Blumich and H.W. Spiess, Macromolecules, 1992, 25, 3315.
[38] B. Blumich and W. Kuhn (Eds.), Magnetic Resonance Microscopy, VCH-Verlags-
gesellschaft, Weinheim, 1992.

S-ar putea să vă placă și