Sei sulla pagina 1di 651

Magnetic Resonance in

Colloid and Interface Science


NATO Science Series
A Series presenting the results of scientific meetings supported under the NATO Science
Programme.

The Series is published by lOS Press, Amsterdam, and Kluwer Academic Publishers in conjunction
with the NATO Scientific Affairs Division

Sub-Series

I. Life and Behavioural Sciences lOS Press


II. Mathematics, Physics and Chemistry Kluwer Academic Publishers
III. Computer and Systems Science lOS Press
IV. Earth and Environmental Sciences Kluwer Academic Publishers
V. Science and Technology Policy lOS Press

The NATO Science Series continues the series of books published formerly as the NATO ASI Series.

The NATO Science Programme offers support for collaboration in civil science between scientists of
countries of the Euro-Atlantic Partnership Council. The types of scientific meeting generally supported
are "Advanced Study Institutes" and "Advanced Research Workshops", although other types of
meeting are supported from time to time. The NATO Science Series collects together the results of
these meetings. The meetings are co-organized bij scientists from NATO countries and scientists from
NATO's Partner countries - countries of the CIS and Central and Eastern Europe.

Advanced Study Institutes are high-level tutorial courses offering in-depth study of latest advances
in a field.
Advanced Research Workshops are expert meetings aimed at critical assessment of a field, and
identification of directions for future action.

As a consequence of the restructuring of the NATO Science Programme in 1999, the NATO Science
Series has been re-orllanised and there are currently Five Sub-series as noted above. Please consult
the following web sites for information on previous volumes published in the Series, as well jiS details of
earlier Sub-series.

htlp://www.nato.inVscience
http://www.wkap.nl
http://www.iospress.nl
http://www.wtv-books.de/nato-pco.htm

-~­
~
I

Series II: Mathematics, Physics and Chemistry - Vol. 76


Magnetic Resonance in
Colloid and Interface Science
edited by

Jacques Fraissard
Universite Pierre et Marie Curie,
Laboratoire de Chimie des Surfaces,
Paris, France

and

Olga Lapina
Boreskov Institute of Catalysis,
Russian Academy of Sciences,
Novosibirsk, Russia

SPRINGER SCIENCE+BUSINESS MEDIA, B.V.


Proceedings of the NATO Advanced Research Workshop on
Magnetic Resonance in Colloid and Interface Science
St. Petersburg, Russia
26-30 June 2001

A C.I.P. Catalogue record for this book is available from the Library of Congress.

ISBN 978-1-4020-0787-3 ISBN 978-94-010-0534-0 (eBook)


DOI 10.1007/978-94-010-0534-0

Printed on acid-free paper

AII Rights Reserved


©2002 Springer Science+Business Media Dordrecht
Originally published by Kluwer Academic Publishers in 2002
Softcover reprint of the hardcover 1st edition 2002
No part of this work may be reproduced, stored in a retrieval system, or transmitted
in any form or by any means, electronic, mechanical, photocopying, microfilming,
recording or otherwise, without IlVritten permission from the Publisher, with the exception
of any material supplied specifically for the purpose of being entered
and executed on a computer system, for exclusive use by the purchaser of the work.
TABLE OF CONTENTS

Preface xi

Plenary Lectures

Analysis of slow motion by multidimensional NMR 3


B. Bliimich, S. Han, C. Heine, Roo Eymael, M. Bertmer, S. Stapf

Quantifying structural and dynamic disorder in ionically conducting solid solutions 15


Hellmut Eckert, Eva Ratai, Thorsten Torbrugge. Michael Witschas

14N MAS NMR spectroscopy. An instrumental challenge and informatory technique 43


H.J. Jakobsen, H. Bilds¢e, J. Skibsted, T. Giavani

NMR diffusion studies of molecules in nanoporous materials 57


1. Karger, F. StaUmach

NMR studies of the mesomorphism, structure and dynamics of some new pyramidic 71
liquid crystals
Z. Luz, R. Poupko, E.J. Wachtel, V. Bader. H. Zimmermann

Investigation of conformational changes of organic molecules sorbed in zeolites by 83


HR MAS NMR spectroscopy
J. Roland. D. Michel. A. Pampei

"Lighting up" NMR and MRI in colloidal and interfacial systems 97


A. Pines. J. W. Logan. M.M. Spence

Progress in high resolution solid state NMR of quadrupolar nuclei: applications to 107
porous materials and catalysts
M. Pruski, J.P. Amoureux. C. Fernandez

Applications of hyperpolarized 129Xe NMR spectroscopy to the study of materials 115


I.L. Moudrakovski. A. V. Nossov. V. V. Terskikh. S. Lang. E.B. Brouwer.
D. V. Soldatov,C.1. Ratcliffe. J.A. Ripmeester

NMR in colloid science with special emphasis on self-aggregating systems 123


OUe Soderman. Carin Melander, Magnus Nyden. Daniel Topgaard

Solid state NMR characterization of polymers and surfactant molecules as confined 139
to porous silica materials
Roberto Simonutti. Angiolina Comotti. Silvia Bracco, Piero Sozzmii

Characterisation of porous materials by NMR 155


J.H. Strange. L. Betteridge. M.1.D. Mallett

Keynotes and oral presentations

Investigation of radical pairs in micelles using spin polarization techniques 173


E.G. Bagraynskaya, N. V. Lebedeva. M. V. Fedin. R.Z. Sagdeev
vi

Monitoring ultraslow motions in organised liquids 185


F.A Grinberg

Characterization of mass transport and related phenomena in porous catalysts 197


and sorbents by NMR imaging and displacement NMR spectroscopy
I. V. Koptyug, L. Yu. Ilyina, A V. Matveev, R.Z. Sagdeev, V.N. Parmon

NMR spectroscopy contribution to the study of biomaterial mineralisation 209


AP. Legrand, B. Bresson, R. Guidoin, R. Famery, J.-M. Bouler

Multinuclear magnetic resonance characterization of solid catalysts and their 219


reactions in the adsorbed state
J.B. Nagy, P. Lentz, A Fonseca, F. Testa, R. Aiello, Z. Konya, I. Hannus, /. Kiricsi

Orbital order and orbital fluctuations in colossal magnetoresistive manganites. 231


An investigation with sSMn and 139La NMR
G. Papavassiliou, M. Belesi, M. Fardis, e. Dimitropoulos, M. Pattabiraman,
G. Rangarajan

The features of PFO NMR technique and some methodical aspects of its application 245
V.D. Skirda

129Xe NMR of adsorbed xenon used as a probe to study microporous solids 255
M.-A Springuel-Huet

The mechanism for ionic and water transport in nafion membranes from resonance 267
data
Vitali /. Volkov, Evgeny V. Volkov, Serge F. Timashev

An aggregation number-based definition of the ionization of a micelle 277


Barney L. Bales

Use of IH NMR imaging to study competitive adsorption of hydrocarbons in zeolites 285


J.-L. Bonardet, P. N'Gokoli-Kekele, M.-A Springuel-Huet, J. Fraissard

EPR study of photo-induced surface modifications of nanocrystalline Ti02 samples 297


J.M. Coronado, AJ. Maira, J.e. Conesa, J. Soria

Double resonance NMR study of mesoscopic interaction between surfactant and 307
silica-alumina during the direct synthesis of AlSBA-15 mesoporous solids
J.B. D'Espinose de la Caillerie, E. Haddad, A Gedeon

MR profiling of drying in alkyd emulsions: origins of skin formation 317


J.-P. Gorce, J.L. Keddie, P.l. McDonald

Catching a falling drop by NMR: correlation of position and velocity 327


Song-I Han, Siegfried Stapf, Bernhard BlUmich

Robust characterisation of flowing emulsions using regularisation and velocity- 337


compensating pulsed field gradient (PFO) techniques
M.L. Johns, K. Hollingsworth, G.M. Davies, L.F. Gladden
vii

Phospholipids' sera and mononuclear cells in acute leukemia, malignant lymphoma 347
and multiple myeloma-evaluation by 31p MRS in vitro
M. Kuliszkiewicz-lanus, B. Baczynski

Vanadium-51 3QMAS NMR and its application for the studies of vanadia based 355
catalysts
O.B. Lapina, P.R. Bodart, 1.-P. Amoureux

Simultaneous EPR and TPR study of the V-Ce-O catalysts redox properties 365
J. Matta, E. Abi-Aad, D . Courcot, A. Aboukai's

Static and dynamic NMR studies on cosmetic emulsions 375


J. Plass. D. Emeis

Water magnetic relaxation in superparamagnetic colloid suspensions: the effect 383


of agglomeration
A. Roch, F. Moiny, R.N. Muller, P. Gillis

A stray field imaging study of the drying process of precasting materials used in a 393
steel making converter
Koji Saito, Yoshitoshi Saito, Peter 1. McDonald, John Godward

The adsorption of polyelectrolytes to colloidal particles monitored by IH relaxation 403


of the solvent
B. Schwarz, M. SchOnhoff

I-D and 2-D double heteronuclear magnetic resonance study of the local structure 409
of type B carbonate fluoroapatite
H. Sfihi, C. Rey

Spatio-temporal correlations in gravity-driven and pressure-driven fluid transport 423


processes
S. Stapf, C. Heine, S. Han, B. Bliimich

Pulse gradient spin echo measurement of flow dynamics in a porous structure: 433
NMR spectral analysis of motional correlations
lanez StepiSnik, Ales Mohoric, Andrej Duh

Mesoporous transition metal aluminosilicas: incorporation of alkylphenothiazines 445


and their photoionization
Sunsanee Sinlapadech. Larry Kevan

99Tc NMR of technetium and technetium-ruthenium metal nanoparticles 455


V.P. Tarasov, Yu.B. Muravlev. N.N. Popova, K.E. Guerman

Size effects on the nuclear magnetic resonance of sodium metal confined in 469
controlled pore glasses
V. V. Terskikh, I.L. Moudrakovski. c.l. Ratcliffe. 1.A. Ripmeester, c.J. Reinhold.
P.A. Anderson. P.P. Edwards
viii

Poster Presentations

EPR investigation at 4K of ceria and Cu-Ce ox.ide under S02 and H2 atmosphere 479
E. Abi-Aad, 1. Matta, R. Flouty, C. Decarne, S. Siftert, A Aboukai's

Characterization of mesoporous materials by IH NMR 485


D. W. Aksnes, L. Gjerdaker, L. Kimtys

29Si and 27 Al MAS NMR study of alkali leached kaolinite and metakaolinite 491
N. Benharrats, AP. Legrand, M. Belbachir

Electron accepting properties of zr0 2 based catalysts studied by EPR of 497


paramagnetic complexes of probe molecules
M. V. Burova, A V. Fionov, AO. Turakulova

Characterization of organic nanoparticles synthesized in microemulsions by 2H NMR 507


F. Debuigne, L. Jeunieau, 1.B. Nagy

Study of the anomalous solubility behaviour of solutions saturated with fullerenes 513
by 13C_NMR
ADemortier, R. Doome, A Fonseca, 1.8. Nagy

EPR study of pol yamine copper complexes 519


N. Guskos, V. Likodimos, J. Typek, H. Fuks, S. Glenis, M. Wabia, L. Lin, E. Grech,
T. Dziembowska

Anisotropy of transverse IH magnetization relax.ation in strained elastomers by the 525


NMR-MOUSE®
K. Hailu, R. Fechete, D.E. Demeo, B. Blumich

Synergy phenomenon in bulk ruthenium-vanadium sulfides: Sly NMR 531


and ESR studies
R. Hubaut, A Rives, 0. Lapina, D. Khabiiulin, C.E. Scott

Correlation of Sly NMR parameters with local environment of vanadium sites 537
D.F. Khabibulin, AA Shubin, 0.8. Lapina

Solid-state NMR studies of mesostructured alumino-phosphates: structure and 545


dynamics of the inorganic network and of the organic component
Yaroslav Z. Khimyak, Jacek Klinowski

Dispersion in porous bead packs studied by velocity exchange spectroscopy 553


Alexandre A Khrapitchev, Siegfried Stapf, Paul T. Callaghan

Alumina and zeolites as catalysts for decomposition and transformation of 559


chlorofluorocarbons studied by multinuclear NMR methods
Z. K6nya, l. Hannus, P. Lentz, J.B. Nagy, I. Kiricsi

Solid state NMR studies and reactivity of Silica-supported 12-tungstophosphoric acid 565
Wenxing Kuang, Alain Rives, Michel Fournier, Robert Hubaut
ix

Molecular motion in crosslinked polymers as studied by stimulated spin echo 571


T.P. Kulagina. G.E. Karnaukh. G.T. Avanesyan. F. Grinberg

Structural changes in Zr02 catalyst doped with Fe and Cu. EPR study 577
M. Labaki. J.-F. Lamonier. S. Siffer.t. E.A. Zhilinskaya. A. Aboukai's

IH-NMR-investigation of the phase transition of thermo-reversible polymers in 585


solution and at interfaces
A. Larsson. D. Kuckling. M. SchOnhoff

The dilute to semi-dilute transition in petroleum colloids. as studied by IH NMR 591


relaxometry and viscosimetry
D. Mastrojini, L. Barre, S. Gautier. E. Richard

51V MAS NMR studies of inorganic vanadates with very small and large chemical 597
shift anisotropies
Ulla G. Nielsen, Hans J. Jakobsen, J¢rgen Skibsted

iH• 13C. and i29Xe NMR study of changing pore size and tortuosity during 603
deactivation and decoking of a naptha reforming catalyst
X.-H. Ren, M. Bertmer. H. Kuhn, S. Stapf, D.E. Demeo, B.Bliimich. C. Kern, A. Jess

The EPR and electronic spectroscopy study of polyarylene-sulfophthalides and 609


polydiphenylenephthalide reduction by alkali metals
N.M. Shishlov, V.N. Khrustaleva, Sh. S. Akhmetzyanov. N.G. Gileva. V.S. Kolosnitsyn,
O.G. Khvostenko

195pt NMR-Fourier spectroscopy in the analysis of the mechanism of the cytostatic 615
activity of platinum complexes
V.E. Stefanov, A.A. Tulub

Nitroxyl radicals as spin probes for the study of Lewis and Bronsted acid sites 625
of oxide catalysts
O.Yu. Ovsyannikova, A. V. Timoshok. A.M. V%din

Porous structure of cellulose fiber walls studied with NMR diffusometry 631
Daniel Topgaard. Olle SiJderman

Alternation in the structure of water-in-oil microemulsions by the action of 637


poly(ethylene glycol)
N.N. Vylegzhanina, B.z. Idiyatullin, Yu.F. Zuev, V.D. Fedotov

In search of the nature of the anisotropic diffusion in nervous tissue - 643


MR microscopy of the excised rat spinal cord
W.P. W~glarz, A. Hilbrycht. D. Adamek, 1. Pindel. A. Jasinski

Structural factors in micellar catalysis: NMR self-diffusion study 649


Yu.F. Zuev, B.Z. Idiyatullin, V.D. Fedotov. A.B. Mirgorodskaya, L. Y.A. Zakharova.
L.A. Kudryavtseva

Index 655
Preface

Colloids and interfaces playa crucial role in modern research and technology. Among
the various approaches used to study these systems, the techniques of magnetic resonance, in
particular nuclear magnetic resonance (NMR) and electron paramagnetic resonance (EPR),
have a special place because of their unique ability to elucidate structure, dynamics and
function at the atomic and molecular level. Traditionally, discussions of colloids and
interfaces and their study by magnetic resonance do not take place under a coherent, unified
organisational framework; rather, they occur within specialised colloquia devoted to the
disparate subject areas defined by different systems of colloids and interfaces, for example
polymers, membranes, zeolites, etc. Such a separation is a pity since many common elements
and themes of these systems remain unrecognised and unexploited. The purpose of this
Advanced Research Workshop on "Magnetic Resonance in Colloids and Interface Science"
had an interdisciplinary objective, namely to bring together the most active researchers in the
broad, general area of colloid and interface science using NMR and EPR. Such an
interdisciplinary workshop provided an opportunity for researchers to learn about the broad
applicability of advance NMR and EPR techniques and to appreciate the power of advanced
new methodologies for the study of molecular structure and dynamics and interfaces, thereby
helping to bridge the gaps between the various disciplines under investigation by magnetic
resonance.

This meeting should be considered as a "mise au point" of the combined state of the art,
illustrated by applications of innovative NMR and EPR techniques, including magnetic
resonance imaging (MRI) and microscopy, to the field of gas-solid and liquid-solid interfaces,
organic and biological surfaces, microemulsions, liquid crystals, membranes, structure and
dynamics of polymers and of micellar systems, and diffusion in heterogeneous systems.
Attendants of the workshop have been stimulated by the interdisciplinary vitality of the
research and got benefit from exposure to an important domain of the physical, chemical and
biomedical sciences.
Many people have contributed to the success of the ARW on which this volume is
based. We thank of course all participants for contributing to the intellectual dialogue. It is
xi
xii

also a great pleasure to acknowledge the main financ;ial support provided by the Scientific
Affairs Division of the North Atlantic Treaty Organization (NATO) .
We thank also the
- International Association for the Promotion of Cooperation with Scientists from the -
- New Independent States of the former Soviet Union (INT AS),
- Centre National de la Recherche Scientifique (CNRS, France),
- Universite Pierre et Marie Curie (Paris, France),
- Boreskov Institute of the Catalysis (Novosibirsk, Russia),
- Russian Foundation of Basic Research (Moscow, Russia),
- Bruker Analytik GMBH, (Germany)
- St.Petersburg Scientific Centre,
- French Ambassy at Moscow,
- Meriya ofSt.Petersburg

Many individuals whose great help with the organisation of the ARW we acknowledge with
gratitude include V.V.Lunin (Moscow), A.AShubin, L.Ya.Startseva, V.V.Terskikh,
G.M.Zhidomirov, I.E.Beck, K.P.Bryliakov, D.F.Khabibulin from Novosibirsk, and
G.V.Nikitin, M.V.Matetskaya, N.B .Matetskaya from St Petersburg
Plenary Lectures
ANALYSIS OF SLOW MOTION BY MULTIDIMENSIONAL NMR

B. Bllimich, S. Han, C. Heine, R. Eymael, M. Bertmer, S. Stapf


Institute for Technical Chemistry and Macromolecular Chemistry
RWTH, 0-52056 Aachen, Germany

Slow translational motion is conventionally probed by magnetic gradient fields which are
manipulated in time so that different moments of the gradient modulation function are either
adjusted to zero or stepped through a range of values for subsequent Fourier transformation
to obtain the displacement propagators, and probability densities of motional parameters
like position, velocity and acceleration. It is shown, that this formalism is not restricted to
time-dependent linear fields to probe translational motion, but can be applied to time-
dependent offset fields in general with arbitrary parameter dependences including angular
dependences to probe rotational motion. Three cases are considered in particular: a linear
space dependence, a quadratic space dependence, and an angular dependence of the offset
field following the second Legendre polynomial. Experimental examples concern position
exchange NMR of laminar flow through a narrowing pipe and velocity exchange NMR for a
hollow fiber filtration module with pulsed linear fields, laminar flow thi-ough a pipe in a
time-invariant parabolic field profile, and I3 C solid-state exchange of dimethyl sulfone in a
homogeneous polarization field.

1. Introduction

Pulsed field gradient NMR is a well established method to study effects of translational
molecular motion [I - 7]. Gradient echoes are generated by a sign inversion of the gradient
field . An effective sign change of the space dependent polarization field can also be
achieved by 180 0 pulses, which is well known from the Hahn echo [8]. In fact, the
sensitivity of multiple Hahn echoes to translational molecular motion has been observed by
Carr and Purcell [9]. The common formalism for describing the effects of translational
motion in gradient fields is the moment analysis of the time-dependent gradient field [1 - 5].
Usually the somewhat misleading terminology offield gradient NMR is used, which directs
the attention to space-invariant linear fields, although the moment analysis is more general.
In fact, it can be applied to any space or angle dependence of the polarization field in the
rotating frame, which is called the offset field or sometimes the fictitious field [10, 11]. In
the following the moment analysis is applied to describe the time dependence of the
precession phase in the rotating frame for offset fields with a linear and a quadratic field
dependence as well as with an angular dependence (Fig. 1). The resulting moments are the
Fourier conjugates to motional parameters like initial position, velocity, and acceleration in
translational (zo, vo, ao) and rotational (P, aplat, &plat2) space as well as combinations of
3
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 3-14.
© 2002 Kluwer Academic Publishers.
4

PFG NMR NMR-MOUSE solid-state NMR

*z -1Lz
Gzz Fzz Z2 -Q =- ~(3COS2P-1)

off-set fields
~cosp
Boft <p(t) =-y j B.m df /
on/off - +/-
unit function


0

•t = -y [ jG,(f) df Zo z~ n(~)1

10 0
+ jF,,(f) df + fEet) df
0 0 o /00

•t + fG,(f) f df v., + fF,,(f) f df 2zoVo, + fE(f)fdf an £Q.I


0 0
o a~ at ",0

+ + ... ) +
Figure 1. Offset fields and their dependence on space z, Z2 or angle ~(top), examples of time-dependences of the
offset fields which can be introduced by use of 1800 pulses (left), and time dependence of the magnetization phase
<p (center and bottom right). Here t1n denotes the anisotrpy ofthe spin interaction in frequency units, and E is a unit
function which changes sign upon application of a 1800 pulse and goes to zero during the action of a z filter.

these variables like Z02, 2 ZOVo, and v02 + zoao. The formal treatment is illustrated by
experimental examples from flow NMR and solid-state spectroscopy.
Fourier-NMR experiments which aim at the measurement of motional parameters by
scanning the respective moments can readily be conducted in a multi-dimensional fashion in
terms of multi-dimensional exchange NMR. Along the ID diagonals of the Fourier
transforms of such exchange spectra, moments of different orders are traced. These
diagonals provide probability densities of motional parameters like position, velocity and
acceleration, which can also be measured in 1D experiments [10, 11]. Applied to solid-state
exchange NMR, the moment description complements the established analysis by Wefing
and Spiess [12, 13].

2. The phase of the transverse magnetization

The phase q> of the transverse magnetization of spins ~ in the rotating frame is given by the
time integral oft Q(t') dt' of the offset frequency Q = -y (B z - Bo) = - Y Boff' where Bz is the
polarization field in the laboratory frame and for simplicity Bo is defined by the frequency
Wrf of the rotating frame as Bo = - wrriy. For magnetic fields inhomogeneous in z direction,
the offset field is expanded into a Taylor series,

where the origin of z has been chosen to coincide with the origin of the offset field.
A time dependence of this expression results from two sources: the Fourier
coefficients can be time dependent from time-dependent currents in the field generating
coils (gradient coils including shim coils) and from effective sign changes due to application
5

of 180 0 pulses, and the spin position z may be time dependent from coherent flow or
Brownian motion in fluids, rotating solids or granular flow. If the offset field is defined by
the orientation of the magnetic field in the principal axes frame of a coupling tensor in solid
state NMR for linear interactions of spins ~, it can conveniently be manipulated through
sign inversion by 180 0 pulses (for higher order spin systems the discussion needs to address
different flip angles and the Hamiltonian instead of the offset field) .
Given the path of motion, the time dependence of position and angle can be
accounted for again by Taylor series expansions, leading to the phase terms listed in Fig. I.
For reorientational motion an angular dependence following the second Legendre
polynomial is considered for simplicity. The pulse sequences on the left side of Fig. I depict
the effective time behaviour of the offset field which can be introduced in NMR
experiments by means of 180 0 pulses. As a result, the Taylor expansion coefficients G" and
Fzz change their signs whenever a 180 0 pulse (or tWo 90 0 pulses forming a z filter) are
applied, and so does the angle-dependent offset frequency represented by the unit function
E. By variation of the duration of the pulse sequences the different time moments of E, F,
and G can be scanned so that by Fourier transformation of the measured signal the
probability densities of the Fourier conjugate variables of position and angle as well as of
position change and angle change, etc., can be obtained. Clearly, gradient fields can be
pulsed on modem spectrometers, and possibly the offset frequency be switched rapidly, so
that other schemes for phase encoding of these moments may be employed. This is true in
particular in imaging and flow NMR where pulsed gradient fields are standard. But neither
the approximately quadratic fields encountered in single-sided NMR like the NMR-
MOUSE® [14] nor the offset frequency in solid-state exchange NMR are traditionally
modulated in time other than effectively by rf pulses.

3. Measuring Flow Velocities in Time-Invariant Fields

Following Fig. I, the phase evolution for translational motion in z direction can be described
in terms of Fourier conjugate variables for linear fields

<pet) = kz Zo + qvz VOz + Ez ao z + ... , (2)

and for quadratic fields

<pet) = Kz z02 + ~zz 2 Zo VOz + Szz (vo/ + Zo aoz ) + .. . . (3)

Here kz is the wave number in z direction familiar from NMR imaging, and qvz is the Fourier
conjugate to velocity. It is proportional to the wave number qv for displacement familiar
from diffusion studies and scattering theory [3, 4, 10]. The definition of these and the other
Fourier variables Cz, Kz, ~zz, and Szz can be identified by comparison of eqns. (2) and (3) with
the expressions of the phase contributions given in Fig. I.
In linear fields the probability density of velocity is measured by acquisition of
transverse magnetization as a function of qvz> and in quadratic fields the distribution of
velocity times position is measured by variation of2 ~zz or the distribution of velocity square
for steady flow (ao z = 0) by variation of Szz. Experimental measurement schemes for phase
detection of these Fourier variables in time invariant linear and quadratic fields are the Hahn
6

B", =G, Z (a)

-~Z
--t (b) --t
Figure 2. Velocity measurements of laminar flow
TX~ through circular pipes in time invariant
I- t, - I inhomogeneous fields with a linear profile (left)
and a quadratic profile (right). a) Field profiles.
k. =0\ b) Pulse sequences and Fourier variables.
q ., =-1 /4 .,r 2;IE.
{1 c) Experimental data acquired by the Hahn-echo
technique on a high-field spectrometer (left) and
(c) the CPMG method for flow across the gap of an
NMR-MOUSE® (right) [10, 15]. d) Sampling
grids for acquisition of the Fourier transforms of
the probability densities of velocity in linear fields
(left) and velocity square in quadratic fields
(right).
2vm.. ..... v.. [em/s] K =0
~:= ·1 /4 "f F. f.
s(f.> (d) s(t.>

t '" tFT""" 'f, t I I IIIII1 r te

P(v,,) P(Vo2J

echo for qvz and 1;zz, and the double Hahn echo for Szz by suitable variation of the echo time
tE (Fig. 2b). This can readily be understood from the left column of Fig. I, where each sign
change of the offset field is achieved by a 180 0 pulse on the transverse magnetization. The
measurement schemes for acquisition of probability densities of velocity and velocity square
in linear and quadratic fields, respectively, are illustrated in Fig. 2. By variation of the echo
time fE, qvz should b~ sampled with constant increments on a time raster of tl and Szz on a
time raster of IE3 (d). The limitations of sampling on a logarithmic time scale are
encountered, when the decreasing echo-time increments interfere with the duration of the rf
pulses. For this reason the use of pulsed fields is far more favorable, because qvz and Szz vary
linearly with the field amplitudes Gz and F zz (cf. Fig. 2). Alternatively Fourier-
transformation algorithms with variable variable sampling grids can be explored.
Experimental data measured in time invariant inhomogeneous fields are shown in
Fig. 2c (left) for laminar water flow through a circular pipe. The box profile of the velocity
distribution P(v) = 21t r/(1t R2 dV/dr) = l/vrnax for 0 :s; v :s; Vrnax , where r is the radial variable
and R the diameter of the pipe, agrees well with the maximum velocity Vrnax = 2 <v>, where
<v> is the average velocity. The velocity distribution was measured in a linear field by
variation of tl in constant increments. In the quadratic field of an NMR-MOUSE, the
probability density of velocity square could not b~ measured, because of the more rapidly
decreasing echo time increments (Fig. 2d, right). However, an odd/even effect of the echo
amplitudes in a CPMG sequence was observed (Fig. 2c, right) [9], and quantitative average
flow velocities <v> appear to be accessible from the echo decays as a function of the echo
time. For laminar flow through a circular pipe in quadratic field profiles we observed
7

CPMG echo 2 CPMG echo 3

....,.
:i :::J
& &
Figure 3. Decays of echo amplitudes
:l ;l for the second (left) and the third
(right) echo of a CPMG train acquired
in the quadratic field profile of an
0,00 0,01 0,02 0,00 0,10 0,20 NMR-MOUSE® for laminar flow of
tE3 [ms3) tE2 [ms2) water through a 3 mm inner diameter
pipe. The initial decays are approxi-
700 mated by exponential functions with
6,5
600
.,.6,0
inverse decay time constants which are
~ proportional to average velocity and its
E
500 E 5,5
square (cf. eqns. 4 and 5).
~ 400 -:5,0
300
a:: 4,5
4,0
128 192 256 8 10 12 14 16
<V",>2 [cm2/s2) <v",> [cm/s)

experimentally, that the echo magnitudes as a function of the echo time IE can be
approximated in the initial regime by the following exponential decay functions [15],

odd echoes: Ip(~z)1 = IFTl{p(ZOVOZ)} I :::;exp{-/Ez/Rz}, (4)

even echoes: Ip(l;zz) I = IFTl{p(VO/)} I :::;exp{ - /E 3/R z'}, (5)

where, within experimental error, R z and R z' are proportional to <vo.,> and <vo.,>z,
respectively, and the time dependence follows from the dependence of the Fourier variables
~ and S on the echo time (cf. Fig. 2). This was observed for echoes 2 to 5 of a CPMG train,
Figure 3 provides experimental data for echoes 2 and 3. For a sufficiently small range of
average velocities <voz> such a linear relationship can indeed be expected based on a Taylor
expansion of the relaxation rate on <voz>. This observation indicates, that in spite of highly
inhomogeneous fields single-sided NMR may be suitable to collect quantitative information
about average flow velocities, while the measurement of velocity distributions will require
more sophisticated experiments, for example, multidimensional experiments,

4. Multidimensional Exchange NMR with Pulsed Gradient Fields

In gradient field NMR the offset fields (cf. Fig, 1, left) are usually pulsed. By means of
pulse sequences with the effective gradient fields given in Fig. I, probability densities of
position, velocity or displacement, acceleration, etc. can be measured [3, 10], However, the
gradient fields can also be varied independent of each other leading to encodings of k space
at different times Ii and the acquisition of joint probability densities of translational motion
(Fig, 4). For example, a 20 exchange experiment is obtained by varying two gradient field
pulses for encoding of kl and kz independent of each other. The condition kz = -kl for q-
space encoding is fulfilled on the secondary diagonal of the k-space exchange data matrix,
while along the diagonal kz = kl is valid, that is, along the diagonal average position is
8

Figure 4. 40 exchange NMR for translational motion.


a) Four pulsed gradient fields separated by mixing
times ImI. 1m2, and 1m) encode position at different times
by stepping through k space independently. b) Imposing
two boundary conditions, k2 = -k) and k4 = ok) reduces
the 40 position exchange experiment to a 20 velocity
exchange experiment. c) Imposing a third condition qv2
= -qv) results in a 10 experiment which measures the
probability density of velocity difference or
acceleration by stepping through E space. In the 20 q
space exchange experiment E appears along the
secondary diagonal, while average q appears along the
diagonal. d) Similarly q appears along the secondary
diagonal in a 20 k space exchange experiment and
average k along the diagonal (top right). For the 20
exchange experiments the effective modulations of the
offset field applicable to the primary and secondary
diagonals are indicated.

encoded, while displacement or, in the slow motion limit, velocity is encoded along the
secondary diagonal. This experiment has been published as POXSY for position exchange
spectroscopy [10, 11, 16]. In the 2D k space frame the spectrum after 2D Fourier
transformation is interpreted as a joint probability density, while in the frame rotated by 45°
the joint probability density is transformed into a position-velocity correlation map.
By introduction of boundary conditions, the dimensionality of the k space exchange
experiment is reduced and slices through the original multidimensional k space are defined.
For the 4D exchange experiment, particularly interesting conditions are k2 = -kl and k4 = -k3'
The resulting 2D cross section defines the velocity exchange spectroscopy (VEXSY)
experiment [6, 17], where velocity is encoded on both axes in terms of qv = (kfinal -
kinitial) /))2 (Fig. 4b). Similar to 2D k space exchange, the joint probability density of initial
and final velocities obtained by Fourier transformation of the 2D q space data transforms
upon 45° rotation of the coordinate axes into a velocity-acceleration correlation map,
because the condition qv2 = qvl is valid on the primary diagonal which encodes average
velocity and qv2 = -qvl is valid on the secondary diagonal. The latter encodes the probability
density of acceleration according to the pulse sequence in Fig. 4d, where the Fourier
conjugate variable to acceleration is given by E = (qv,final -qv,initial) 11.12.
In a similar way the three-pUlse SERPENT experiment [18 -20) can be understood to
scan a 2D plane in 3D k space. The echo condition kl + k2 + k3 = 0 defines the 2D
SERPENT spectrum on the plane perpendicular to the principal diagonal of the 3D (k" k2'
k3) space (Fig. 5). It has been shown, that this SERPENT spectrum is closely related to the
VEXSY spectrum, taking into account that difference or differential coordinates are used for
time variables in VEXSY in the tradition of multi-dimensional NMR, while the coordinates
in SERPENT are sum or integral variables measuring the timing of the gradient pulses in
9

Integral time
coordinates

• I

.j + - - - :..l1 -----+ ~ + - - ..l2


.
---+~. differential time
coordinates

k, + k, + k," 0 Figure 5. Pulse sequence (top) and definition plane


(bottom) of the three-pusle SERPENT exchange
experiment in 3D kspace. The SERPENT plane is
shaded in grey. The coordinates of this plane are
obtained by rotation of the k3 axis of the (k " kl , k3)
coordinate system into the magic angle along the
principal diagonal. Note, that the parameter along
the principal diagonal is proportional to the Fourier
(3'~.11k,- ,n k, conjugate of average position.
I"-- ~:"

terms of durations from a common reference time. The three-pulse-SERPENT slice


perpendicular to the principal diagonal in 3D k exchange space is equivalent to a projection
along the direction of average position (rJ + r2 + r3)/3 in the 3D position exchange spectrum,
so that the remaining axes are labelled by position differences corresponding to
displacements or velocities.
Experimental examples of the POXSY and VEXSY experiments on flowing water
are given in Fig. 6. The POXSY experiment (left) of water flowing through a narrowing
pipe (top) [16] clearly demonstrates that the 2D joint probability density of initial and final
positions given a mixing time fm is equivalent to an average position-velocity correlation
experiment (bottom). For the narrow section of the pipe high velocities are observed and the
velocity distribution is wide (middle). For the wide section of the pipe, lower velocities are
observed and the velocity distribution is narrow.
The VEXSY experiment (Fig. 6, right) has been applied to study the function of
hemodialyzers, i.e. hollow-fiber filtration modules [21]. The hollow fibers accommodate the
blood flow and are washed from the outside in counter flow to remove the filtered
compounds (top). In the experiment water was used for both flows. Along the diagonal of
the VEXSY spectrum the velocity distribution is observed (middle). For negative velocities
corresponding to water flow within the hollow fibers it is the box function, while for the
counterflow in the channels in between the fibers a more complicated velocity distribution is
observed (middle). Along the secondary diagonal the distribution of velocity change or
acceleration is measured. The acceleration distribution is obtained from the VEXSY
spectrum by integration over the velocity dimension. Two different membrane materials
have been investigated, SMC (synthetically and modified cellulose) and SPAN (special
poly(acrylo nitrile». They show different filtration efficiency, which is demonstrated by
comparison of the VEXS Y spectra [21] as well as by comparison of the distributions of
acceleration (bottom). The interactions of the molecules with the membrane wall appears to
be enhanced for the SPAN membranes, where the acceleration distribution has a more
10

positon exchange veloc ity exchan g e

-
dialysate in
5 mm r-- v< 0 +
blood i~blOOd out
V>O~~
<v> = + v<O

y:
I S.7cmls dialysate out
203 _

~ y
~o.o
w
Figure 6. Experimental examples of
water flow through a narrowing pipe
,;-
by position exchange (POXSY, left)
t~= 11ms and through a hemodialyzer by
-20."'-_ _ _~______'
-20.3 0,0 20.3
velocity exchange (VEXSY, right)
v, [mm /s] spectroscopy.

POXSY frame :
jOint probability
position-
velocity ~.)[]t]

+
density correlation

-200 0 200
acceleration [mm/s1

Lorentzian shape as compared to the SMC material, where the shape is close to being
Gaussian. Trans-membrane flow depends on the pressure difference between both
reservoirs. Under regular operating conditions it is unidirectional and leads to an
asymmetric distribution of accelerations, which was not observed for the experimental
conditions given in this case.

5. Exchange NMR for Rotational Motion

Multidimensional exchange NMR is well established in solid state NMR [\3] including its
analysis in terms of joint probability densities [12]. However, the analogy to exchange NMR
of translational motion is known less well [10,17,22]. In particular an analysis of solid-
state exchange spectra in terms of distributions of angular velocity and acceleration is new.
This analogy, however, comes about naturally when exploring the similarities between
exchange NMR of translational motion and rotational motion (cf. Fig. I). The offset field
associated with the angle-dependent offset frequency is manipulated in the same way by
180 0 pulses as the offset fields arising from position differences in space-dependent fields .
Given that the amplitude of these fields cannot be varied, the moments of the offset fields
are scanned in solid-state exchange NMR spectroscopy by variation of evolution and
detection times (Fig. 7). The zero-order moment is scanned by stepping time in a linear
fashion (top), while the first moment should be scanned on a logarithmic time scale. It is
noted that scheme (b) is reminiscent of the so-called reduced 4D exchange NMR experiment
[23] and is equivalent to scheme (b) in Fig. 4, while scheme (a) of Fig. 7 is equivalent to
scheme (d) in Fig. 4. The parameters accessible by use of the pulse sequences in Fig. 7 are
11

exchange of angular positions


90° +90° +90° (e
rf ~~'----+"';"'~B_oJ._t)_dt_=0
Figure 7. Pulse sequences and offset fields for solid state
:0 !-tIE dotE exchange NMR spectroscopy. Evolution and detection times
i i : -time are varied to scan different moments of the offset field. In the

80ff I" "I 0 551-+1 0,"",00-00-0'--0t-<0HO,...O,...OHO,"",Oo---


conventional exchange experiment the zero-order moment is
scanned in both dimensions by linear variation of time using
phase encoding in the first dimension and frequency
encoding in the second dimension (top). The first moment
exchange of angular velocities should be scanned on a time scale linear in tE2 by pure phase
90° 180°+90° -90° 180° ; detection in both dimensions (bottom, cf. Fig. 2, left).

rf~~ Clearly, the 1800 pulses can be spl it into pairs of 90° pulses
forming z filters with additional mixing times to yield a 40
10 d-IE !IE ItE !-ttE j21E exchange experiment. Then the zero-order moments of the
Ii! ! i i --+time
offset field can be scanned with linear time increments in

BOff-9:t:'
I S.'I ~ i r/E
~lllIlIi
i
each dimension, and the 20 spectrum for exchange of
angular velocities is obtained as a 20 cross-section through
this 40 angular orientation exchange space.
r'a (I) dt =0 ~ BOff(t) dt =0
1
~ t off 2/E

Boff(t) tdt = 0

probability densities of angle ~ from the zero-order moment, angular velocity a~/at from the
first moment, and angular acceleration (j~/aP from the second moment (cf. Fig. 1).
A simple experimental example is given by the I3C 2D exchange NMR spectrum of
the 180 0 jump motion of dimethylsulfone which leads to a reorientation of the principal axis
of the chemical shielding tensor by 108 0 [24]. The approximately axial symmetry of the
coupling tensor leads to an unambiguous relationship between orientation angle f3 and offset
frequency Q which is given by the second Legendre polynomial Q = !l.d2 (3 cos 2f3 -1) (cf.
Fig. 1). Figure 8 depicts experimental (top) and simulated (bottom) spectra in the frequency
(left) and the angle (right) representations. The latter is obtained from the former by
transformation of the axes according to the second Legendre polynomial. Following the
interpretation of POXSY spectra (cf. Fig. 4d and Fig. 6, left), the distribution of the average

Figure 8. DC exchange spectra of


dimethylsulfone at 125 MHz and a temperature
of 280 K. The chemical shielding tensor of this
molecule is approximately symmetric. Top:
experimental data. Bottom: simulated data.
Left: frequency representation. Right: angle
representation.
12

experi ment simulation

~~
0' 45'
li=(fl ,+[l,)12
90 ' 0' 45'
li=(fl,+fl,)/2
90'
B,
Figure 9. Probability densities of average
orientation angle (top, left) and average
reorientation angle (bottom, left) derived from 20
exchange NMR of dimethylsulfone (cf. Fig. 8).
Right: the accessible angle difference f'4J= f32 - f31

~~
is the projection of the reorientation angle t:;.O on
to the direction of the magnetic field.

-50' O' 50' ·50' O' 50'


"'fl =fl,-fl, "I1=Jl,-fl,

orientation angle 13 = (/32 + f31)/2 is observed along the diagonal in the angle representation
(right) and the distribution of angle difference or angular velocity apia! ::::: (/32 - f31) I tm ,
where tm is the mixing time (cf. Fig. 7), is observed along the secondary diagonal. These
distributions are plotted in Fig. 9 with experimental (left) and simulated (middle) data. The
average angle distribution (top) assumes a complicated shape, because it is the projection of
the angle distribution on a sphere on to the z axis parallel to Bo (right). The average
reorientation angle distribution (bottom) consists of a sharp peak and a wide foot. The foot
contains the exchange signals, which build up with the correlation time of the jump motion
as a function of the mixing time [22], but the foot is hard to measure.
The data shown in Fig. 9 are obtained without resort to a model of the motion. The
jump-angle distribution on a sphere can be obtained with the help of additional information
such as a model of the molecular dynamics [25]. The distribution of angular acceleration
&plat2 requires the measurement of the VEXSY analogon, varying the first moment
independently in both dimensions by stepping the evolution and detection times on a time
grid linear in tE 2, or by scanning the second moment in a ID experiment (cf. Fig. 1., right)
on a grid linear in tl
At least in the latter case the same difficulties are expected as in
measuring the probability density of velocity square in time invariant quadratic field profiles
(cf. Fig. 2, right), so that this experiment is unlikely to be realized in this way on current
NMR spectrometers. However, the the VEXSY analogon can also be obtained as a 2D slice
through a 4D exchange spectrum, where each evolution time is stepped with constant time
increments. Then a linear time scale for data acquisition is obtained at the expense of a
higher dimensionality of the exchange experiment.

6. Summary

The effects of slow rotational and translational motion on the phase of the transverse
magnetization have been described. Particular profiles of the offset field have been
considered. These are a linear and a quadratic variation with space and a variation with
angle according to the second Legendre polynomial. Motional parameters corresponding to
initial position or orientation, velocity and acceleration can be measured both for position z
and angle f3 in the slow motion limit by suitable variation of moments of the offset field. If
the fields can be pulsed like in gradient field NMR, the moments scale linearly with field
amplitude. This is not so for quadratic field profiles given the hardware of conventional
spectrometers and for angle-dependent offset fields. In these cases the moments are scanned
13

by variation of evolution times. Only the zero-order moment is proportional to time in this
case and yields access to the square of position z and to the orientation angle p, respectively.
The time derivatives of position and angle are accessible in such fields by non-linear
variation of time to achieve a linear variation of the higher order moments of the offset
fields . Alternatively the zero-order moment can be scanned at different times in
multidimensional experiments, so that higher order moments appear along the diagonals,
avoiding non-linear sampling variations of evolution times by jncreasing the dimensionality
of the experiments, or the offset field may be scaled at fixed evolution times by chemical
shift scaling methods.
After Fourier transformation the signal amplitude in such experiments is described
by joint probability densities. The same is true for projections of such spectra to reduce the
dimensionality of the data set. Such projections correspond to slices in the acquired data set
before Fourier transformation. On the other hand a transformation of coordinates achieved
for example by rotation of the coordinate frame transforms the joint probability densities to
correlation spectra of dynamic variables like velocity and position in 2D POXSY NMR and
acceleration and velocity in 2D VEXSY NMR.

Acknowledgements: Continuous support of this work by DFG (Sonderforschungsbereich


SFB 540, Forschergruppe BI 231123-1) is gratefully acknowledged. This work also
benefitted from several enlightening discussions with Paul Callaghan and with Alexandre
Khrapitchev.

References

1) Stejskal, E.O., Tanner, lE. (1965) Spin Diffusion Measurements: Spin Echoes in the
Presence of a Time-Dependent Field Gradient, J Chem. Phys. 42, 288.
2) Karger, J., Heink, W. (1983) The Propagator Representation of Molecular Transport
in Microporous Crystallites, J Magn. Reson. 51 , 1.
3) Callaghan, P.T. (1993) Principles of Nuclear Magnetic Resonance Microscopy,
Clarendon Press, Oxford,.
4) Pope, J. M., Yao, S. (1993) Quantitative NMR Imaging of Flow, Concepts Magn.
Reson. 5, 281.
5) Seymour, J.D., Maneval, J.E., McCarthy, K.L., Powell, R.L., McCarthy, MJ. (1995)
Rheological Characterization of Fluids Using NMR Velocity Spectrum
Measurements, J Texture Studies 26, 89-101.
6) Callaghan, P.T., Codd, S.L., Seymour, J.D. (1999) Spatial Coherence Phenomena
Arising From Translational Spin Motion in Gradient Spin Echo Experiments,
Concepts Magn. Reson. 11, 181.
7) Fukushima, E. (1999) Nuclear Magnetic Resonance as a Tool to Study Flow, Annu.
Rev. Fluid Mech. 31,95.
8) Hahn, E. L. (1950) Spin Echoes, Phys. Rev. 80, 580.
9) Carr, H.Y. , Purcell, E.M. (1954) Effects of Diffusion on Free Precession in Nuclear
Magnetic Resonance Experiments, Phys. Rev. 94, 630.
10) BlUmich, B.(2000) NMR imaging of materials, Oxford Science Publications,
Oxford,.
14

11) Han, S., Bltimich, B. (2000) Two-Dimensional Representation of Position, Velocity


and Acceleration by PFG-NMR, Appl. Magn. Reson. 18, 101.
12) Wefing, S., Spiess, H.W. (1988) Two-Dimensional Exchange NMR of Po\Vder
Samples. I Two-time distribution functions, J Chem. Phys. 89, 1219.
13) Schmidt-Rohr, K., Spiess, H.W. (1994) Multidimensional Solid-State NMR and
Polymers, Academic Press, London.
14) NMR-MOUSE is a registered trademark of RWTH Aachen.
15) R. Eymael (2001) Methoden und Anwendungen der Oberflachen-NMR: Die NMR-
MOUSE, Dissertation, RWTH, Aachen.
16) Han, S., Stapf, S., Bltimich (2000) Two-dimensional PFG-NMR for encoding
correlations ofposition, velocity and acceleration in fluid transport, J Magn. Reson.
146,169.
17) Callaghan, P.T., Manz, B. (1994) Velocity Exchange Spectroscopy, J Magn. Reson.
A 106,260.
18) Packer, KJ. , Stapf, S., Tessier, J.J ., Damion, R.A. (1998) The Characterisation of
Fluid Transport in Porous Solids by Means of Pulsed Magnetic Field Gradient NMR,
Magn. Reson. lmag. 16, 463.
19) Stapf, S., Packer, KJ., Graham, R.G., Thovert, J.-F., Adler, P.M. (1998), Spatial
Correlations and Dispersion for Fluid Transport through Packed Glass Beads, Phys.
Rev. E 58, 6206.
20) Bltimich, B., Callaghan, P.T., Damion, R., Han, S., Khrapitchev, A., Packer, K.,
Stapf, S. submitted Two-dimensional NMR of velocity exchange: VEXSY and
SERPENT
21) Han, S., Stapf S., Bltimich, B. in press Fluid transport and filtration in a
hemodialyzer module by 2D PFG NMR, Magn. Reson. lmag..
22) Heine, C. (200 I) NMR von rotatorischer und translatorischer Dynamik, Dissertation,
RWTH, Aachen.
23) Schmidt-Rohr, K., Spiess, H.W. (1991) Nature of Nonexponential Loss of
Correlation above the Glass Transition, Phys. Rev. Lett. 66, 3020.
24) Favre, D.E., Schaefer, D., Chmelka, B.F. (1998) Direct Determination of Motional
Correlation Times by 1D MAS and 2D Exchange NMR Techniques, J Magn. Reson.
134,26 1.
25) Hagemeyer, A., Brombacher, L., Schmidt-Rohr, K., Spiess, H.W. (1990)
Reconstruction of Angular Distribution from Two-Dimensional NMR Spectra of
Powder Samples, Chem. Phys. Lett. 167,83 .
QUANTIFYING STRUCTURAL AND DYNAMIC DISORDER IN
IONICALLY CONDUCTING SOLID SOLUTIONS

New Results from Solid State NMR

HELLMUT ECKERT, EVA RATA!, THORSTEN


TORBRUGGE, MICHAEL WITSCHAS
Institut fur Physikalische Chemie, WWU Munster
SchlofJplatz 7, D-48149 Munster, Germany

1. Abstract

Crystalline and glassy solid solutions play an important role in tailoring solid
electrolyte materials to specific applications. Both iso- and aliovalent
substitution of the mobile ions can greatly affect the magnitude of the electrical
conductivity and the mechanism of ion transport in such materials, as a
consequence of the spatial arrangement and ordering of the mobile ions
involved. Solid State NMR techniques are uniquely suitable to address these
kinds of structural issues and, in addition, can provide a detailed
characterization of the ion dynamics involved. In particular, information
available from temperature-dependent studies of the dipole-dipole and nuclear
electric quadrupolar interactions, as well as the spin-lattice relaxation times,
lend themselves to quantifyable structural and dynamic information. In the
crystalline solid solution system (LiNb03kx-(W0 3)x the substitution of
pentavalent Nb by hexavalent W creates cation vacancies in the lithium
sublattice, producing a marked increase in cation mobility. A similar principle
obtains in the plastic-crystalline solid solution system (Na3P04)1.x-(Na2S04)x
(0 ~ x ~ 0.1) where the substitution of phosphate by sulfate ions creates
vacancies not present in the original lattice, thereby altering the mechanism of
cation transport. Peculiar transport anomalies related to isovalent cation
substitution are also well-known in the glassy state. Replacement of a mobile
alkali cation species by one of its homologues results in a dramatic reduction in
cation mobility ("mixed alkali effect", MAE). We have studied the structural
origins of the MAE by a combination of dipolar solid state NMR techniques
providing intricate information on the structural arrangement of both like and
unlike cation species in relation to each other. These results indicate that the
unlike cations are randomly distributed within the entire cation inventory and
support the view that the MAE is a consequence of inhibited cation transfer
among mismatched sites created by the glassy network.
15
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 15-41.
© 2002 Kluwer Academic Publishers.
16

2. Introduction

Order/disorder phenomena possess fundamental importance in the design and


function of solid state electrolyte materials. In many solid electrolytes, the ionic
conductive behavior may be tailored to special applications by using solid
solution systems. These may be either based on a crystalline lattice or be
entirely amorphous, and the ionic conductivity in them can often be controlled
and varied over many orders of magnitude by adjusting the chemical
composition. For developing a fundamental understanding of ion transport in
solids, it is important to establish such composition-:property relations and to
provide an explanation for them on a structural basis. Nuclear magnetic
resonance techniques are uniquely suitable for providing such information on a
local level in an element-selective, inherently quantitative manner. The
structural formation is contained in the internal solid state NMR Hamiltonian
[1 ],

H = Hz + Hes + HDbomo + HDbetero + HQ (1)

which summarizes the contribution of distinct anisotropic internal interaction


mechanisms that influence the NMR resonance in the solid state. Hz (the
Zeeman Hamiltonian) represents the basic interaction of the nuclear magnetic
moments with the externally applied magnetic field. H es , the chemical
shielding Hamiltonian arises from the magnetic screening effects produced by
the electrons surrounding the nuclei and is sensitive to chemical bonding
effects. H D , the dipolar Hamiltonian, describes the magnetic dipole-dipole
interactions between the nuclei whose resonance is being observed and
neighboring magnetic moments; for this interaction one has to distinguish a
homonuclear and a heteronuclear contribution, respectively. Finally, the
quadrupolar Hamiltonian, H Q, describes the interaction of the nuclear electric
quadrupole moment (exhibited by all those nuclei with spin quantum number>
1/2) with local electric field gradients generated by asymmetric electron
distributions around the probe nucleus. Each individual type of interaction
bears useful structural information, which can be extracted with the help of
either empirical correlations or ab-initio calculations. In particular, the dipole-
dipole interaction bears special structural significance, as the corresponding
coupling constant can be calculated from internuclear distances (or distance
distributions) in a straightforward manner.
Since the solid state NMR lineshape is influenced by all of the interactions
together (eq. (1», a direct evaluation of an interaction from Hamiltonian from
the static lineshape will be restricted to those exceptional cases, where this
interaction makes the dominant contribution. In all other cases, the use of
modem multipulse, multidimensional solid state NMR methodology is
necessary. During the past decade, a plethora of such selective averaging
17

techniques have been developed, which simplify the above Hamiltonian, by


averaging out certain interactions while maintaining others [2]. Furthermore, in
the course of two- and three dimensional NMR experiments, it has become
possible to separate the effects of certain interactions in different evolution
periods.
Due to this progress in recent NMR methodology, it has become possible to
characterize the structural aspects of disorder in crystalline and glassy solid
solutions in great detail. Furthermore, there are also powerful NMR approaches
for characterizing the mobility of the ions involved, over a wide-ranged
timescale. Atomic and molecular mobility occurring with correlation times in
the millisecond region produce stochastic averaging of the anisotropic static
interactions. Thus information on cation motion on the kilohertz timescale is
available from a temperature dependent measurement of static solid state NMR
spectra. Furthermore, temperature- and frequency dependent measurement of
nuclear spin-Iattic relaxation rates afford a comprehensive characterization of
the dynamics occurring on the megahertz timescale. During the past two
decades, such NMR techniques have been employed to study ionic mobility in
a wide variety of crystalline and glassy solid electrolytes [3,4].
In the present contribution, we will discuss some recent results from our
laboratory, regarding the structural and dynamic details of three different
crystalline and glassy solid solution systems. We will investigate systems with
both iso- and aliovalent substitution, and their effects on the spatial
arrangement and the dynamics of the mobile ions involved. The results offer
important insights into the mechanism of ionic transport in crystalline and
glassy solid electrolytes.

3. Aliovalent Substitution within the Crystalline LiNb03- W03 System

The (LiNb03)I_x-(W03)x phase field is characterized by the existence of a


stoichiometric compound LiNbW06 and a wide solid solution range 0 ~ x ~ 0.5
with the hexagonal LiNb03 structure [5]. With increasing value of x, the c-axis
parameter shrinks from 13.868 A to 13.677 A, whereas there is no
corresponding change in the a-parameter within experimental error. Aliovalent
substitution of W(VI) ions on the Nb(V) sites results in the concomitant
creation of cation vacancies in the lithium sub lattice. Furthermore, this
substitution produces a total of five possible lithium environments. In addition
to the unperturbed Li site with four Nb next-nearest neighbors, there are the
possibilities of forming other Li environments having either 3Nb,lW, or
2Nb,2W, or INb,3W or 4W sites in their next nearest coordination sphere. The
quantitative distribution of these lithium sites depends on their corresponding
energies with which they are stabilized in the lattice, and if there are no energy
differences then a simple statistical distribution would be expected (Scenario I).
18

On the other hand, a W(VI) ion on a Nb(V) lattice site represents a positively
charged defect which can be most effectively compensated by direct interaction
with a Li vacancy (i.e. a negatively charged defect). If this effect is
energetically significant and leads to maximized W -vacancy interaction, a
completely different lithium site distribution would be expected (Scenario II).
Finally, based on the pronounced influence the cation substitution has on the c-
parameter, one might envision a third scenario, involving a pairing of each W
substitution site with a lithium vacancy along the c-direction (Scenario III). As
shown in Figure I, all three scenarios produce different predictions in the
respective populations of the five lithium sites involved. Figure 2 suggests that
the various lithium sites can be differentiated by 6Li MAS NMR on the basis of
their isotropic chemical shifts [6]. A consistent line shape analysis succeeds on
the basis of three distinct lithium sites, conrresponding to the next-nearest
neighbor environments 4Nb; 3Nb,IW; and 2Nb,2W. These results rule out the
statistical distribution scenario I. Most significantly, the line shape
deconvolution allows a quantitative assessment and a detailed comparison
with predicted values for the different scenarios. Clearly, the pairing model of
Scenario III produces the best agreement with the experimental data. Thus, the
NMR data suggest, that the tungsten - vacancy interactions in these solid
solutions are significant, and result in a specific pairing arrangement, providing
a rationale for the experimentally observed decrease in the c-Iattice parameter.
A second solid solution effect in the (LiNb03)t.x-(W03 )x system is apparent
from the static variable temperature 7Li NMR data summarized in Figure 3.
Except for the x = 0 sample the figure reveals a successive reduction in the line
width (full width at half maximum) as the temperature is increased. The static
7Li NMR line width of these samples is dominated by homonuc1ear 7Li_7Li
dipole-dipole interactions. As the temperature is increased to values above
200-300°C, thermally activated lithium motion on the kHz timescale results in
a successive averaging of these interactions, producing motional narrowing
effects on the lineshape. As revealed in Figure 3, the extent of this narrowing
increases with increasing x, indicating a concomitant increase in the lithium
mobility. This result correlates well with the fact that with increasing
substitution level x the number of lithium vacancies is increased [6]. A similar
effect has also been observed in electrical conductivity measurements in the
solid solution system LiTa03 -W03[7]. Thus, the NMR results of this study are
an important demonstration of the general principle that aliovalent substitution
with concomitant vacancy generation in a crystalline lattice can have a marked
influence on the ionic mobility in solid electrolyte materials. Furthermore,
detailed models on the spatial arrangement of the vacancies and the substituent
atoms can be developed on the basis of detailed NMR line shape analyses.
19

Scenario I
100

80

60
~
40

20

0.1 0.2 X 0.3 0.4 0.5

Scenario II
100

80

60
~
40 3Nb 1

20 0
··2~~2W
.. ' .. '
A ••••
A
0
0 0,1 0,2 0,3 0,4 0,5
X

Scenario III
100

80

60
~
40

20 ..~.....
~ ••••• , 2Nb2W
•• .L:r ••••••••
0
0 0,1 0,2 0,3 0,4 0,5
X

Fig. I: Predicted lithium site distributions for the five lithium environments in the (LiNb03) ••
x(W03)x solid solution system for the three scenarios discussed in the text. The symbol denote

°
experimental results from 6Li MAS-NMR Iineshape devonvolution (see Figure 2).They denote
lithium sites with: 4Nb; 3Nb; 1W; 6 2Nb, 2W next nearest neighbors.
20

x=O.O x =0.3
Z.O 1.0 .0.0 -1.0 -Z.O Z.O
,!.! I,! I! I"" I ,!., I,.! , I t I !
.,! I 't"

x=0.1 x=O.4

Z.O
!
1.0
I,,! ! I".! I,,!,
.0.0 -1.0
,!",,!
-2.0
, Z.O
! t!!"
1.0 .().O
I.!,! I !! "
-1.0
I • • !
·Z.O
,I ! ,

x=O.2 x=0.5

Z.O ,
, I !
1.0
tI ! t "
-0.0
It, t
-1.0
! I
-Z.O
,,!,'!!
Z.O
1 •• , . 1 , !
1.0 .0.0
I!!!, I • • "
-II.0
t!
-Z.O
, t!., ,I

Fig. 2: 6Li MAS-NMR spectra of the (LiNb03)._.(W03). solid solution system. Gaussian
devonvolutions into three Li sites are included.
21

8000

8 3 • • • • •
••
Q

• •
Q
0

0

3

0 • • •
.. .
0 0
.
0 0 0 •
..'"
0 0
t, +
6000 '" '" .A '"
.A
+
.A
D 0
A A A A
'"
.....N ..
A 0

•• •
A


A

d D

....
..c:: •

+
"C:f • 0

'i • X=o • A

'..c::
GI

..J
4000
0 x • 0;1
" D

• X = 0.1S • A +
•0
D x=< 0.2
.A
X = 0.25
2000 .'" x .. 0.3

A .'
X" 0.4
"
A


• x'" 0.5
..•
I
00
100 200 300 400 500 600
T(°C)

Fig. 3 Temperature dependent 7Li NMR linewidths in the (LiNb03)1_x(W0 3)x solid solution
system.

4. Anion and Cation Motion in Crystalline (Na3P04)1-x-{NazS04)x Solid


Solutions

Among the large group of crystalline solid electrolytes there are a number of
excellent cation conductors in the plastic crystalline state. In general these
materials are high-temperature phases of simple ionic compounds such as
LhS04, LiNaS04, and Na3P04, where the anions are rotationally disordered.
These phases are of specific fundamental interest, as it has been suggested that
cation and anion motion in these systems are dynamically coupled with each
other {"paddle wheel mechanism"} [8]. Unfortunately, their high ionic mobility
cannot be exploited at near-ambient temperatures because the plastic crystalline
state is not stable below the phase transition temperature. For example, pure
Na3P04 undergoes a first order solid-solid phase transition at 598 K from the
high-conducting cubic modification to a low-conducting tetragonal structure
22

(LT -Na3P04) [9]. On the other hand, the plastic-crystalline state can be
maintained in Na3P04-Na2S04 solid solutions which are known to form over a
wide compositional region [10]. Whether the cubic phase is thermodynamically
stable or merely metastable for these solid solutions is not entirely clear at the
present time. However, current studies suggest that solid solutions within the
compositional region 0.05 < x ~ 0.25 seem to have long time stability upon
extended annealing [11]. DSC studies of these solid solutions indicate the
presence of an extended second-order phase transition region with an onset
temperature near 380-400 K, suggesting the onset of motional degrees of
freedom in the solid state in this temperature region [12]. The structural
consequences of sulfate substitution can be discussed in connection with Figure
4: In high-temperature Na3P04 the anions form an fcc lattice, with sodium
occupying all of the octahedral and tetrahedral sites, resulting in a fully
occupied lattice with no vacancies [9]. Aliovalent substitution of phosphate by
sulfate ions, however, creates new vacancies in the sodium sublattice, thereby
potentially altering the cationic transport mechanism. Furthermore, the more
extended region of stability offers a dynamical characterization of both cation
and anion motion over a much wider range of temperatures than possible in
high-temperature Na3P04.

7.423 A

Fig 4 Crystal structure of high-temperature a3P04 and the ( a3PO.)I _.(Na2S0.) solid solu tion
system_
23

4.1. DYNAMICS OF CATION MOTION

4.1.1. 31 P NMR Lineshape Measurements


To characterize the dynamics of cationic motion, various experimental NMR
aproaches have been exploited. First of all, the stationary 31p nuclei in the
sample can be used as a convenient probe of sodium ionic motion [12]. To this
end, Figure 5 summarizes the experimental 31p linewidth data for various solid
solution compositions. Since there is no chemical shift anisotropy contribution
owing to the Td point symmetry of the phosphate group, the 31p NMR static
linewidth L\ is dominated by the static 23Na}lp dipole-dipole coupling. As Na

o. __ Rigid lattice IineMdth 3600 Hz


o-~.o"4·¢'4.o.

-
N 3500 3500
I ~v \...
o & <>\ ¢ pureN~P04
<13000
\.'0 \.
~
v x=O.06 3000
o x=O.10
2500 o ~ \\ o x=O.25 2500
o o¢
2000 <J <>\ 2000

I
oo v: ~... PhaseTransl'ti on
1500 1500
DO ~ \

1000 ~
o ~ .. 'O"",~h,."
~ .........
¢ ..,-o-......<> ........<>
.Q .. .. 1000

500~--~~--~~--~~~--~~--~~~500
200 300 400 500 600 700 800
T/K

Fig. 5 lip NMR linewidth as a function of temperature in Na3P04 and in three


(NaJP04)I.x(Na2S04)x solid solutions,

diffusion becomes thermally activated the linewidth is gradually diminished


when the correlation time tc characterizing the cation motion becomes
comparable to L\' I. For tcL\ « 1 at sufficiently high temperatures this
contribution is completely averaged away, and the static 31p lineshape is finally
dominated by the homonuclear contribution between the 31p spins. At these
24

temperatures a constant Gaussian linewidth of 850 Hz is observed, which is in


excellent agreement with the van Vleck value calculated from the 31p_31p
internuclear distances within the structure. These results clearly reflect the fact
that although the phosphate anions undergo reorientational motion (see below),
the P atoms at their centers are stationary. Figure 5 illustrates further that the
motional narrowing observed for the 31p resonance due to the onset ofNa ionic
motion can be divided into two distinct temperature regions. Within the
temperature range 250 K ~ T ~ 380 K, the 31p NMR linewidth narrows only
gradually. The onset temperature Tc of this gradual narrowing effect (300 K)
can be exploited to estimate an activation energy using the Waugh-Fedin
expression [13]:

Ea (kJ/mol) = 0.156 TclK, (2)

yielding an estimated activation energy of 0.48 eV. At temperatures above the


region of the second order phase transition (380 to 400 K, the decrease of the
31p NMR linewidth observed with increasing temperature is much more
dramatic, suggesting that Na motion is greatly accelerated at temperatures
above the second-order phase transition. Also included in Figure 5 are data for
the low-temperature tetragonal phase ofNa3P04. In this compound, (which has
no second-order phase transition in this temperature range) much higher
temperatures appear to be necessary to activate sodium motion. Using eq. (2)
we estimate Ea = 0.69 eV [14]. The gradual 31p line-narrowing effect continues
over an extended temperature range spanning nearly 200 K. It is only near the
151 order phase transition temperature of 598 K that the 23Na dipolar field has
been completely averaged out at the 31p site.

4.1.2. 23Na NMR spin-lattice relaxation studies


A second approach to monitor sodium motion and to study solid solution
effects thereupon employs 23Na NMR spin-lattice relaxometry, where the
return (relaxation) of non-equilibrium spin state populations back to thermal
equilibrium is monitored. Relaxation is produced when magnetic or electrical
interactions caused by molecular or atomic movement fluctuate at the nuclear
Larmor frequencies, thereby inducing spin state transitions.
A frequently employed theoretical description by Bloembergen, Purcell and
Pound (BPP) [15] indicates that the spin lattice relaxation rate depends on the
spectral density components JI(ooo) and J2(200 0 ):

(3)

Here the relaxation strength C characterizes both the origin and the ,magnitude
of the fluctuating local field. For 23Na spin-lattice relaxation the electric field
gradient fluctuations usually produce the dominant contribution. For many
25
mobile solids the spectral density function can be decomposed into two or
more statistically independent motional processes (i) with distinct associated
correlation times. In such cases the experimentally measured spin-lattice
relaxation rate is the sum of the respective separate contributions. Since 'tc can
be assumed to exhibit a temperature dependence according to an Arrhenius
law,

'tc = 'teo exp(EJRT) (4)

the activation energy Ea(i) and the preexponential factor 'teo(i) characterizing
each motional process can be extracted from an analysis of temperature
dependent spin lattice relaxation data. Substitution of eq (4) into eq (3) results
in a plot of log T I-I vs. inverse temperature, revealing a relaxation rate
maximum for each motional process at that temperature where OOo'te(i) = 0.62.
At significantly higher temperatures, where OOo'te(i) « 1 (the "extreme
narrowing limit") T I-I becomes independent of Larmor frequency 000 , whereas
in the low temperature region eq. (3) implies that T I-I is proportional to 000 - 2,
provided the BPP theory is applicable. These considerations illustrate that
NMR spin lattice relxation rates are most sensitive to motional correlation
times in the micro- and nanosecond regime.
Figure 6 summarizes temperature dependent relaxation rate data measured at
79.4 MHz for Na3P04 and several solid solution samples with different sulfate
contents. Note that in all of the solid solution phases the relaxation rate curves
give evidence of two distinct contributions: one relaxation rate maximum
occurs uniformly at temperatures near 500 K (process 1)_ The sulfate content
appears to have only a minor influence on the position of this maximum, and
the TI values at the corresponding temperatures are compositionally invariant
within the limits. of experimental error_ A second relaxation rate maximum is
observed at higher temperatures, and the importance of this mechanism
increases markedly with increasing sulfate concentration (process 2). The
corresponding theoretical analyses of typical temperature- and frequency
dependent data sets are shown in Figure 7 for Na3P04 and a representative
solid solution sample[12]. The two activation energies are similar and agree
well with the values obtained by electrical conductivity measurements (0.43 ±
0_02 eV for pure Na3P04 and slightly lower values for the solid solutions [16].
Process 2 is almost absent in pure high-temperature Na3P04, and its relaxation
strength increases significantly with increasing sulfate concentration (see
Figure 8). Along with this trend, the temperature at which the relaxation
maximum is observed, decreases, indicating that the process becomes faster.
These results suggest that process 2 is a conventional. ionic hopping
mechanism, which proceeds via the vacancies created on the sodium sublattice
by the sulfate substitution process_ In contrast, process 1, which is essentially
26

T/K
1000 900 800 700 600 500
10'-~r-r-.-'--.--'---.----r----~----~~ 10

+++++
+++
++ + + 0
+
+
++
+
+
CJ) +
+
E +
+ o
+ 0
0 o o
/),.
+ 0 0
+ 0
o
f),.
o
0
+ 0 ¢
00 f),.

+
o 00 f),.
o
o o
<f>
o
o o 1
o
o
o
0 0
<:19 0
o~
of),. 00
o
o
o
o

"" Content Na2S04 :


x=O.25
""""
+

~ " 0 x=O.10
;Y" "
x=O.06
""""
0

't>-"" x=O.04
0.1 f),.
0.1
.;F " 0 x=O.02
,,~ x=O.OO
"
""

1.0 1.2 1.4 1.6 1.8 2.0 2.2


(1000 K) I T

Fig.6 23Na spin-lattice relaxation rates in (Na3P04)I.x(Na2S04)x solid solutions.


27

1000 900 800 700 600 T/K


Ii)
E
-; o 0

I-
0 79.4 MHz
~ 0.0 0 16 MHz 0 0.0

-0.5 -0.5

-1.0 -1.0

1.0 1.2 1.4 1.6 1.8


(1000 K) I T

T/K
1000900 800 700 600 500

U> 1.5 1.5


E 0 16.0 MHz
-; 0 79.4 MHz
I-~

0) c. T 1p(79.4 MHz)
..Q 1.0 1.0

0.5 0.5

0.0 0.0

1.0 1.5 2.0


(1000K) / T
Fig. 7: BPP analysis of frequency dependent 23Na spin-lattice relaxation rates in terms of two
independent processes for pure Na3P04 (top) and (Na3P04)09(Na2S04)OI solid solution (bottom).
28

o
.....

<>--------------------------~---
0.1
o _-------------------------0----
--

0.01
,,
,0

0.00 0.05 0 .10 0.15 0.20 0.25


X
NazSO.

N
!/)
N

iH
";' 2.0

f
0
':-
U
c:
o 2
1.5

1.0 D.] o C1

0.5 ¢

0.0
0.00 0.05 0.10 0.15 0.20 0.25
XNa,SO,

Fig. 8 Values of teo and C for processes I and 2 in (Na3P04)l .•(Na2S04). solid solutions as a
function of x. .
29
not influenced by the anion substitution might be attributable to a cation
motion process that is assisted by the anion reorientation process. Based on the
23Na relaxation data alone, however, no definite conclusions can be drawn on
the existence of the paddle-wheel mechanism and complementary
characterization of the anion motion is necessary.

4.2. DYNAMICS OF ANION MOTION

Data regarding the anion motional process have been obtained from 170 NMR
spectra of isotopically labelled samples. Temperature dependent 170 NMR
lineshape analysis has been carried out to study the details of anion
reorientation in a solid solution with ~ = 0.1 [17]. Figure 9 summarizes typical
spectra. At 135 K, the 170 NMR line shape is characterized by static 2nd order
quadrupolar perturbations, with a nuclear electric quadrupolar coupling
constant of 4.7 MHz. The quadrupolar interaction is dominated by the axially
symmetric electric field gradient arising from the electron distribution in the p-
O bond [18]. As the temperature is increased, characteristic lineshape changes
are observed indicating the onset of a three-fold rotation around the p-o axes.
Figure 10 compares typical experimental spectra [17] with those measured
recently in the low-temperature phase of pure Na3P04[19]. Also included in
Figure 10 are simulations based on a dynamic model involving reorientation
about the four Cr axes present, albeit with two distinct rates. Altogether, the
results suggest that the anion reorientation mechanisms in both compounds are
quite similar. Note, however, that the temperatures where similar lineshapes are
observed (reflecting similar reorientation rates) are significantly shifted with
respect to each other, indicating that at any given temperature the anions
undergo C3- reorientation at much much faster rates in the cubic solid solution
system than they do in the tetragonal LT-Na3P04 compound. The reason for
this difference is probably related to the excess free volume present in the cubic
phase. Furthermore, at temperatures above 450 K the 170 spectra are simple
Lorentzians, reflecting a more or less isotropic rotational diffusion of the
phosphate tetrahedra, quite similar to the situation in high-temperature
Na3P04[19]. Our recent results indicate that it is this rotational diffusional
motion that is dynamically correlated with sodium cation transport [14, 17, 19].
For high-temperature Na3P04, a more detailed discussion of this issue in
relation to our NMR results as well as additional dynamic characterization by
inelastic neutron scattering and conductivitiy spectroscopy has been published
recently [19], and the situation appears to be similar in the Na3P04-Na2S04
solid solution system [17]. Here, we simply wish to emphasize that in the
present system, the aliovalent substitution of phosphate by sulfate ions leads to
significant mechanistic and/or rate changes in the dynamics of both the cationic
30

and the anionic species, which can be identified and quantified using
temperature dependent NMR lineshape and relaxation analysis.

313K /~~
-~ ~

293K

270K

230K~
,[,--/,-

373K

160K

333K

i i I
400
I I
200
I
o .200 ·400 400 2~0
ppm ppm

Fig. 9: Temperature dependent 170 NMR spectra of a (Na3P04)09(Na2S04)OI solid solution.


31

a) d)

Q=OkHz

i
, , , ,
400 ax) 0 -ax)-400
. I I I I

400 ax) 0 -200-400


I

400 200 0 -200 -400


O/ppn O/ppn
01 ppm

e) f)

, , I , i • , i • •
• i • I • iii' i '
400 ax) 0 -ax)-400 400 ax) 0 -ax)-400 400 ax) 0 -200-400
O/ppn O/ppn
OIPPll

Fig. 10: Experimental 170 NMR spectra of (Na3P04)O.9(Na2S04)OI (left column) and LT-Na3P04
(middle column) in two different dynamic regimes. The right column shows simulated spectra on
the basis of C3 reorientation processes with two distinct rates n and Q. Note that given dynamic
regime is reached at significantly lower temperatures in the solid solution than in LT-Na3P04

5. Isovalent Substitution in the Glassy State: the Mixed Alkali Effect

The phenomenon of ion transport is not limited to crystalline materials but


equally common in the glassy state. As a matter of fact, glassy ionic conductors
have unique advantages for applications as electrolytes in solid state batteries:
(l) electrode/electrolyte interfacing is easier when using glassy materials, (2)
the composition of glasses can often be varied over wide ranges, allowing their
physicochemical properties to be tailored to specific application demands, and
(3) the magnitude of the ionic conductivity is often enhanced in the glassy
state, compared to crystals having the same chemical composition. This
32

conductivity enhancement is presumably related to the intrinsic disorder in the


glassy state, which provides a manifold of available target sites into which the
mobile ions can be transferred. A considerable amount of work has been
devoted to covalent oxide glasses, the structural organization of which is
sketched in Figure 11 .

.-.
Short range 0-. o-®
Medium range .-® ©-©

Fig. 11 : Sketch of the structural organization in covalent oxide glasses.

These glasses are based on a network structure formed by polyvalent ions (such
as Si(IV), B(IU) or P(V» that are linked by oxygen. The network is modified
by the addition of alkali or alkaline-earth oxides (network modifier species),
which create anionic sites (typically non-bridging oxygen species) whose
charge is being compensated by the cationic species that are located in the
network's interstices. In most alkaline oxide based glasses the modifier cations
exhibit considerable ionic mobility. There is an extensive literature
characterizing the ionic conductivity of glasses as a function of their
composition and attempting to relate ionic ally conductive behavior to specific
structural features [20]. From these studies, a number of general principles
have emerged:
33
(1) The ionic conductivity increases dramatically with increasing ion
concentration, and it depends also on the nature of the network former
species,
(2) there is some evidence £Or cation clustering, particularly at low alkaline
oxide concentrations, although the extent of clustering depends on the
respective network former species present and
(3) owing to their different ionic sizes and potentials, the sites occupied by
different alkali ion species are clearly different and can be uniquely
distinguished by EXAFS.

Isovalent substitution of one alkali ion species by a homologue, at a fixed total


cation content, leads to a dramatic reduction in ionic mobility and electrical
conductivity by several orders of magnitude. This observation, known as the
"mixed alkali effect" [21] has puzzled researchers for several decades. In the
past, various physical models have attempted to explain this behavior on the
basis of the spatial distribution of the cations in the glass network. For various
types of glass systems, the mixed-alkali effect has been considered a
consequence of preferred interactions ("pairing") among unlike cations within
clustered arrangements, with the implication that such preferred interactions
would pose impediments on cation transport in glasses [22]. In apparent
conflict with this concept, there seems to be some evidenc.e for "like-cation
segregation" in a number of mixed-alkali systems [23,24]. More recent
thinking is based on the concept of "site mismatch", [25] originating from the
EXAFS result mentioned above that each type of cation is located in its own
distinct local site [26]. The mixed alkali effect then originates from the
hindrance of alkali ions of one type to migrate to sites previously occupied by
alkali ions of another type, if the two types of cations are intimately mixed.
Random mixing of unlike cations has been suggested also in molecular
dynamics simulations [27] . During the past five years, more direct
experimental evidence concerning the spatial relationship between different
types of modifier ions has been accumulated on the basis of solid state NMR
data. For example, the 7Li and 23Na chemical shifts respond quite sensitively to
the presence of cation homologues, suggesting strong interactions between the
two types of cations [28] . A more quantitative NMR method probing ion-ion
interactions in mixed alkali glasses is based on the selective measurement of
the magnetic dipole-dipole interactions between the nuclear spins associated
with the two different kinds of cationic species present. According to van
Vleck theory [29], the direct contribution to the dipolar second moment, M 2,
for a nucleus dipolarly coupled to surrounding heteronuclei, can be calculated
directly from internuclear distance distributions:

MIS
2
= ~(llo)2
15 \411
./i 2y2y
I s
2S(S + 1). '" -6
LJ IS (5)
34

where YI is the gyromagnetic ratio of the observed nuclei and Ys and S are the
gyromagnetic ratio and the spin quantum number of the nonresonant nuclei
coupled to the observed nuclei. Figure 12 shows the NMR pulse sequence used
to determine M~s selectively. The experiment, called SEDOR ("spin echo
double resonance") [30] consists of two distinct parts. In the first part, the spin

1800

7\
Echo

1800

s D
Fig. 12 SEDOR Pulse sequence used to measure heteronuclear dipolar second moments.

echo intensity of nuclear species I is measured by a 90 0 -t\-180° Hahn-spin echo


pulse sequence as a function of the dipolar evolution time t). In this
experiment, the heteronuclear I-S dipole-dipole coupling is refocused and does
not contribute to the decay ofthe I - spin echo as a function oft\. In the second
part of the experiment, the Hahn-spin echo of the I spin species is recorded,
while also applying a 180° pulse to the S- spin species. Since this additional
pulse serves to reverse the sign of the I-S dipole-dipole coupling constant, the
heteronuclear dipolar coupling is now no longer refocused, and the decay of the
I spin echo as a function of t1 is accelerated accordingly, depending on the
magnitude of the dipole-dipole coupling. Mixed alkali glass systems containing
lithium and sodium as the mobile ions lend themselves particularly well to this
kind of analysis, because of the favorable NMR properties of the 6Li, 7Li, and
23Na isotopes. In general, the experiment is most conveniently carried out with
23Na as the observe (I-spin) nucleus and one of the lithium isotopes as the S-
spin nucleus (for 23Na{6Li} SEDOR experiments isotopic labelling is required).
For a multi-spin system the normalized decay 1(2t\)/1o of the 23Na spin echo
amplitude is typically observed to be Gaussian, and M2 Li-Na can be calculated
according to the simple expression [31]:

(6)
35
In this expression, F(2t))/Fo is the 23Na spin echo decay in the absence of the
1800 pulses applied to the lithium spins. Figure 13 shows the experiment
applied to the crystalline model compound LiNaS04, illustrating the respective
advantages and disadvantages of using 7Li and 6Li as the S (non-observe) spin
species. Owing to the larger gyromagnetic ratio of 7Li, the magnitude of the
SEDOR effect is larger than with 6Li. On the other hand, because of the large
first-order quadrupolar splitting of the 7Li resonance, only a fraction of the 7Li
spin states are inverted, thereby attenuating the effect below, the one expected
on the basis of eq.(6).

• 23 Na-Spin-Echo

1,0 " 23 Na - eli} SEDOR


~<><>
\ .'l>
-e:-
0,8
...-..<>
\ ...;:,.
-
-
~
0,6 \ ...~
\ ...."'.A.
0.4 ,
\
..
L;:;.,.

0,2
\'",""..,..."........ .
" ............ .
0,0
"
'"
" """
0,0 0,2 0.4 0,6 0,8 1,0 1,2 1.4

• 23Na-Spin-Echo
I \iNaSo4 1
1,0 " 23Na _ {6Li}-SEDOR

0,8

g
-- 0,6

0,4
,
A,6
,
, ,"
~ '-"
-
" "
0,2 "" " "
" " '--"
0,0 +--....--r--...---,.--~-~-~-=-:..::.:.,...-~
0,0 0,5 1,0 1,5 2,0
2 t 1 ' ms

Fig.I3: 23Na ~ 7Li ~ SEDOR (top) and 23Na ~ 6Li ~ SEDOR curves in LiNaS04. The dashed lines
are curves calculated from eqs. (5) and (6), based on the internuclear Na-Li distances in the
crystal structure. The dotted curve represents a SEDOR calibration based on incomplete 7Li spin
inversion because of the electric quadrupole splitting.
36

U12
1.0 • DHe-5p/n-Echo

.,
A D He. (\I) SEDOR
..~,
0.8 a.·t... , •••• D Na-Spin-Echo (fit)

'.
~
....... SzenIrioI
"4 .. ,
o
:::'0.6
"A ........
Szllllllo "

',. .
'.A. "

:r "."
e!. 0.4 '"
"
"'
.....
0.2 ······.A

O.O+---''---r-~--'-'''''''''-",---'--'---'
0.0 0.5 1.0 1.5 2.0
2 t,I rna

....
1.0 • 22 He-Spin-Echo
A 22 He • (\J) SEDOR
~ ,~
~ •••• 22 Na-Spft-Echo (fit)
0.8 ' "
.... f> ... , ....... Sztnlllol
". .'11- SztnIrioU
§;
_ 0.6 'A" <I\ "'.,

-
··A.. ",-=-:,_ _ _ _ _ _....1
.: ····4 "-""_
'. .
~0.4 ····4 ..........
"'
...... .....
0.2 ......... ,.
O.O+-~--r--....--.---..---.--..--;;:;::;--.
0.0 0.5 1.0 1.5 2.0
2t,/ms

U24
• 22 He Spin-Echo
1.0
A 22 He • (\J) SEDOR

0.8
.... 22 He s.*-Echo (F"d)
·······S~I
~II
S
:::' 0.6
••· •. A

~0.4 ........~
••••• A
.....
0.2 .••••••
" .....A
0.0 +---.--,...---.---r-~-:-.-::::::::;:=:::;..---.
0.0. 0.5 1.0 1.5 2.0
2t,/ms

Fig. 14: 23Na ~6Li~ SEDOR curves in [(Li 20) .(Na20)I.Jo3(B203) glasses. and comparison with
predicted curves for two scenarios (see text). Top x = 0.4, middle x = 0.6, bottom x = 0.8.
37

As a consequence, the M2 values determined on unknown systems need to be


calibrated against a reference compound. This situation is more favorable with
the 6Li spins, where the small quadrupolar splitting does not interfere with
complete spin inversion. Figure 14 displays typical experimental data on
[(Li20MNa20)\.x]o.3(B203)o.7 glasses [32] for three different compositions x.
The experimental data are compared with values calculated theoretically
assuming random lithium-sodium mixing within two distinct scenarios
describing the spatial distribution of the entire cation inventory of the glass.
Scenario I corresponds to a homogeneous distribution calculated via isotropic
dilution of a cubic lattice model, whereas scenario II shows the prediction
according to a statistically clustered distribution as calculated from a decimated
lattice model. The generation of both scenarios has been described previously
[33-35]. Clearly, our results are most consistent with scenario I and rule out
cation distributions models based on strongly clustered arrangements.

Na+: eo Lt eo
LI·: 100% u·, Na·= 1: 2

U· ' Na = 2 :1 Na· :100%

o o o ~O
o bj

Fig. 15: Schematic representation of the site mismatch concept in mixed alkali glasses.Favorable
migration paths are shown by arrows.
38
Our results also do not show any evidence for unlike cation pairing or like-
cation segregation. Rather, they are consistent with statistical mixing of lithium
and sodium cations within mixed alkali borate glasses and thus lend support to
the concept of site mismatch as the fundamental principle underlying the mixed
alkali effect. The basic idea is illustrated in Figure 15: for each of the ions there
are a number of well-matched empty target sites that were previously occupied
by this ion and thus represent favorable size and bonding characteristics for a
successful ion jump. As Na is being replaced by Li, the random mixing of both
ions results in a corresponding decrease in suitable target sites, producing a
marked reduction in cation mobility. Ionic jumps may still occur, but because
of the site mismatch they are strongly correlated with a backward hopping
motion, hence making the jump unsuccessful. As the contribution of a given
ionic species to the total inventory of modifier cations decreases, .this ion
becomes successively immobilized. This prediction well-supported also by
tracer diffusion experiments [36] and by temperature dependent 23Na and 7Li
NMR line shape studies (see Figure 16), which reveal that ion transport (and
hence electrical conductivity) are generally dominated by the majority cation.
Thus, by providing direct experimental evidence for random mixing of the
unlike modifier cations, solid state NMR techniques have provided strong
support for the cation site mismatch being the fundamental physical
phenomenon underlying the mixed alkali effect in glasses.

6. Conclusions

Owing to the technological demands of the modern information society,


fundamental research in the area of solid electrolytes will remain a rewarding
enterprise in the decades to come. Solid state NMR applications for the study
of structure and dynamics of such materials will continue to make important
contributions to this area, resulting in quantitative information of intricate
detail. The applications discussed in this Plenary Lecture serve only as
examples for the extent of information that can be generated even when using
NMR methodology (static variable-temperature NMR, magic-angle spinning,
NMR relaxometry and spin echo double resonance) that is widely
commercially available. However, the technique itself is subject to continued
development, producing ever more powerful multidimensional and multi-
resonance approaches that await applications to interesting materials. The
adaptation of such methodology to solid electrolytes to produce new insights
not available by standard NMR techniques contiriues to be an important part of
the agenda of our research group at the University of Muenster.
39

&v 1Hz

8000

6000

4000

•• TG: 771 K
2000
•• I
..................
O+-~~----r-~-r--~--~~---r---,----.
100 200 300 400 500 600 700 800 900
T/K

t. y I Hz

8000

6000

4000

2000

300 400 500 600 700 800 900


T/K

t. y I Hz
8000

6000

4000
......~' •••

2000 • T G: 725 K
••• I
O+-~~~-,----~---r~~----r-~~--~
••• ... il
100 200 300 400 500 600 700 800 900
T/K

Fig. 16 Temperature dependence of the 7Li static linewidth in (Li 20)03(B 20 3)07 glass, (top) in
[(Li20)0Is(Na20)od(B203)o7 glass (middle) and in [(Li 20)006(Na20 )O24](B 20 3)07 glass
(bottom). Onset temperatures for motional narrowing are indicated. The increase in Tc with
decreasing Li 20 content reflects the successive immobilization of the Li+ ions.
40
7. Acknowledgments

Financial support of this work by the Wissenschaftsministerium Nordrhein-


Westfalen and by the Sonderforschungsbereich 458, funded by the Deutsche
Forschungsgemeinschaft is most gratefully appreciated.

8. References

1. Ernst, R. R., Bodenhausen, G., and Wokaun, A. (1987) Principles of


Nuclear Magnetic Resonance in One and Two Dimensions, Oxford
University Press Oxford U.K.
2. Schmidt-Rohr, K., and Spiess, H. W.(1995), Multidimensional solid State
NMR and Polymers, Academic Press 1994.
3. Brinkmann, D. (1992) Prog. NMR Spectrosc. 24, 527
4. Bjorkstam, J.L. and Villa, M. (1980) Magn. Reson. Rev. 6, 1.
5. Weigel M., Emond, M. H., DeBruin T.H.M., and Blasse, G. (1994),
Chern. Mater. 6, 973.
6. Xia, Y., Machida, N,. Wu, X., Lakeman, C., van Wiillen, L., Lange, F.,
Levi, C., and Eckert, H. (1997),1. Phys. Chern. 101,9180-9187
7. Kawakami, S., Tsuzuki, A., Sekiya, T., Ishikuro, T., Masuea, M., and
Torii, Y. (1985) Mater. Res. Bull. 20, 1435.
8. Jansen, M. (1991), Angew. Chern. 103, 1574 and references therein
9. Lissel, E., Jansen M., Jansen, E., and Will, G. (1990) Z. Kristallogr. 192,
233.
10. Wiench, D.M., Jansen M. (1982) Z. Anorg. Allg. Chern. 486,57.
11. Putnis, A. , to be published
12. Witschas M, and Eckert H. (1999),1. Phys. Chern. 103, 10764-10775
13. Waugh, J.S and Fedin, E. I. (1963), Sov. Phys. Solid State 4, 1633
14. Witschas, M., Eckert, H., Freiheit, H., Putnis, A., Korus, G., and Jansen,
M . (2001) J. Phys. Chern. A 1056808-6816.
15. Bloembergen, N., Purcell, E. M., Pound, R. V. (1948) Phys. Rev. 73, 679
16. Funke, K., Wilmer, D., Banhatti, R. D., Witschas, M., Lechner, R. E.,
Fitter, M. Jansen, M., and Korus, G. (1998) Mater. Res. Soc. Symp. Proc.
525, 469.
17. Torbriigge, T., Witschas, M., Eckert, H. and Jansen, M. to be published
18. Masuda, Y., Sano, M., and Yamatera, H. (1985) 1. Chern. Soc. Faraday
Trans. 81, 127
19. Witschas, M., Eckert, H., Wilmer, D., Banhatti, R.D., Funke, K., Fitter, 1.,
Lechner, R. E., Korus, G., and Jansen, M. (2000) Z. Phys. Chern. 214,
643-673
41

20. Bunde, A., Funke, K., and Ingram M. D. (1998) Solid State Ionics 105, 1-
13 and references therein
21. Tomozawa, M, and Yoshiyagawa, M. (1983), Glastechn. Ber. 56, 939
22. Ingram, M.D. (1987) Phys. Chern. Glasses 28, 215.
23. Ingram, M.D. (1980) J. Am. Ceram. Soc. 63,248
24. Emerson, J. F. and Bray, P. J. (1994), J. Noncryst. Solids 169,87
25. Bunde, A., Ingram, M. D., and Maass, P. (1994), 172-174, 1222.
26. Vessal, B., Greaves, G. N., Marten, P. T., Chadwick, A. V., Mole, R., and
Houde-Walter, S. (1992), Nature 356, 504.
27. Balasubramanian, S., and Rao, K. J. (1993), J. Phys. Chern. 97, 8835.
28. Ratai, E., Janssen, M., and Eckert H. (1998) Solid State Ionics 105,25-37
and references therein.
29. Van Vleck, J. H. (1948) Phys. Rev. 74, 1168
30. Makowka, C. D., Slichter, C. P., and Sinfelt, J. H. Phys. Rev. Lett (1982)
49,379
31. Gee, B., and Eckert, H. (1996), J. Phys. Chern. 100,3705-3712
32. Ratai, E., and Eckert, H. to be published
33. Gee, B. and Eckert H., (1995) Solid State Nucl. Magn. Reson. 5, 113-121
34. Van Wtillen, L., Gee, B. Ztichner, L, Bertmer, M., and Eckert, H. (1996)
Ber. Bunsenges. Phys. Chern. 100, 1539
35. Alam, T. M., McLaughlin, J., Click, C. C., Conzone, S., Brow, R. K.,
Boyle, T., and Zwanziger, J. W. (2000) Phys. Rev. B 104, 1464-1472.
36. Day, D. E. (1976) J. Noncryst. Solids 21,343 .
14N MAS NMR SPECTROSCOPY. AN INSTRUMENTAL CHALLENGE AND
INFORMATORY TECHNIQUE

H. J. JAKOBSEN, H. BILDS0E, J. SKIBSTED, AND T. GIA V ANI


Instrument Centre for Solid-State NMR Spectroscopy
Department of Chemistry, University of Aarhus
DK-8000 Aarhus C, Denmark

1. Introduction

Nitrogen constitutes one of the most important elements in chemistry and in the life and
materials sciences. Both naturally occurring nitrogen isotopes are NMR active: 14N (I =
1) and 15N (I = 112) with abundances of 99.63% and 0.37%, respectively. Due to the
unfavorable properties of the 14N isotope it has rarely been targeted for NMR
observation because it is considered a low-frequency nucleus of low sensitivity and
because of the quadrupolar broadening. This is especially true in the solid state where
the combination of low-resonance frequency and the quadrupolar effects makes the
observation of 14N extremely difficult. Thus, 15N is the nucleus preferred in both solid-
and liquid-state NMR studies of nitrogen materials, usually employing '5N-enriched
materials.
Compared to other low-y quadrupolar nuclei (e.g., 170, 35C1, 39K, 25Mg, 40Ca, 67Zn,
95Mo, 137Ba), one reason that makes 14N NMR in solids a much greater challenge is the
fact that it does not possess a central transition. Applications of solid-state 14N magic-
angle spinning (MAS) are quite few and are generally limited to ammonium ions and a
few other special cases [1-3]. However, taking advantage of state-of-the-art
Transmission-Line Tuning (TLT) CP/MAS probes recently developed by others and in
our laboratory for high-field magnets we find that high-quality 14N MAS NMR spectra
can be acquired for samples with 14N quadrupole coupling constants up to at least 1
MHz. A preliminary study which illustrates these findings has most recently been
communicated [4].
This lecture is intended to demonstrate some of many challenges encountered in
solid-state 14N MAS NMR of this low-frequency quadrupolar nucleus, even at 14.1
Tesla, using standard high-Q double resonance MAS probes and how these difficulties
are coped with using todays available NMR instrumentation. Potential applications of
14N MAS NMR to chemistry, life and materials sciences are numerous and a few will be
presented here. These include studies of solid-solid phase transitions, zeolites,
mesoporous materials, and clay mineralogy in relation to oiVgas formation in the North
Sea.
43
J. Fraissard and O. LapiTUl (eds.), Magnetic Resonance in Colloid and Interface Science, 43-55.
© 2002 Kluwer Academic Publishers.
44

2. Experimental

All compounds used in this study are commercially available and were used without
further purification.
14N MAS NMR experiments were performed at 43.34 MHz on a Varian INOVA-600
spectrometer with a 14.1 T widebore magnet using a VarianlChemagnetics broadband
low-y frequency 7.5 mm T3TM CP/MAS probe (Q-200) with transmission-line tuning
and a specified maximum spinning speed of 7 kHz. The spinning speed of the 7.5 mm
o.d. zirconia rotor was stabilized to <0.05 Hz using the VarianlChemagnetics MAS
speed controller. For spinning speeds above 7 kHz (7 - 15 kHz) we used a homebuilt 5
mm CPIMAS probe. The magic angle was adjusted to the highest possible accuracy
(i.e., about <±0.015°) employing the 14N MAS NMR spectrum of (NH4)2HP04 to give
the best possible resolution of the two different 14N resonances for this sample. Spectra
were recorded using single-pulse excitation with 'tp equal to 2.0 and 0.5 Ils (yB/2n = 40
and 46 kHz corresponding to 'tp90 = 6.3 and 5.4 IlS for the 7.5 mm VarianlChemagnetics
and 5.0 mm homebuilt probe, respectively). Spectral widths of either 0.5 or 1.0 MHz
were employed for the acquisition of the 14N MAS spectra while the repetition delay
could vary from 1 to 1200 s depending on the sample. IH decoupling was employed
during acquisition for proton containing samples. 14N chemical shift are referenced to
the narrow 14N MAS resonance (-0.3 ppm) of an external sample of solid NH4Cl.
Experimental 14N MAS NMR spectra have been analyzed by computer simulationsl
iterative fittings on a Sun Microsystems Ultra 5 workstation using the program STARS,
a solid-state NMR software package developed in our laboratory [5-8] and incorporated
into the Varian VNMR software. Because of the low 14N resonance frequency, even at
14.1 T (vo = 43.34 MHz), and the high Q (-200) of the IH_X MAS probe, the
acquisition of 14N MAS NMR spectra are far from ideal. For example under our
experimental conditions the radiofrequency (rf) bandwidth is highly suppressed during
acquisition of the FID (i.e., 50% reduction for a width of vofQ = 43.34/200 =' 215 kHz)
compared to the corresponding bandwidth for detection at much higher frequencies (e.g.
23Na or 27 AI) for the same Q value. This effect has already been incorporated into the
original version of STARS. To achieve optimizations almost to perfection under these
conditions the original version of the ST ARSIVNMR program has been modified to
include: (i) variation in the rf offset for detectionlexcitation to compensate for non ideal
cable length in the duplexer of the preamplifier, which will otherwise result in a "tilt" of
the manifold of spinning sidebands (ssbs); (ii) a phase shift, dubbed "Q-phasing",
introduced by the narrow rf bandwidth and thus partly caused by the high Q of the
probe; (iii) the second-order cross-term between the quadrupole coupling and chemical
shift anisotropy (CSA) in the average Hamiltonian, a term proportional to voOoCdvo and
usually neglected [7], but of importance for the appearance of the lineshapes for the
individual ssbs in 14N MAS NMR (e.g., for the nitrate ion) [4]. The quadrupole coupling
constant (CQ), its associated asymmetry parameter (l1Q)' the CSA (0 0 ), and its
asymmetry parameter (l1Q) as determined using STARS are related to their tensor
elements by the following equations
45

(1)

where Oiso = 1I3(Ou+Oyy+OzJ The principal tensor elements of the CSA (8) and electric
field gradient (V) tensors are defined using the convention

I Au - 113 Tr(A) I ~ I Au - 1I3Tr(A) I ~ I Ayy - 1I3Tr(A) I (2)

where Au = Oii, Vii. The Euler angles ('II, X, ~), describing the orientation of the CSA
tensor relative to the quadrupole coupling tensor, correspond to positive rotations about
0" ('II), the new Oyy (X), and the final 0" (~) axis. These angles are defined in the ranges
0::; 'II ::; nand 0 ::; X, ~ ::; nl2 [8].

3. 14N Quadrupole Coupling and Chemical Shift Anisotropy

14N MAS NMR spectra for most of the ammonium compounds (and also other nitrogen
samples), that we have obtained, display a manifold of spinning sidebands (ssbs) which
can be covered within a spectral width of 0.5 MHz. With the carrier frequency in the
center of the spectrum and the correct setup of the experimental conditions, the
sidebands are usually placed with symmetrical intensities around the carrier, i.e., with
respect to the isotropic resonance peak. These spectra can be analyzed in a
straightforward manner as described above using STARS considering only the
quadrupole coupling parameters (CQ and II Q) and Oiso. However, in a few cases the
intensities of the ssbs appear with an asymmetric distribution with respect to the
isotropic peak. In such cases the 14N CSA must be taken into account in addition to the
quadrupole coupling in the spectrum analysis using STARS. Moreover, it is
recommended that for such samples the spectra are recorded and analyzed at different
spinning speeds (Dr) (in particular at slow spinning speeds and at the highest possible
magnetic field strength) in order to improve on the precission of the CSA determination.
As an example Figure 1 shows the 14N MAS NMR spectrum of hexagonal boron nitride
(BN) for Dr = 4000 Hz at 14.1 T. This spectrum, the optimized simulation from the
iterative fitting, and the corresponding parameters should be compared with the results
obtained for this material at Dr = 12 kHz and 9.4 T by Jescheke and Jansen [1] . The
determination of an improved value for the CSA (0 0 = 127 ppm versus 160 ppm [1]) is
clearly illustrated by the more well-defined intensities of the ssbs in the manifold for the
present spectrum.
Considering the general tetrahedral symmetry of ammonium ions we were more
surprised to observe an apparent asymmetry within the manifold of ssbs for the 14N
46

(c)

(b)

(a)

i i i i I i
3000 2000 1000 o -1000 -2000 (ppm)

Figure 1. 14N MAS NMR spectra of hexagonal boron nitride at 43.34 MHz. (a) Experimental spectrum with
u, = 4000 Hz and a repetition delay of 20 min. (b) Simulated spectrum for the optimized parameters: ~ = 145
kHz, TlQ =0.09, 00 = 127 ppm, Tlo =0.07, 0;$0 =61.1 ppm, IV =0°, X =6°, and ~ =0°. (c) Simulated spectrum
using the above quadrupole coupling data, but without the CSA parameters. Note the symmetry in this
spectrum as opposed to the asymmetry in (a) and (b).

MAS NMR spectrum at Dr = 5000 Hz of tetrapropylammonium bromide (TPAB) as


illustrated in Figure 2c. This asymmetry becomes even more pronounced in the 14N
MAS spectrum obtained at the lower spinning speed of Dr = 3000 Hz (Figure 2a).
Ascribing the asymmetry to the effect of a possible 14N CSA is confirmed by
performing an iterative fitting of both experimental spectra involving parameters for the
quadrupole coupling as well as the chemical shift interaction as variables. An identical
set of optimized parameters resulted from both spectra and the corresponding simulated
spectra (Figure 2b and 2d) show an excellent agreement with the asymmetries in the
experimental spectra. The determined CSA (0 0 = -23 ppm) for TP AB most likely results
from a distortion of the TP A ion from ideal tetrahedral symmetry caused by packing
with the large bromine ion as has also been deduced from an early X-ray crystal
structure determination [9].
47

(a) (c)

(b) (d)

I iii I iii I I I ' I iii I iii I ' I i I I iii


I1
I iii I I I iii I ii' I
L
iii I i I

1000 400 -200 (ppm) 1000 400 -200 (ppm)


Figure 2. 14N MAS NMR spectra of tetrapropylamrnonium bromide (TPAB) at 43.34 MHz. (a) and (c)
Experimental spectra with u, = 3000 Hz and u, = 5000 Hz, respectively. (b) and (d) Illustrate the
correspondingly simulated spectra employing the same set of optimized parameters: ~ 47.7 kHz, 'lQ = =
0.00,00 = -23 ppm, '10 = O.Q2, 0;", = 28.8 ppm, '" = 0°, X = 20±15°, and ~ = 0°.

The two examples of 14N MAS NMR spectra in Figure 1 and 2 illustrate the effect of
CSA on the ssbs in MAS NMR of the (1 H 0) and (0 H -1) 14N transitions. This effect
is similar to that observed for the inner satellite transition in MAS NMR of half-integer
quadrupolar nuclei [7].
48
4. The Challenge: 14N MAS NMR of the Nitrate Ion

The determination of parameters for the solid-state 14N quadrupole coupling and CSA
interaction in nitrate ions employing 14N MAS NMR represents a major challenge to the
NMR spectroscopists. The reasons contributing to the experimental difficulties have
been considered in our early MAS NMR studies of half-integer quadrupoles [5,6] and
have been enlightened in the recent 14N MAS NMR studies by Fung et at. [2,3]. For the

(c)

(b)

(a)

iii I I
8000 -8000 (ppm)

Figure 3. "N MAS NMR spectra of the nitrale ion in Pb(NO)2 at 43 .34 MH z. (a) Experimental specttum for
U, =6000 Hz. (b) Simulated specttum for the optimi zed parameters: ~ =539 kHz. 1'\Q =0.00, 00= 159 ppm.
1'\0 =0.20. 0... =335 ppm, IjI = X = ~ = 0·, and Q =200. (c) Simulated spectrum using the same parameters as
in (b) except that the probe Q = O. i.e .. infinitely low and corresponding to the ideal specttum.
49
past few years we have worked quite intensively on contributions which could improve
the 14N MAS NMR methodology. The quite successful preliminary results have recently
been presented in a short communication [4]. Generally speaking it is our experience
that a successful outcome of 14N MAS experiments is highly related to the quality of

(a)
Iii I I Iii Iii r I i I I Iii i I I i i i

500 450 400 350 300 250 (ppm)

(b)

(c)

Figure 4. I'N MAS NMR of Pb(N0 3h at 43.34 MHz. (a) Expansions of the centerpart (around 0'10 = 335 ppm)
of the experimental spectrum in Figure 3a. (b) Further expansions illustrating the details of the experimental
lineshapes for the three different resonances in (a). (c) Simulations of the lineshapes in (b) when including the
second-order cross-teno in the average Hamiltonian and using the parameters shown in the captions of Figure
3. (d) The same simulations but excluding the second-order cross-teno.
50

Figure 5. Expansions selected from 14N MAS NMR spectra of the nitrate ion for Pb(N03)2 at 43.34 MHz and
for a spinning speed of 12000 Hz. (a) Experimental spectra with expansions for two ssbs and the isotropic
resonance at 0;", =335 ppm (note the second-order quadrupolar shift). (b) Simulated spectra for the optimized
parameters: CQ =539 kHz, 11Q =0.01 , 00 = 160 ppm, 110 =0.1 - 0.4, 0;", =335 ppm, IjI =X =~ =0°, and Q =
135. (c) The same simulations as for (b) but excluding the second-order cross-term. (d) The identical
simulations as for (b) but for 00 =0 ppm.
51

the MAS probe with respect to observation of low-y nuclei. Thus, it is noted that the
TLT probe technology greatly reduces probe ringing which usually hampers the
observation of low-y nuclei employing high-Q MAS probes with ceramic capacitors as
tuning elements.
As stated above in the experimental section, the high Q of H-X MAS probes is the
main reason for the severe distortions from ideal spectral observation in 14N MAS
NMR. For example the high suppression of the rf bandwidth during acquisition is
clearly illustrated in Figure 3 for the 14N MAS NMR spectra of the N0 3• ion in
Pb(N0 3h- For the experimental spectrum in Figure 3c the dramatic decrease in intensity
for the ssbs with increasing order from the centerband can easily be simulated (Figure
3b) using STARS and taking the ratio volQ into account The effect of the high Q for the
probe is best appreciated by a comparison of the simulated spectra in Figure 3b
(Q = 200) and 3c (Q =0).
Several other effects on the experimental 14N MAS spectra of Pb(N0 3h related to
the probe Q, such as "Q-phasing" and rf offset as discussed in the experimental section,
are considered in all experiments and simulations. Furthermore, the second-order
lineshapes/splittings observed for the individual ssbs throughout the manifold of ssbs
are extremely sensitive to the deviation in angle from the exact magic angle of 54.734°.
Just as importantly, the lineshapes observed for all ssbs cannot be reproduced in
simulations without taking the second-order cross-term between the quadrupole
coupling and CSA in the average Hamiltonian into account This is illustrated by the
lineshapes observed for example for the ssbs around the central transition and shown as
expansions in Figure 4. An excellent agreement is observed between the experimental
spectra in Figure 4b and the simulated spectra including the second-order cross-term
displayed in Figure 4c.
Because of the second-order nature for the observed influence of the CSA
interaction on the lineshapes for the ssbs, the effect of this second-order CSAI
quadrupole coupling cross-term is independent of the applied spinning speed. Thus, the
effect of the CSA is apparent in the lineshapes of the ssbs independent of the spinning
speed and should allow determination of the CSA for any spinning speed.
To confirm these findings experimentally the 14N MAS NMR spectrum of Pb(N0 3)2
has been recorded at a spinning speed v, = 12000 Hz, i.e., at twice the spinning speed
used in Figure 3, using our homebuilt 5 mm CPIMAS probe. Expansions for two of the
ssbs and for the isotropic resonance are shown in Figure 5 along with the corresponding
optimized simulated spectra with and without the second-order cross-term. These
spectra allow determination of exactly the same CSA value (0" = 160 ppm) as obtained
from the slower spinning speed spectra in Figure 4. This value is in good agreement
with the value 0" = 149 ppm we determined from 15N MAS NMR for a 15N-enriched
(95%) sample of Pb(N0 3h.

5. An Application: Phase Transition in NH4N03

The great improvements presented here to studies based on 14N MAS NMR
spectroscopy obviously give potential for numerous applications of this NMR nucleus
in many areas of chemistry. These include a) ammonium and alkylammonium ions, b)
52

ceramics, c) binding studies of nitrogen bases to acid sites of active surfaces, d) ion-
exchanged zeolites, e) templates in microporous and mesoporous materials, f) "pillared"
inorganic layered compounds, g) dynamic studies, and h) phase transitions.
As an example of one of these many potential applications we here present the
application of 14N MAS NMR to a study of the phase transition in NH4N03 at about
34°C. The many phase transitions for NH4N03 have earlier been studied by 15N MAS
NMR employing the change in isotropic 15N chemical shifts as benchmarks for the
phase transition [10]. The potential of using 14N quadrupole coupling parameters as

(a)

(b)

(c)

(d)
i i • I'" i," iii' i' i' i i' iii I. iii ii, Iii. iii iii iii i i i ' i' iii i i

4000 2000 0 -2000 -4000 (ppm)

Figure 6. 14N MAS NMR spectra of the ammonium ion for NH4NO l at. two different spinning speeds. (a)
=
Experimental spectrum for v, 5000 Hz, a line broadening of \0 Hz, and 2000 scans. (c) Experimental
=
spectrum for v, 7000 Hz, a line broadening of \0 Hz, and 789 scans. (b) and (d) illustrate the optimized
simulated spectra corresponding to the experimental spectra in (a) and (c), respectively, and to the parameters
for (a) CQ = 245 kHz, llQ = 0.85, and Ii;.. = -17.6 ppm and for (c) ~ = 127 kHz, llQ = 0.41, and Ii;.. = -15.8
ppm. In addition to the experimental conditions given in the experimental section the deviations from the
exact magic angle were determined to be 0.015° for (a,b) and 0.0\0° for (c,d).
53
greatly improved and alternative benchmarks in studies of phase transitIOns is
highlighted from our study of the NH/ as well as of the N0 3- ion in the course of the
phase transition at about 34°C
The phase transition for NH4 N0 3 at about 34°C is most conveniently approached by
gradually increasing the spinning speed, thereby taking advantage of the frictional
heating [11] as a heat source_ This is illustrated for the NH/ ion in Figure 6 which

(a)

(b)

(c)

Figure 7. "N MAS NMR spectra of the NO,· ion in ISN-enriched ,sNH.NO, illustrating the phase transition at
about 34°C. (a) Experimental spectrum for v, '" 4000 Hz. a line broadening of 10 Hz. and 40000 scans. (c)
=
Experimental spectrum for v, 5000 Hz following a slight decrease in spinning speed and temperature from
34'C (v, '" 7000 Hz) to 26'C (v, =4000 Hz). a line broadening of 10 Hz, and 14659 scans. (b) and (d) show
the optimized simulated spectra corresponding to (a) (Co '" 620 kH z. llQ = 0.25. and 0;", = 335.3 ppm) and (c)
(Co = 662 kHz, llQ =0.00, and 0,., '" 338.9 ppm), respectively, and for deviations from the exact magic angle
of -0.012' (b) and -0.010' (d). A low-intensity manifold of ssbs from the NH: ion (1 % non-enriched) is
observed in the experimental spectrum (c).
54
shows the experimental 14N MAS NMR spectrum for v, = 5000 Hz in Figure 6a.
The simulated spectrum for this MAS condition, which corresponds to a sample
temperature of 29°C (from 207Pb MAS NMR of Pb(N03)2), is displayed in Figure 6b and
corresponds to CQ = 245 kHz, TJQ = 0.85. Increasing the spinning speed to v,= 7000 Hz,
corresponding to a sample temperature of 34°C, results in a 14N MAS spectrum with a
quite different appearance (Figure 6c). The parameters resulting from an iterative fitting
of this spectrum are CQ = 127 kHz and TJQ = 0.41, i.e., these parameters are reduced by
approximately 50% compared to the spectrum for v,= 5000 Hz in Figure 6a. However,
the change in (\soC 4N) (1.7 ppm) is much less pronounced but within' error limits of the
15N isotropic chemical shift difference, Miso( 15 N) (1.9 ppm) reported earlier for the
15NH/ ion for this phase transition [10]. Clearly, the 14N quadrupole coupling
parameters, CQ and TJQ' for the NH4 + ion determined from 14N MAS NMR constitute a
highly sensitive set of NMR parameters compared to the isotropic chemical shift in
characterizing the studied phase transition.
To characterize the same phase transition in NH4N03 by the behavior of the N0 3
ion when subjected to 14N MAS NMR spectroscopy, a sample of 15N-enriched
15NH4N03 was used. This material was selected in order to suppress the intensity of the
14NH/ manifold of ssbs and thereby to better appreciate an undisturbed appearance of
the manifold of ssbs for the N0 3- ion. We should note that it is possible to obtain 14N
MAS NMR spectra for the N0 3- ion without overlap of the NH4 + resonances and of
sufficient quality using ordinary non-enriched NH4N03 and an appropriate spinning
speed. However, for the sake of clarity and for a more clear illustration of the 14N
spectral change which takes place during the phase transition, the 15N-enriched
15NH4N03 sample has been employed in the present study.
14N MAS NMR spectra of the N0 3' ion, acquired using a spectral width of I MHz
and for v, equal to 4000 Hz and 6000 Hz, are presented in Figure 7a and 7c,
respectively. It is noted that the spectrum for v,= 6000 Hz is obtained upon decreasing
v, from the v,= 7000 Hz experiment and clearly indicate the effect of hysteresis since by
further decreasing v, to 5000 Hz gives a spectrum of similar appearance. The two
spectra in Figure 7 exhibit clear differences in their overall appearances and can be used
to characterize the 14N quadrupole coupling parameters for the phase below and above
the phase transition at about 34°C. The corresponding simulated spectra, based on the
parameters resulting from the iterative fitting (see the Figure 7 caption), are shown in
Figure 7b and 7d below the experimental spectra. Indeed the phase change for the NH/
ion can also be observed in the centerpart of Figure 7c as the low-intensity manifold of
ssbs resulting from the 1% 14NH4N03 present in 99% 15N-enriched 15NH4N03. The
c1earcut phase transition observed for the N03- ion leaves great promises to 14N MAS
NMR spectroscopy as a method in studies of phase transitions for nitrates in general.
Finally, we should note that the difference in isotropic 14N chemical shift determined in
this study for the N0 3' ion, ~8isoC4N) = 3.5 ppm, for the phase transition is similar to
that reported by Andersen-Altmann and Grant [10] from the 15N MAS NMR study,
~8isl5N) = 2.4 ppm, taking the error limits into account.
55

6. Conclusion

The techniques of solid-state I'N MAS NMR spectroscopy have been investigated from
an experimental and theoretical point of view. The results achieved have lifted t'N MAS
NMR to a level of "state-of-the-art" techniques. Numerous applications of '~N MAS
NMR within chemistry. life and materials sciences are foreseen.

7. Acknowledgment

The use of the facilities at the Instrument Centre for Solid-State NMR Spectroscopy,
University of Aarhus, sponsored by the Danish Natural Research Council. the Danish
Technical Science Research Council , Tekno logistyrelsen, Carlsbergfondet. and Direkl0f
Ib Henriksens Fond, is acknowledged.

8. References

I . Jcsche ke, G. and Jansen, M. (1998) High -reso luti on "N solid-state NMR spectroscopy. Aligew. Ch~m.
/11/. &/. 37, 1282- 1283.
2. Ermolaev. K. and Fung. B. M. (1 999) High-reso lutinn "N NM R in polycrysl.a.lline solids. J. Chern. Phys.
110. 7977-7982.
3. Khitrin. A. K. and Fung, B. M. (1999) "N nuclear ITl3gnetic resonance of polycr)'stalline solids with fas t
spinning at nr ver)' near the magic angle, J. Ch~rrl Phys, III . 8963-8969.
4. Jakobsen. H. J., Bilds0e, H.. Skibsted, J. and Giavani. T. (200 1) "N MAS NMR spectroswpy: The nilrate
ion, J. Am Cht rrl Soc. 123, 5098-5099.
5. Jakobsen, H. J., Skibsled, J., BildS0e. H. and Nielsen. N. C. (1989) Magic.angle spinning NMR spectra of
satellite transilions for quadropolar nuclei in sol ids. J. Magn . R~son. 85, 173- 180.
6. Sldbsted, J., Nielsen, N. c., Bilds~. H. and Jakobsen, H. J. (1991) Satellite transitiollS in MAS NMR
speclra of quadrupolar nuclei. J. Magn. Ru on. 95, 88-1 17.
7. Skibsted. J., Nielsen, N. C ., B ilds~. H. and Jakobsen, H. J. (1992) >Iv MAS NMR spectroscopy:
Determination of quadrupole and aniSOlropic shie ldi ng lensors, including the relative orientation of their
principal·axis systems. Ch~m. Phys. Leu. 188,405-412.
8. Skibsted, J ., Nielsen, N. C , Bilds~. H. and Jakobsen. H. J. ( 1993) Magniludes and relalive orienUlion of
"V quadrupole coupling and anisotropic shie lding lensors in ~Uvanadales and KV)O, from "V MAS
NMR speCIra. llNa quadrupole coupling parameters for a - and II -NaVOJ' J. Am. Chern. Soc. 115, 735 1-
7362.
9. Zalnn, A. (1957) The crysw-slructure of [elra-n-propyl ammonium bromide, Acta Crrst. 10.557-560.
10. Anderson-AltlTl3nn, K. L. and Granl. D. M. (1993) A solid-Slate UN NMR study of Ille phase transitions
in ammonium nitrate. J. Phys. Chern. <n, 11096-11102.
II. Bjorbolm. T. and Jak obsen H. J. ( 1989) lip MAS NMR of p.SJ. Crystalline-to-plastic phase transition
induced by MAS in a double air-bearing sutor, J. Magn. Reson . 84, 204-211 .
NMR DIFFUSION STUDIES OF MOLECULES IN
NANOPOROUS MATERIALS

1. KARGER, F. STALLMACH
University of Leipzig
Department ofInterface Physics
Linnestraj3e 5
04103 Leipzig. Germany

Recent developments in the instrumentation of pulsed field gradient NMR, in particular


the design of the hardware for the application of ultra-high intensity pulsed magnetic
field gradients in high-field sc magnets and of operation routines for an automated
matching of such pulsed gradients, gave rise to a number of diffusion studies with
porous materials, where new features of interrelation between host structure and guest
mobility were observed. These studies include the detection of internal transport
resistances in MFI-type zeolites, the investigation of the channel architecture in
ordered mesoporous materials of type MCM-41, and the observation of fractal surface
properties of sand grains.

1. Introduction
Since their introduction by Stejskal and Tanner in the mid-sixties [1,2], NMR
diffusion studies by the pulsed field gradient method (PFG NMR) have steadily
deepened our understanding of the internal processes of molecular migration in soft
matter. The potentials of diffusion measurement by NMR have been further reinforced
by implementing the advantages of high-resolution techniques [3] (including the
possibilities of diffusion-ordered spectroscopy [4]) as well as - just on the other side of
the spectrum - by profiting from the large constant gradients provided by the stray
fields of superconducting magnets [5,6]. Among the 8477 citations of NMR papers in
the issues of current contents 2000 (Physical, Chemical and Earth Sciences) as much as
441 papers deal with diffusion phenomena.
NMR studies of molecular diffusion in nanoporous materials are complicated
by a number of particular circumstances:
(i) As a consequence of the interaction with the internal surface, molecular mobilities
are generally reduced ·with respect to the neat liquid. This leads to a reduction of
both the time constant of transverse nuclear magnetic relaxation and of the
diffusivity, which in tum poses higher quality requirements on the measuring
procedure [7].
(ii) Typical samples of nanoporous materials are heterogeneous. They may thus give
rise to the formation of internal magnetic field gradients. Their interference with
the externally applied pulsed field gradients may notably affect the measurements
[8].
(iii) Generally representing loose assemblages of particles, PFG NMR samples often
lack the necessary internal mechanical stability. In order to exclude unwanted
57
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science. 57-70.
© 2002 Kluwer Academic Publishers.
58
displacements of the particles with respect to each other during the measurements,
special precaution is recommended [9].
(iv) Sample heterogeneity often leads to a distribution of observed diffusivities and/or
exchange phenomena, which require a separate consideration [10].

A survey about recent progress in the hardware and operation routine of the
PFG NMR spectrometer [11], which has helped us to successfully meet these
problems, is provided in the subsequent section 2. Section 3 presents the results ofPFG
NMR self-diffusion measurements of alkanes in MFI-type zeolites, where for the first
time a distinct dependence of the intracrystalline diffusivity on the length of the
considered mean diffusion paths has been observed [12]. This experimental finding
may be taken as an indication of the presence of internal transport barriers in MFI-type
zeolites. In the measurements described in section 4, PFG NMR diffusion data are also
used to get structural information, viz. about the architecture of the channel system in
ordered mesoporous materials of MCM-41 [13]. Section 5 presents the results of a
systematic study of molecular diffusion in the free space between beds of sand grains.
The observed dependences of the grain surface on the grain radii reveal that the grain
surface obeys fractal geometry.

2. PFG NMR Instrumentation

Key features of the instrumentation of PFG NMR for application to diffusion studies
with nanoporous materials are the accessibility of large pulsed gradient ·amplitudes
g p , associated with the possibility to generate pulsed field gradients with alternating
direction in order to suppress the disturbing influence of internal magnetic field
gradients. Besides hardware limits determined by the design of the gradient coil and
the gradient current power supply, the most crucial limitation for g p is essentially
brought about by the requirement of equality ("matching") of the time integrals over
the effective field gradients (e.g. g pt5 for rectangular shaped field gradient pulses of
the duration t5) in the preparation and read intervals [14-16].

! I 'I ,

I I
i,6. 6 6\ I',
i : i
g.~-+~~.-.-+---------~--~-r.4-~--
III
: i :

!
!
!
! :':',

g""------1Ji++-;---+-~f-L---;'
'-

Figure I. 13-interval PFG NMR pulse sequence according to COtls et al. (17] with additional field gradients
in the prepare and read intervals (11].
59

Figure 1 shows the pulse scheme, which we have used to generate large
bipolar field gradient pulses [17] and which allows us to detect mismatches of the
pulsed gradient intensities by observing the shift LIt of the spin echo in time domain
[11]. As indicated in the lowest line of the scheme, varying the duration of the
additional gradient g read by a time increment LIt in a subsequent experiment provides
a straightforward possibility of pulsed field gradient matching. In order to do so, one
has simply to detennine the shift LIt of the location of the spin echo due to the
mismatch of the field gradient pulses. In principle, this time shift can be measured by
comparing the positions of the echo maximum for the measurements with and without
pulsed field gradients. However, this method is limited to sufficiently high signal-to-
noise ratios and may yield problems if samples with small initial intensities andlor high
attenuations due to pulsed field gradients are considered. As an alternative possibility,
we have considered, therefore, the convolution of the re~pective transient NMR signals

Q)

A(t')= IMo(go=O,t-t')M(go,t)dt, (1)


-Q)

which attains its maximum at ,'=


LIt [II]. The advantage of this procedure compared
with the direct observation of the spin echo shift for small signal-to-noise ratios is a
consequence of the general property of convolutions that uncorrelated signals (noise)
do not contribute to the convolution function. This approach likewise works with the
Fourier transfonned spin echoes yielding

A(t') = F- 1 {F[Mo(go = O,t -t')]* F[M(go,t)]} , (2)

As gradient coils, actively shielded anti-Helmholtz coils allowing 7.5 mm


sample diameters were used [18]. They have an ohmic resistance of 1.40, an
inductance of 220 JlH and a current-to-gradient conversion ratio of 0.35 (T/m)/A. The
current was generated by means of two power supplies of the type ·TECHRON 8606,
operating in the so-called push-pull configuration, in which the gradient coil is wired in
the middle of two power supplies [II]. With the maximum available output current of
±IOO A it yields a maximum gradient amplitude of ± 35 Tim. The gradient rise times
of 120 JlS pennitts PFG NMR self-diffusion measurements with samples whose
transverse nuclear magnetic relaxation times are well below I ms.
While such small transverse relaxation times are not untypical for molecules
in nanoporous systems, the generally much larger T2 of macromolecular systems
allow the application of much broader field gradient pulses and thus the measurement
of the smallest diffusivities. As an example of the efficiency of the PFG NMR
spectrometer FEGRIS 400 NT [II] used in our studies, Figure 2 shows the temperature
dependence of the self-diffusion coefficient of a PEE-PDMS diblock copolymer (molar
weight ~ 10 kg/mol). The measurements have been carried out with the 13-interval
PFG NMR pulse sequence shown in Figure I with a maximum gradient amplitude of
g p = 35 Tim, with gradient pulse durations and separations of 0 = I ms and
LI = 500 ms, respectively, and with a separation of r = 2.5 ms between the pairs of
TC12 and TC rf pulses. The self-diffusion coefficient measured at the lowest temperature
of 285.5 K is (3.0 ± 0.1) x 10. 15 m2/s. This ·corresponds to a mean square displacement
of (z2)0.5 = 55 nm.
60

1--
~
VI

3.1 3.2 3.3 3.4 3.5


!¥W
Figure 2. Self-diffusion coefficient of PEE-PDMS as a function of the inverse temperature [II].

3. n-Alkanes in MFI-Type Zeolites

Consisting of a system of two sets of parallel channels, perpendicular to each other,


zeolites of type MFI (viz. ZSM-5 and silicalite-l [19]) possess a number of unique
features of intracrystalline molecular transport. The crystallographic x-direction of
these zeolites is traversed by a set of "sinusoidal" channels, while a set of "straight"
channels is aligned along y-direction. Since there is no set of channels in z-direction,
molecular diffusion in z-direction has to proceed by alternating periods of molecular
migration along elements of the straight and sinusoidal channels. In this way, the
principal element of the diffusion tensor in z-direction cannot be independent of the
diffusivities in x- and y-direction. The proposition [20,21] and experimental validation
[22,23] of the relevant correlation rules continues to be a challenging task of zeolite
research.
In view of the two sets of channels, Derouane and Gabelica suggested the
possibility of reactivity enhancement by "molecular traffic control" in catalytic
reactions, if the reactant and product molecules prefer different channel types on their
diffusion paths into and out of the zeolite crystallites [24]. Recent MD simulations [25]
have demonstrated that the occupation probability of the two channels may in fact be
substantially different for the two components of binary mixtures adsorbed in MFI.
Combining the features of "single-file" diffusion [26,27] with the prerequisites of
molecular traffic control, in ref. [28,29] the benefit of molecular traffic control for the
enhancement of catalytic reactivities in channel networks could be elaborated
quantitatively by dynamic Monte Carlo simulations.
Thus being in the focus of numerous diffusion studies, the results on
intracrystalline diffusion in zeolite MFI most vividly demonstrate the enormous
divergence in our present knowledge of zeolitic diffusion in these systems. As an
example, Figure 3 presents the data of intracrystalline diffusion of the n-alkanes in
zeolite ZSM-5 at 300 K for small sorbate concentrations. Though the different
techniques unanimously yield the expected tendency of decreasing mobility with
increasing chain length, there is an unacceptably large difference in the obtained
absolute values.
61

Figure 3. Averaged diffusion coefficients


obtained at 300 K by different techniques for
small concentrations of n-alkanes in the MFI
structure: (0) MD simulation [30], (e)
hierarchical simulation [31], (+) QENS [32], (6)
PFG NMR [33], ('V) single-crystal permeation
[34]. (A) ZLC [35] (after [32]).
o 2 • 6 8 '0 12 " 16 18
Number of carbon aloms

So far, the analysis of diffusion studies has been based on the assumption that the
individual zeolite particles may be considered as homogeneous crystallites. Recent
PFO NMR diffusion studies, however, shed some doubt upon the general validity of
this assumption [12]. Figure 4 shows the results of PFO NMR self-diffusion
measurements of n-butane in two different samples of silicalite, plotted as a function of
the mean diffusion paths. It turns out that for sufficiently low temperatures the obtained
diffusivities notably increase with decreasing lengths of the mean diffusion path. Since
for both samples the mean crystallite sizes were of the order of 20 x 30 x 120 11m3, this
dependence cannot be explained by confinement effects due to the external crystallite
surface. As a most plausible explanation, one therefore has to assume the existence of
internal transP9rt resistances with mutual spacing of the order of or less than I /lm. It is
remarkable that this effect is not observable anymore in the measurements at higher
temperatures. This may be attributed to the fact that now the thermal energy is so high
that the internal barriers do not represent an essential transport resistance in
comparison with diffusion through the genuine zeolite pore structure. Moreover, for
mean diffusion paths sufficiently exceeding the spacing between adjacent transport
resistances, further increase of the diffusion paths will not affect anymore the observed
diffusivities. Unfortunately, the relatively high molecular mobility at the higher
temperatures excludes the possibility of diffusion measurement with smaller
displacements. The PFO NMR diffusion measurements leading to the data presented in
Figure 4 have been carried out by means of the 13-interval pulse sequence as shown in
Figure 1. For comparison, as well included are the results of measurement by the
conventional stimulated echo sequence. They are in complete agreement with the
results of the 13-interval pulse sequence. Thus, for the system under study and the
given experimental conditions, the influence of internal field gradients (which is
annihilated by the action of the 13-interval pulse sequence) results to be negligibly
small.
Alternatively, the influence of internal field gradients on PFO NMR diffusion
measurements by the stimulated echo sequence may also be checked by considering
the dependence of the thus obtained effective diffusivities on the spacing TI between
the first two 7( 12 pulses. Under the assumption that the PFO NMR signal intensity
exponentially depends on the square of the intensity of the applied field gradient
pulses, the influence of the internal field gradients may be shown to lead to a linear
dependence of the effective diffusivity o~ TI2 with the true diffusivity resulting from
extrapolation to the effective diffusivity for TI2 = 0 [36].
62
lE·9

.~~ ..
c c o-a~~"
Figure 4. Effective diffusivity of n-butane in two
samples of silicalite (open and closed symbols)
as a function of the mean molecular
• displacement during the diffusion time at a

.
,
·E
;; lE·l0
o·~;ttv
Ocr
..
siiealite I n-butane
sample •
sample •
383K
297 K
temperature of 213 K, 297 K and 383 K. Results
were obtained with the 13-lnterval PFG NMR
pulse sequence of Figure I. Measurements at
00
'0 •
0
-pie •
....pIe.
297 K
2l3.K 297 K by the conventional stimulated echo
0\ ° v
0 _pleb
sample b
383K
297 K
sequence are as well included (*) (12) .

" ° ....pleb 2131(

4. Structure-Related Diffusion in MCM-41

The propagation rate of guest molecules in porous materials depends strongly on the
pore structure of the host system. Measurements of intraparticle diffusion may
therefore provide information about structural features, which is scarcely to attain by
conventional techniques of structural analysis such as scattering and diffraction
methods. This is particularly true for such structural effects like pore blockage and
leakage in the pore walls. As an example, investigations of this type have recently
contributed to an improvement of our understanding of the real structure of
mesoporous materials of the MCM·41 type [37]. MCM-41-type materials are known to
consist of an array of parallel channels in hexagonal arrangement. In contrast to
zeolites of one-dimensional channel structure (19], which owing to their relatively
narrow channel diameters represent excellent host-systems for single-file diffusion
[26,27], the channel diameters in MCM·41 are of the order of3 up to 10 nm. However,
so far very little has been known about the extension of the channel segments and the
permeability between different channels. As a consequence of the irregular sh~e of
the adsorbent particles, which is quite common for mesoporous materials, analysis of
transient adsorption/desorption measurements is not free from some ambiguity [38].
Being sensitive to the distribution of molecular shifts in the interior of the adsorbent
particles, PFG NMR offers much better possibilities for the measurement of
intraparticle mass transfer.
Information about the intraparticle diffusivities parallel and perpendicular to
the mean channel direction have to be based on an analysis of the shape of the PFG
NMR echo attenuation curves [14,15,39]. This type of analysis may only properly be
performed if all disturbing influences are negligibly small. Besides the influence of the
internal field gradients this concerns in particular the fact that during the observation
time the vast majority of the diffusants has to remain in one and the same crystallite
and, even more stringent, within one region uniform with respect to its transport
properties. In view of the inevitable lower limit of the molecular displacements still
detectable by PFG NMR, this requirement implies the application of highly ordered
mesoporous materials, which are only available with rather modest particle sizes of up
to 5 Ilm longitudinal extension [40]. With such relatively small adsorbent particles it
cannot be excluded that during the observation time of the PFG NMR experiment a
notable part of the diffusants is able to leave the individual particles, giving rise to an
additional constituent in the PFG NMR attenuation curve [41,42], which dramatically
impairs the evidence of an analysis of its shape. We have circumvented this
complication by monitoring the water diffusivity in MCM-41 samples overloaded with
63

water at temperatures below O°C [13]. In this way it is possible to find experimental
conditions under which the phase of the intraparticle water is still mobile, while the ice
fonned in the space between the particles definitely excludes any exchange of water
molecules between different adsorbent particles. Thus, the PFG NMR signal
attenuation curve is in fact found to be in excellent agreement with the theoretical fit
based on the assumption that intraparticle diffusion is described by an axisymmetrical
diffusion tensor. Figure 5 shows the main elements of the diffusion tensor as
detennined for an observation time of 10 ms in comparison with thediffusivity of free
water. Figure 6 gives a survey about the complete dependence of the resulting
diffusivities on the observation time,

10,9

10,9

,
10,10
0
10,10 CoO
.
, ii",- '7 O[:J
en til

'"E , ......
N
E
[:J

"
..... 10. 11
Q 10,11 Q

e (SO
e e 0
10, 12
•...•......•........ 10,12 ~ ~~

3.6x1O,03 4.0x1O,03 4.4xlO,03 0.001 0.01 0.1

6 / s
Iff / Kl

Figure 5. Arrhenius plot of the components of Figure 6. Dependence of the parallel (0, ., +)
the diffusion tensor of water in MCM-41, and perpendicular (0, e, X) components of the
measured parallel (II) and perpendicular (e) to axisymmetrical self-diffusion tensor on the
the axis of symmetry with an observation time of observation time for water in MCM-41 at 263 K
10 ms. For comparison, the corresponding (squares and circles, two samples) and 269 K
diffusivities of neat water [43] are as well (crosses, only one sample) [13].
included [13].

The origin of the anisotropy of self-diffusion must be sought in the


morphology of the MCM-41 material. There are two possible reasons for the observed
behaviour, viz,(i) the "microscopic" anisotropy of the individual particle itself as
observed by SEM or light microscopy and (ii) the inherent "nanoscopic" anisotropy of
the axisymmetrical pore structure inside the domains of the hexagonally arranged
channels fonning the individual particle. Using Einstein's relation

(3)
64
the observed time dependence of the effective self-diffusivities may be transferred into
the corresponding time dependence of mean square displacements. Figure 7 shows
these data for the parallel and perpendicular components of the self-diffusion. Since
both mean square displacements do not increase linearly with the observation time, the
observed time dependence reflects some kind of restricted self-diffusion.

10 1 r--------------,
Figure 7. Dependence of the parallel (_,0) and
perpendiCUlar (.,0) components of the mean
square displacement on the observation time for
water in two MCM-4l samples at 263 K. The
horizontal lines indicate the limiting values for
the axial (full lines) and radial (dotted lines)
components of the mean square displacements
for restricted diffusion in cylindrical rods, which
....·__..........·1.3 were calculated via eqs.(4) and (5). The lengths I

./
o and diameters d of the rods are given in units of
...... ~...... 1.0 micrometers. The oblique lines (45°), which are
plotted for short observation times only,
/ ·...._-.. _· ..·0.7 represented the calculated time dependence of
.,.-6.' Q the mean square displacements for unrestricted
q,/ • (free) diffusion with Dpo< = 1.0 X 10. 10 m1s l
...~ (full line) and with Dpe'l' = 2.0 X lO.Jl m1s· 1
(dotted line), respectively [I3].
0.01 0.1

t:, I s

The origin of the restriction becomes clear if one compares the mean square
displacements observed with the limiting values expected for restricted self-diffusion
inside a cylinder of length I and diameter d at infinite observation times. These
limiting values may be calculated to be [14, 15,39]

/2
(Z2(fj...." 00») =6 , (4)
and
d2
(z2(fj"""00»)=S' (5)

respectively.
The horizontal lines in Figure 7 indicate these limiting values for I = 3 - 5 11m
and d = 0.70 - 1.3 11m. Obviously, the mean square displacements measured for an
observation time of 60 ms are consistent with restricted diffusion in a cylinder of about
4 11m length and about 1.0 - 1.3 11m diameter. Since these values agree satisfactorily
with the length and the diameter of the rod-like MCM-41 particles as determined by
SEM, it must be the "shape" anisotropy of the particles, which determines the diffusion
anisotropy at long observation time. With decreasing observation times, however, the
observed diffusion anisotropy must be expected to be an immediate consequence of the
internal particle structure, where the direction of the larger diffusivity may
unambiguously attributed to the longitudinal extension of the MCM-41 particles. This
65

conclusion is in agreement with the fact that owing to the axial symmetry of the system
the channel axes must in fact be expected to be preferentially aligned in the direction of
the longitudinal extension of the particles.
The observed similarities in the temperature dependences (activation energy)
of the water self-diffusion in MCM-41 for both the parallel and perpendicular
components and the bulk (Figure 5) suggest that the self-diffusion in the bulk and
inside the nanopores follows the same thennodynamic processes. Obviously, the
nanopores of 3 nm diameter do not represent an additional energetic barrier.
Consequently, the reduced diffusivity in the direction of the channel axis (D par) must
be caused by steric obstructions, which the water molecules experience on their
diffusion path. Though being much smaller than the diffusivity in channel direction,
there is also a finite diffusivity perpendicular to it. There are essentially three models,
by which this finding might be explained:

(i) The channel walls are penetrable for the water molecules.
(ii) The channels consist of only relatively short segments, allowing matter
exchange between different channels at the end of each segment.
(iii) Over the considered diffusion paths, the channel axes notably deviate from
straight lines, so that displacements along the mean of the channel axes are
always associated with displacements in the direction perpendicular to them.

On the basis of the presently available experimental data, a distinction


between these different cases is impossible.

5. Molecular Diffusion in Beds of Sand Grains


In contrast to zeolitic adsorbate-adsorbent systems (cf. the representations in
Figure 4), for PFG NMR studies with natural porous rock samples internal field
gradients are well known to be generally of substantial influence. This is illustrated in
Figure 8 by comparing the dependence of the respective effective diffusivities on the
diffusion time as derived from PFG NMR measurements with the conventional
stimulated-echo sequence and with the 13-interval pulse sequence. In this case, the
influence of the internal field gradients still present in the measurements with the
stimulated echo, leads to a time dependence, which is substantially different from the
behaviour observed with the 13-interval sequence. The measurements have been
carried out with a hexadecane-saturated unconsolidated sediment (sand) sample
consisting of grain particles with grain radii between 50 and 200 flm.

1.0

0.1
~~ .. .. • ... 8··
· ....
.• .
.0.' \ •
e
Q '--. e ..
Figure 8. Effective diffusivities of n-
hexadecane in a bed of sand particles
0.4 .... . determined in PF6 NMR measurements
"
...........
0.2
by the stimulated spin echo (e) and by
the 13-interval PFG NMR pulse
0.0
sequence (_) (46) at room temperature
O.OE+OO 5.0E.oI I.OE.o5 1.5E.o5 2.0E-<l5
(293 K).
ID.-" J" I m
66

Measurements of this type are usualiy analysed on the basis of the relation
[44,45]

D{.1) 4 ~ ~
- - =1- r Svv Do.1 + R{v Do.1) (6)
Do 9v 7r
with
(7)

denoting the ratio between the total volume Vp of the liquid phase (coinciding with the
volume of the space between the sand grains, i. e. the pore volume) and the total
surface S of pore system (which coincides with that of the sand grains). The value
R{ ~ Do.1) accounts for higher-order tenns in ~ Do.1, which becomes increasingly
important with increasing observation time. Do denotes the free diffusivity in the pore
space, i.e. the effective diffusivity in the limiting case of zero observation time.
Applying eq. (6) to analyzing the time dependence of the effective diffusivities as
obtained without suppressing the influence of the internal field gradients by use of the
13-interval pulse sequence would obviously lead to completely erroneous data.
As a consistency check of the results for the surface-to-volume ratio obtained
by the 13-interval pulse sequence, for a large variety of systems we have saturated the
pore space with both water and hexadecane. According to eq. (6), plotting the relative
values of the obtained effective diffusivities versus ~ Do.1 should yield a master
representation, independent of the considered diffusant. The experimental data shown
in Figure 9 are in excellent agreement with the expected behaviour. Irrespective of the
fact, that the diffusivities of the considered probe molecules in the free state differ by
about one order of magnitude, for any of the considered systems there is a remarkable
agreement in the initial slopes of the representations. Both probe molecules coincide,
therefore, in their evidence about the ratio between the surface and the volume of the
porous host system. For comparison, also the data obtained for the neat liquids are
presented, which clearly yield horizontal straight lines indicating that the surface-to-
volume ratio approaches zero.

1.0

0.9

0.8

0.7

0.6
0

~
'::J 1.0 Figure 9. Relative apparent self-diffusion
Q coefficients (D(d)/Do) as function of (Dod)o.l for
0.9
water (open symbols) and hexadecane (full
0.8 symbols) in different model systems (Figure 9a:
bulk liquids (*, *), glass spheres (.,0) and
0.7
o 0
Buntsandstein (A, 6)) and in two un-
0.6 consolidated sediments (Figure 9b: sand A (II,
0) and sand C (. ,0)). The lines given show the
O.S
0 averaged initial decays of D(d)/Do [46,47].
67
Being sensItive to the ratio between the total volume occupied by the
introduced liquid and the area of the internal surfaces, which represent the boundary
between the fluid and solid phases, PFG NMR diffusion studies provide a unique
possibility for analysing the dependence of the grain surfaces on the grain size. We
have particularly focused on the investigation of the scaling behaviour of the specific
surface area

(8)

where Sm and mg stand for the grain surface area and mass, respectively. Assuming
Euclidian geometry, due to Sex: dg2 and mg ex: d g3 , the quantity Sm must be expected
to scale inversely proportional with d g , i.e. Sm ex: d;1 .
The unconsolidated sediment investigated with this approach originates from a
glacial sand deposit in Central Germany located about 10 km northwest of Leipzig.
The sand was dried on air and separated into grain size fractions by sieving. From each
grain size fraction a PFG NMR sample was prepared by filling it into 7.5 mm o. d.
glass tubes. Distilled water was added stepwise to the sands until 100 % saturation of
the pore space was achieved. In order to ensure a high packing density and a complete
saturation of the pore space, the samples were centrifuged after each step of saturation.
For all samples, the surface-to-volume ratio of the pore space (Sv ) was determined on
the basis of eq. (6) from representations of D( t1) / Do as a function of the relative
displacement in free space, JDot1, i.e. from the slope of the initial decay of the
normalized effective diffusivity in a representation of the type of Figure 9. From the
experimentally determined surface-to-volume ratio Sv, the specific surface area Sm
may be easily calculated by multiplication with the total volume of the introduced
liquid and division by the total mass of sand. Figure 10 presents the thus determined
values as a function of the averaged grain diameter.

10- 1

Ne--
00

.......................•.....
--cJ Figure 10. Specific surface area Sm as a function
of the averaged grain diameter d, for the grain
size fractions of the sand obtained from PFO
NMR diffusion measurements. The star (*)
10-2
shows the corresponding values measured with
the original sand. The full and dotted lines
represent the regression line of the log Sm-vs.-
log d, fit (slope: 0.80 ± 0.05) and its confidence
interval, respectively. The horizontal error bars
100 1000
represent the remaining size distribution in the
individual grain size fractions resulting from the
width of the meshes used for sieving (48).

The data of Figure 10 show that the scaling law notably deviates from the
behaviour expected from Euclidian geometry. The specific grain surface is in fact
found to decrease less pronounced with increasing diameter as to be expected for
68
simple Euclidian bodies. This experimental finding can be rationalized on the basis of
the probably quite reason(ible assumption that, as a consequence of their special
genesis in a terminal moraine during the ice age, smaller grains have been subjected to
a more intense erosion by mutual grinding down ending up in a less pronounced
surface ruggedness. Since for compact matter, grain volume and mass is still expected
to scale with the third power of the grain diameter, according to eq. 8, the grain surface
results to be best described by a fractal dimension of2.2 rather than 2.
The described procedure of dimension analysis is based on the comparison of
different length scales. It should be kept in mind that these length scales are subjected
to certain conditions. Obviously, the scaling law leading to the fractal dimension of the
grain surface has been observed with grain diameters ranging from roughly 100 Ilm up
to 800 Ilm. The length of the diffusion paths considered in the PFG NMR experiments
( ::::: 5 Ilm - 20 Ilm) represents the length scale of the "probe", which is applied to attain
structural information. The fact that the obtained dependences (Figure 9) are in
satisfactory agreement with eq. (6) indicates that at least for the smallest displacements
attained in the measurements, the grain boundary appears as a plain surface. This
means in particular, that the radii of the curvature giving rise to the observed fractal
behaviour must be larger than the smallest displacements. Finally, one must not
confuse the presently chosen way of fractal surface analysis with the procedure by
Avnir et al. [49,50]. In their analysis the mean diameter (and not the mean diffusion
path length) of the molecules served as a scale of their measurements. Structural details
with curvature radii much smaller than the minimum diffusion path lengths remain
unobservable in the surface analysis by PFG NMR.

6. Conclusion

Irrespective of the fact that already half a century ago NMR has been
recognized as a useful technique for diffusion measurement, so far there is no end in
sight in a rather prosperous development of the application of NMR in this particular
field. Progress in the application of a particular technique is often associated with the
development of the experimental basis. In a first section, we have therefore described
the recent progress in our PFG NMR instrumentation, in particular the mode of
automated matching of high-intensity field gradient pulses. In the subsequent sections,
we have illustrated the benefit of these new possibilities by presenting our most recent
results of diffusion studies with liquids in porous materials. As a common feature of
these investigations, all of them have to be considered as incomplete:
Though the time dependence of intracrystalline diffusion for various alkanes
in MFI-type zeolites, as reported in section 3, rather stringently suggests the existence
of internal transport barriers, this finding cannot be considered as anything more than a
faint hint, how the huge differences in the results obtained by different techniques for
this very system might be explained. The problem how to generally explain the
discrepancies between the various techniques continues to exist.
It is true that the diffusion studies presented in section 4 provide completely
new insights into the pore architecture of meso porous material of type MCM-41. It has
to be admitted, however, that we failed with similar studies for other specimens of
MCM-41. So far, our only explanation is that these materials even stronger deviate
from the conception of the ideal MCM-41 structure.
Finally, in section 5 the diffusion studies with saturated beds of sand have
been interpreted by assuming a fine structure on the surface of the individual sand
grains, which becomes more pronounced with increasing grain sizes. Though this
69
conclusion appears to be stringent from the point of view of the NMR measurements,
in future studies the involvement of complementary techniques of surface structure
analysis is inevitable.

Acknowledgement: PFG NMR dIffusion studies of porous materials substantially


benefited by the accessibility of high-quality zeolite material. Over many years, this
material has been kindly provided by Sergey Petrovich Zhdanov, Russian Academy of
Sciences, St. Petersburg, one of the leading scientists in the field of zeolite synthesis.
We dedicate this contribution, held in his city, to the occasion of his 90 th birthday in
April 2002 and thank him for all his help and friendship. Financial support by the
Deutsche Forschungsgemeinschaft (SFB 294) is gratefully acknowledged.

7. References
1 Stejskal, E. O. (1965), 1. Chem. Phys. 43,3597-3603.
2 Stejskal, E. O. and Tanner, 1. E. (1965), J. Phys. Chem. 42,288-292.
3 Stilbs, P. (1987), Progr. in NMR Spectroscopy 19,1-45.
4 Barjat, H., Morris, G.A., Smart, S., Swanson, A.G. and Williams, S.C.R. (1995), J. Magn. Reson. B
108, 170-172.
5 Kimmich, R., Unrath, W., Schnur, G. and Rommel, E. (1991), J. Magn. Reson. 91, 136-140.
6 Chang, I., Hinze, G., Diezemann, G., Fujara, F. and Sillescu, H. (1996), 76, 2523-2526.
7 Pfeifer, H. (1972), NMR-Basic Principles and Progress 7,53-153.
8 Karlicek, R. F. and Lowe, I. J. (1980),J. Magn. Res. 37, 75-91.
9 Blir, N.-K., Klirger, J., Krause, C., Schmitz, W. and Seiffert, G. (1995). J. Magn. Reson. A 113.278-
280.
10 Karger, J. (1985). Adv. Coli. Inter! Sci.23. 129-148.
II Galvosas. P.• Stall mach. F.• Seiffert. G.. Karger, J.• Kaess. U. and Majer. G. (2001). J. Magn.
Reson 151, in press.
12 Vasenkov, S.• B6hlmann. W.• Galvosas. P.• Geier, 0 .• Hui Liu and Karger, J. (2001). J. Phys.
Chem .• in press.
13 Stall mach, F.• Karger, J., Krause. C.• Jeschke. M. and Oberhagemann. U. (2000), J. Amer. Chem.
Soc. 122.9237-9242.
14 Karger, J., Pfeifer. H. and Heink. W. (1988), Adv. Magn. Reson.12.• 2-89.
15 Callaghan. P. T. (1991) Principles of Nue/ear Magnetic Resonance Microscopy, Oxford University
Press. New York.
16 BIOmich, B. (2000) NMR imaging of materials. Clarendon Press. New York.
17 Cotts, R. M., Hoch, M. J. R.• Sun. T. and Markert, J. T. (1989). J. Magn. Reson 83, 252-266.
18 Karger, J.• Blir, N.-K .• Heink. W., Pfeifer, H. and Seiffert. G. (1995). Z. Natur/orsch. 50a, 186-190.
19 Meier. W. M.• Olson, D. H. and Baerlocher. Ch. (1996) Atlas o/Zeolite Structure Types, Elsevier.
London.
20 Karger, J. (1991), J. Phys. Chem. 95, 5558-5560.
21 Karger. J., Demontis, P.• Suffritti. G. B. and Tilocca. A. (1999),J. Chem. Phys. no, 1163-1172.
22 Caro. J.• Noack, M.• Marlow, F., Peterson, D., Griepenstrog, M. and Kornatowski. J. J. (1993), J.
Phys. Chem. 97. 13685-13690.
23 Hong, U., Karger, J., Pfeifer. H.• Muller, U. and Unger, K. K. (1991), Z. Phys. Chemie 173.225-
234.
24 Derouane, E. G. and Gabelica. Z. (1980). J. Catal. 65.486-489.
25 Clark, L. A., Ye, G. T. and Snurr, R. Q. (2000), Phys. Rev. Lell. 84,2893-2896.
26 Kukla. V., Kornatowski, J., Demuth, D., Girnus, 1.. Pfeifer, H., Rees. L. V. C.• Schunk. S.• Unger.
K. K. and Karger. J. (1996), Science 272. 702-704.
27 Hahn, K., Karger, J. and Kukla, V. (1996), Phys. Rev. Lell.76, 2762-2765.
28 Neugebauer. N., Brauer, P. and Karger, J. (2000),J. Catal. 194. 1-3.
29 Brauer, P., Karger, J., Neugebauer, N. (2001), Europhys. Lells. 53. 8-14.
30 Runnebaum, R. C. and Maginn. E. J. (1997),J. Phys. Chem. B. 101.6394-6408.
31 Maginn. E. J., Bell, A. T. and Theodorou. D. N. (1996);1. Phys. Chem. 100,7155-7173.
32 Jobic, H. (2000), in N. K. Kanellopoulos (Edt.), Recent Advances in Gas Separation by
Microporous Ceramic Membranes, Elsevier. Amsterdam. 109-138.
33 Heink, W., Karger, J.• Pfeifer. H., Datema and Nowak, A. K. (1992), J. c. S. Faraday Trans. 88.
3505-3509.
70
34 Talu, 0., Sun, M. S. and Shah, D. B. (1998), AIChE Journal 44, 681-694.
35 Eic, M. and Ruthven, D. M. (1989), Studies in Surf. Sci. Catal. 498, 897-913.
36 Vasenkov, S., Galvosas, P., Geier, 0 ., Nestle, N., Stallmach, F. and Klirger, J. (2001), J. Magn.
Reson. 149,228-233.
37 Kresge, C.T., Leonowicz, M.E., Roth, W.J., Vartuli, J.C. and Beck, J.S. (1992), Nature 359, 710-
712.
38 Campos, D.S., Eic, M. and Occelli, M. (2000), Studies in Surf. Sci. Catal. 129,639-648.
39 Stallmach, F. and Klirger, J. (1999), Adsorption 5,117-133.
40 Overhagemann, U., Jeschke, M. and Papp, H. (1999), Micropor. Mesopor. Maler. 33,165-172.
41 Hansen, E. -W., Courivaud, F., Karlsson, A., Kolboe, S. and Stocker, M. (1998), Micropor.
Mesopor. Maler. 22, 309-320.
42 Courivaud, F., Hansen, E. W., Karlsson, A., Kolboe, S. and StOcker, M. (2000), Micropor.
Mesopor. Mater. 35136, 327-339.
43 Weinglirtner, H. (1982), Z. Phys. Chem., Neue Folge 132, 129-149.
44 Mitra, P. P., Sen, P. N. and Schwartz, L. M. (1993), Phys. Rev. B 47, 8565-8574.
45 Latour, L. L., Mitra, P. P., Kleinberg, R. L. and Sotak, C. H. (1993), J. Magn. Reson. A 101,342-
346
46 Vogt, C. (2001), Diploma Thesis, Leipzig University.
47 Vogt, C., Galvosas, P., Klitzsch, N. and Stallmach, F. J. Appl. Geophysics, submitted.
48 Stallmach, F., Vogt, C., Karger, 1., Helbig, K. and Jacobs, F. Phys. Rev. Lells., submitted.
49 Avnir, D., Farin, D. and Pfeifer, P. (1985), J. Colloid Interface Sci. 103, 112-123.
50 Pfeifer, P. and Avnir D., (1983), J Chem. Phys. 79,3558-3565.
NMR STUDIES OF THE MESOMORPHISM, STRUCTURE AND DYNAMICS OF
SOME NEW PYRAMIDIC LIQUID CRYSTALS

Z. LUZ', R. POUPKO', E. J. WACHTEL', V. BADER2 and H. ZIMMERMANN 2


IThe Weizmann Institute ofScience. Rehovot 76100. Israel. 2Max-Planck-Institute for
Medizinische Forschung. Jahnstrasse 29.69120 Heidelberg. Germany

1. Introduction

The molecules of discotic liquid crystals consist of disc-like cores to which flexible side
chains are linked at the periphery (I]. In most cases the cores consist of flat medium size
aromatic moieties, such as benzene, naphthalene and triphenylene, but more recently
discogens with much larger cores, such as hexabenzocoronene have also been prepared (2,3].
The side chains are usually normal aliphatic groups, which are linked to the core by ether or
ester bridges, but there are many other modifications including heteroatoms or even direct
carbon-carbon bonds. They serve as "lubricants" for the planar reorientation of the molecules.
Discotic mesogens usually form columnar mesophases, although nematic and lamellar
discotic phases are also known . The columnar phases may be regarded as one-dimensional
liquids, being ordered in the transverse plane of the lattice and liquid-like along the column
direction. Usually the center cores of the discogens are rigid and flat, however flexible, non-
planar cores are also known. These are in most cases macrocyclic moieties, such as ortho- or
meta- cyclophanes [4,5]. In the present work we discuss columnar mesophases derived from a
lower homologue of the orthocyclophane, i. e. tribenzocyclononatriene (TBCN)
R R

R
R

Figure I. The molecular structure (left). crown conformer (center) and saddle conformer ofNTBCN

The TBCN core can acquire two conformations, rigid crown and flexible saddle, as shown in
Fig. I. Unsubstituted TBCN or when only the peripheral sites (R) are substituted (R'=H),
prefer the crown conformation. On the other hand substitution ofTBCN in the ortho aromatic
sites (R'iH) with sufficiently large groups destabilizes the crown form and equilibrium of
both conformations is obtained [6].
71
1 Fraissard and O. Lapina (eds.). Magnetic Resonance in Colloid and inleiface Science, 71-81.
© 2002 Kluwer Academic Publishers.
72

In earlier publications we and others have shown that hexasubstituted TBCN (in the R
positions) with sufficiently long ether or ester chains form columnar liquid crystals of various
two-dimensional symmetries [7-9). Because of the crown conformation of these derivatives
the resulting mesophases were termed "pyramidic". In the present work we extend the range
of the TBCN liquid crystals to nonasubstituted esters, with R, R'=-OqO)C n.)H2n. ). As
indicated above the crowding at the ortho- positions results in an equ ilibrium mixture of the
crown and saddle forms . Alkanoyloxy derivatives of both conformers are mesogenic and
exhibit columnar measophases. In this paper we use I3C MAS NMR to characterize the
structure and dynamics properties of these mesophases.

2. Hexasubstituted TBCN

Before discussing the new nonasubstituted TBCN liquid crystals, we briefly review the
corresponding hexasubstituted homologues, with R being ester or ether side chains and R'=H.
When these chains are sufficiently long the compounds are mesogenic and exhibit columnar
mesophases. They are often polymorphic with a sequence of several phases, including
hexagonal and rectangular. The cone-shaped of the TBCN core leads to special physical
properties for these mesophases.

0$-G)0 00-00
0000 00-00
0000 0000
0000 0000

)v~v
0 C
o 0
)000C
0000

Figure 2. Left: Stacking of the TBCN molecules into columnar structures. Right: Ferroelectric and
antiferroelectric tetragonal and hexagonal two-dimensional arrays of polar columns.

Because of their cone shape these molecules, in the columnar phases, tend to pile on top of
each other in an ordered way as shown in Fig. 2, leading to a macroscopic electric moment
along the column. One can then imagine several arrangements of the columns; ferroelectric
(all columns in a domain are parallel to each other), antiferroelectric (the columns are divided
into two oppositely polarized sub lattices), and paraelectric (where the columns are oriented
randomly up and down). It is interesting to note that, due to "frustration" effects (see Fig. 2)
hexagonal mesophases cannot arrange in an anti ferroelectric lattice. Indeed, in an electric
polarization study of a TBCN-based pyramidic hexagonal phase, it was found to be
ferroelectric [10).
Another quite unusual phenomenon, related to the phase optical anisotropy, 6n, was found in
TBCN hexasubstituted with oqO)C6H40CIOH2) [7). As a function of the temperature the
sign of 6n changes reversibly from positive to negative, apparently reflecting changes in the
73

conformational equilibria of the side chains. In these compounds Lln is determined primarily
by the orientation of the various aromatic moieties in the molecule. In the indicated system
these include the benzene rings of the rigid TBCN core and of the mobile phenylene rings in
the side chains. The average orientations of the latter change with temperature and gradually
cancel and then overcompensate the. Lln of the rigid TBCN core, leading to reversal of its sign.
The dynamic and ordering properties of the molecules in these pyramidic mesophases have
been extensively studied by deuterium NMR of specifically labeled mesogens [11,12). The
molecular reorientation rate in these phases fall in the range of deuterium dynamic NMR, thus
providing very detailed information on the motion of the molecules in these phases. A very
special dynamic feature was observed when comparing the deuterium NMR lineshapes of the
ring methylene in hexasubstituted alkyloxy and alkanoyloxy derivatives. A detailed analysis
showed that the molecular reorientation in the former homologues proceeds via a diffusion
mechanism, while in the latter it involves three-fold jumps. Examples of spectra
demonstrating this effect for the hexaheptyloxy TBCN homologue are shown in Fig. 3.

Figure 3. Deuterium NMR spectra of


hexaheptyloxy TBCN deuterated in the ring
methylenes as a function of the temperature
within the mesophase region (center) and
comparison with simulated spectra calculated
for the three-fold jump (left) and planar
diffusion (right) models.

It is intriguing that the small difference between the nature of the bridging groups, -0- and -
OC(O) -, results in what appears as a major mechanistic difference. In fact it is possible to
show [13] that the different behavior of the two homologous series can be described by a
common three-fold rotational diffusion model in a potential of the form V=-Vocos(3q». The
ether and ester series then correspond to the two ends of this model with low and high Vo
values, respectively.

3. Nonasubstituted TBCN (NTBCN)

On substituting the TBCN core with nine substituents (R=R':;tH), two fundamental new points
need to be considered. The first concerns the crown-saddle equilibrium. The crowding in the
ortho- positions of the benzene ring destabilizes the crown form. On the other hand the open
structure of the saddle (see Fig. I) and its high flexibility result in an increase of the entropy
and corresponding lowering of the free energy to the extent that it becomes the dominant
species at high temperatures. The two conformers are rapidly interconverting in solution and
74

in the isotropic liquid state at high temperatures, but the interconversion is slow in the solid
and meso phase states, as well as in solutions at room temperature and below. Consequently
the crown and saddle forms can be separated by chromatography and their mesomorphic
properties can be studied independent of each other. We describe the results of this study in
the next sections.
The second point that needs to be considered in the nona substituted TBCN derivatives
concerns chirality. The compounds are now asymmetric and their synthesis leads to a racemic
mixture of the L and R optical isomers for both the crown and saddle conformers. The saddle
conformer pseudorotates and thus racemizes rapidly. However the crown form is long lived at
room temperature and in principle it should be possible to separate it into its twp optical
isomers. Attempts to do so using chiral column chromatography failed so far, but their NMR
signal could readily be discriminated in a chiralliquid crystalline solvent [14]. In Figure 4 are
shown deuterium NMR spectra of crown nonamethoxy-TBCN-d6 (deuterated in the ring
methylenes and with R=R'=OCH3) dissolved in the lyotropic liquid crystalline mixture poly
benzylglutamatelDMF (PBGIDMF). The right spectrum corresponds . to th~ racemic
PB(L+D)G/DMF mixture, while the left one to the chiral solvent PBLG/DMF. The doubling
of the spectrum in the latter solvent may clearly be seen.

Figure 4. Deuterium NMR


spectra nonamethoxy TBCN
deuterated in the ring NAD signal of
methylenes dissolved in the DMF
lyotropic liquid crystalline
chiral PBLG/DMF (left) and
...-
racemic PB(L+D)GtbMF
i i i I i i
(right). 20.0 10.0 0.0 -10.0 20.0 10.0 0.0 -10.0
ppm ppm

3.1 PREPARATION, MESOMORPHISM AND HIGH-RESOLUTION NMR OF NTBCN

We have synthesized nona-alkanoyloxy TBCN (NTBCN-Cn) with side chains containing


Cn=2 to 13 carbons in each side chain. All members with Cn>4 (>3) are mesomgenic in the
crown (saddle) form and exhibit hexagonal columnar mesophases. We have studied these
mesophases by Optical Microscopy, Differential Scanning Calorimetry (DSC), X-ray
diffraction, high-resolution IH and I3C NMR, and Carbon-13 MAS NMR. In the following we
first review the chemical and physical properties of the compounds in the neat form and in
solution, and then discuss in some detail the dynamic properties of the meso phases as
manifested in their I3C MAS spectra.

The nona-alkanoyloxy derivatives were obtained by reacting the corresponding acid chloride
either directly with nonahydroxy TBCN (heating to 70°C and recrystallizing from boiling
ethanol) or in pyridine solution (at O°C and recrystallizing from CHCb/ethanolat room
temperature). The purified products obtained at high temperature always exhibited two spots
in column chromatography, which could be identified as the crown and saddle conformers,
while those prepared and purified at room temperature showed only a single spot
75

corresponding to the crown form. As the starting materials, nonamethoxy and nonahydroxy
TBCN exist only in the crown form, the above results show that preparation under "high
temperature" conditions (70°C) results in crown-saddle isomerization, while at room
temperature the process is slow on the time scale of the preparation experiment (a few
hours/days). It was therefore possible to quantitatively separate the two isomers by column
chromatography at room temperature. The various homologues of the NTBCN-Cn series
exhibit similar properties. We therefore mainly discuss one of the homologues (C8) as a
typical example.

, ~-"---r--Ti--~--ri--~i- r--~--ri--~-'i---r--~-'~

8 6 4 2 ppm 8 6 4 2 ppm
*
*

~i~~i~~i~i.~ ~i~~i~i~~i~
160 120 80 40 ppm 160 120 80 40 ppm

Figure 5. lH (top) and llC (bottom) high-resolution NMR spectra of the crown (left) and saddle (right) forms of
NTBCN-C8 in CD2 CI (*),. .

In Fig. 5 are shown high-resolution 'H and DC spectra of the two isomers of C8. The spectra
of the crown form are consistent with a rigid structure of C3 symmetry (note the AX quartet of
the ring methylene in the 'H spectrum, and the six distinct benzene peaks in the DC
spectrum). On the other hand the saddle form is flexible, undergoing fast pseudorotation
(single ring methylene peak and six, rather than eighteen benzene lines in the 'H and DC
spectra, respectively). As the temperature of the solution is raised to above 50°C, the isomers
start to equilibrate at measurable rates, first on the time scale of hours and at higher
temperatures (>1 OO°C) on the time scale of minutes and seconds. We used the 'H NMR
spectra in this range to determine the equilibrium and kinetic parameters for the crown/saddle
system in solutions of C8 . The results, including the relevant thermodynamic parameters, are
summarized in Fig. 6. Note, in particular, the effect of entropy in stabilizing the saddle form
at higher temperature, caused by the flexibility of the cyclononene ring and the open structure
of this form. On the other hand steric repulsion of the methylene hydrogens prefers the crown
form which is dominating at lower temperatures.
76

c,.,.,,~ saddl.
I<,
.~ .
c: .75
0
.~
:;

..
0.
&. .50
c
0
l~
"
A.
ol: .25 H~ - 32.3 KcaVmol
..:
~ -13.7 e.u.

a 2.8 2.9 3.0


H: a 28.S KcaVmol

·100 -50 a 50 100 150 t/OC (IO'IT>'I" S: -3.le.u.


All -3.8 KcaVmol

Figure 6. Equilibrium fractional populations as function of the temperature (left), Arrhenius plots for
interconversion (center) and thermodynamic data for the crown-saddle NTBCN-C8 system in CD2Ch solution.

3.2 OPTICAL MICROSCOPY AND DSC

In Fig. 7 are shown DSC thermograms of the C8 homologues. The top two traces are for the
crown form with clear melting (45°C) and clearing transitions (J64°C). The saddle form
exhibits a more interesting effect. On first heating (bottom trace) there is a broad melting
transition (40°C to 50°C), an endothermic peak (at J37°C) followed immediately by an
exothermic peak and then by a clearing transition at J64°C, identical to that of the crown
form . If the sample is now rapidly cooled from the isotropic liquid into the mesophase and the
heating experiment repeated, a very similar thermogram is obtained on reheating (third trace).
On the other hand heating a sample that was cooled slowly gives a thermogram (second trace)
identical to that of the

sad

:~,::::~ ~,.--
U'C phase C8

__
M_

after slow coolin1

145 Mcrn ~_=~


13TC
after fast cooling
--"-"-"--_ _-"--l

J
C8 saddle M

M"dL-
K
lirst heating
65

! I ,I ! I 1 I I 1 , I ! ,
- -, - - - - • I ,

o 50 100 150 80 7.0 60 5.0 4.0 ppm


ti'C

Figure 7. (left) DSC thermograms of the crown and saddle forms ofNTBCN-C8 and (right) 'H NMR spectra
(low field only) of samples taken during a heating run of the saddle form
77

crown form. These results demonstrate a peculiar mesomorphic driven isomerization. We


note that the clearing temperature of the crown (164°C) is higher than that of the saddle
(l37°C) by some thirty degrees. Thus when heating the saddle isomer at 137°C an isotropic
liquid is obtained, and an equilibrium is immediately established with both saddle and crown
isomers. At this temperature range the latter prefers the meso phase state and it "precipitates"
out of the system producing the exothermic peak of the DSC signal. As the crown form
crystallizes the equilibrium shifts until all the saddle transforms to the crown. On further
heating, at the clearing point of the crown form (l64°C), a second endothermic peak is
obtained, yielding an isotropic liquid with the equilibrium saddlelcrown mixture. Rapid
cooling will retain this equilibrium, but on slow cooling the crown form will "precipitate" out
from the mixture, while the saddle form is depleted in the mixture. This cycle of event was
confirmed by monitoring the IH spectra of samples taken on raising and lowering of the
temperature. Examples of such spectra are shown on the right hand side of Fig. 7.
The above interpretation of the results was also confirmed by optical microscopy and by x-
ray measurements. Optical microscopy of the measophases of both isomers, in particular of
the crown exhibited characteristic features of columnar phases. The X-ray measurements
showed that the measophases are hexagonal (ordered) with lattice parameters (in A) of 25.4
(5.0) and 24.0 (4.8) for the crown and saddle forms of C8, respectively. The figures in
brackets correspond to the intracolumnar stacking distances. They were found to increase
slightly on heating within the mesophase.

3.3 CARBON-13 MAS NMR OF THE SOLID AND MESOPHASES

In this last section we discuss structural and dynamic properties of the columnar phases of the
crown and saddle conformers as reflected in their I3C MAS spectra. Examples of such spectra
(low field only) in the solid and mesophase regions of each conformer are shown in Fig. 8.
The spectra of the crown form in the solid and mesophase are quite similar and, except for the
better resolution in the mesophase, the peak positions correlate well with the corresponding
spectra in solution. The mesophase spectrum shown in Fig. 8 corresponds to 5rC, which
corresponds to the low temperature range of the mesophase. When the sample is further
heated within the mesophase region, gradual line broadening, followed by line narrowing,
takes place, as shown by the sample spectra in Fig. 9. The behavior clearly indicates the
setting in of a dynamic process that modulates the chemical shift interaction of the carbons.
We interpret the effect as due to rotational diffusion of the molecules within the columnar
structure. This interpretation is supported by a rotor synchronized 2D exchange experiment
shown on the left hand side of Fig. 10. It clearly shows cross peaks linking spinning side
bands belonging to the same carbon manifolds. They form series of "lines" parallel to the
main diagonal, indicating that the motion is pure rotational. The rate constants for this process
were derived by comparing the experimental spectra with simulated lineshapes calculated by
the Floquet method [15,16]. The results are plotted on the right hand side of Fig. 9. From the
Arrhenius curve in this figure the activation parameters, t-.Ht=34.0kcallmol and t-.St=51.4e.u,
were derived. Both values are unexpectedly high for a molecular reorientation in a
mesophase. A possible explanation for these unusual values is to assume that the activation
barrier for the reorientation is not constant but rather decreases with increasing tempt<rature.
This will increase the slope of the Arrhenius curve and result in a larger intercept at the T-tO
78

t/ °c t / °c

57 M

20

in
25 CDCI) 25

~
" .. ,
~-,

I' "'" , 1 I" 1 '


170 160 150 ppm 170 160 150 140 130 ppm

Figure 8, IlC spectra in CD2CI 2 solutions (bottom) and MAS spectra in the solids (middle) and mesophases
(top) of the crown (left) and saddle (right) forms ofNTBCN·C8,

trc
135

143,7 111,6 84,1 60,3 tfC


10'

10'
0

o~ o
o.

10'
~

en
32
10'

10'
H' '" 34.0 Kcal/mol
S' '" 51.4 e.u.

2.4 2,6 2,8 3.0


(10'fT)lK"
ppm

Figure 9. (left) I3 C MAS spectra (Iow.field only) of the crown form ofNTBCN·C8 as a function of the
temperature. (right) Arrhenium plot and activation parameters derived from the spectra on the left,
79

20

60

100

.' .' J

+-.i+
120
t"
140 · {(
160
.' ~
180

200·

2(10
1
150
1
100
1
50
-_..
ppm 200 150
1
100
1
50 PP'"

Figure 10. 20 MAS exchange spectra of the crown (left) and saddle (right) forms ofNTBCN-CS.

end. Such an explanation is reasonable since thermal expansion will increase the stacking
distance between the molecules in the columns, thus reducing the barrier to reorientation. As
indicated above, X-ray measurements indeed indicate a small, but definite increase in the
stacking distance within the mesophase region.
The situation is quite different for the mesophase of the saddle. The low-field I3C MAS
spectra in the solid and mesophase (see Fig. 8) are now quite different from those of the
crown conformer and of the saddle conformer in solution. In particular in the mesophase
spectrum one can count 18 peaks as expected for a static saddle conformation, where all
carbons are inequivalent. The signals due to the carboxyl carbons at 170 ppm are also
consistent with 9 partly overlapping peaks. This clearly indicates that the saddle conformation
is static and that the pseudorotation is hindered in the solid and columnar phases. Moreover,
no line broadening is observed on heating the mesophase and the spectrum shown in Fig. 8
for the saddle form at 107°C remains essentially unchanged up to the clearing temperature of
the mesophase (137°C). No motion is detected even in the ultraslow regime as may be judged
from the complete lack of cross peaks in the 2D exchange spectrum shown on the right hand
side of Fig. 10.
These results can be rationalized in terms of the schematic drawings in Fig. II. The center
part shows the arrangement of the columns in a hexagonal arrangement, while the drawing on
the left and right sides of the figure depict the stacking of the molecules in the mesophases of
the crown and saddle conformers. One can easily visualize the slipping of the crown
molecules within their columns and it is likewise easy to see why the saddle molecules remain
stuck. In principle one could imagine a combination of pseudorotation and reorientation
processes that would allow overall rotation of saddle molecules within the columns, similar to
what has been observed in derivatives of orthocyclophane [17]. Apparently such a process
does not take place in the mesophases of the saddle conformers ofNTBCN.
A
80

'O',
.~ , , ,

.'~
R~·~l
J.
l

l OR

It·

.~
:A~~

:~. '~Q:~R
~
l~l~l
R ~ It R It

Figure II . Schematic representation of the hexagonal columnar mesophase (center) and of the stacking of the
crown (seft) and saddle (right) molecules within the columns.

4. Summary

Pyramidic columnar mesophases lend themselves to a variety of dynamic studies by solid-


state NMR methods. In the present work we emphasized the use of I3C MAS and rotor
synchronized 2D exchange experiments for measuring molecular reorientation within the
columns in columnar structures. Several other applications of I3 C MAS NMR to study
discotic liquid crystals can be envisioned. One relates to density modulation within columnar
mesophases. The stacking of discotic molecules into columnar structure may pose packing
problems because of.the incommensurability of aliphatic and aromatic moieties. Under certain
conditions this may lead to density modulation within the columns [18]. Such modulations
may lead to inhomogeneous line broadening of IJC MAS spectra and thus provide a tool for
the investigation of the phenomenon. Related effects on the deuterium NMR of discotic
mesophases were found in columnar mesophases related to derivatives of rufigallol [) 9].
Another application of I3 C MAS NMR to dynamics in discotic mesophases relates to dipolar
recoupling. In most cases the observed carbons are coupled to nearby protons and strong
decoupling of the 13C_1H dipolar interaction is required to observe narrow lines. In the
presence of motion, when the rate of the process is of the order of the decoupling field the
effect of the latter is destroyed. This results in lines broadening, thus providing another tool
for rate measurements [20]. It is likely that the molecular reorientation rate in many discotic
columnar mesophases fall in this time scale.
81

References
I. D. Demus, J. Goodby, G.W. Gray and H.W. Spiess (eds.), Handbook of Liquid Crys/a/s. (1998) Wiley·
VCH, Weinheim, Cannidge, A.N. and Bushby, R.J. Synthesis and Structual Features, V. 2B, Chap. VII
pp. 693·748: Chandrasekhar, S. Columnar, Discotic Nematic an Lamellar Liquid Crystals: Their
Structures and Physical Properties, V. 2B, Chap. VIII pp. 749·780: Boden, N. and Movaghar, B.,
Applicable Properties of Columnar Discotic Liquid Crystals, V. 2B, Chap. IX pp. 78 1·798.
2. Fechtenkotter, A., Saalwachter, K., Harbison, M. A., MUllen, K. and Spiess, H.W. (1999), Angew.
Chern. Int. Ed 38, 3039.
3. Ochsenfeld, c., Brown, S.P., Schnell, I., Gauss, J. and Spiess, H.W. (2001), 1. Am. Chern. Soc. 123,
2597.
4. Spielberg, N., Sarkar, M., Luz, Z., Poupko R. and Zimmermann, H., (1993), Liq. Crys/ais 15,311.
5. Bonsignore, S., Du Vosel, A., Guglielmetti, G., Dalcanale, E. and Ugozzoli, E. (1993), Liq. Crys/a/s.
13,471.
6. Colet, A., (1987), Te/rahedron, 43, 5725.
7. Zimmermann, H., Poupko, R., Luz, Z. and Billard, 1., (1985), Z Na/urforsch. , 40a, 149; (1986) 41a,
1137.
8. Poupko, R., Luz, Z., Spielberg, N. and Zimmermann, H., (1989),1 Am. Chern. Soc .. 111,6094.
9. Malthete,1. and Collet, A., (1985) Nouv. 1. Chim. 9, 151.
10. Jakli, A., Saupe, A., Scherowsky, G. and Chen, Xin Hua, (1997), Liq. Crys/a/s. 22, 309.
I I. Poupko, R., Luz, Z., Spielberg, N. and Zimmermann, H. (1989),1. Am. Chern. Soc. 111,6094.
12. Zamir, S., Luz, Z., Poupko, R., Alexander, S. and Zimmermann, H. (1991),1. Chern. Phys. 94,5927.
13. Zamir, S., Poupko, R., Luz, Z. and Alexander, S. (1991), 1. Chern. Phys. 94, 5939.
14. Lesot, P., (2001), private communication.
15. Schmidt, A. and Vega, S., (1987),1. Chern. Phys. 87,6895.
16. Luz, Z., Poupko, R. and Alexander, S. (1993),1. Chern. Phys. 99, 7544.
17. Kuebler, S.C., Boeffel, C. and Spiess, H.W., (1995), Liq. Crys/a/s. 18,309.
18. de Gennes, P.·G., (1983)1. Phys. Lell. 44, 1·657.
19. Werth, M., Leisen, 1., Boeffel, c., Dong, R.Y. and Spiess, H.W., (1993),1 Phys. /I France. 3,53.
20. Rothwell, W.R. and Waugh, J.S. (1981),1. Chern. Phys. 74, 2721.
INVESTIGATION OF CONFORMATIONAL CHANGES OF ORGANIC
MOLECULES SORBED IN ZEOLITES BY HR MAS NMR SPECTROSCOPY

1. ROLAND, D. MICHEL AND A. PAMPEL


Faculty of Physics and Geosciences, University of Leipzig,
Linnestr. 5, 04317 Leipzig, Germany

1. Abstract

The application of IH NMR spectroscopy to the investigation of organic molecules


sorbed in catalytic systems like zeolites is often limited due to the poor spectral resolution.
During the last years a notable enhancement in resolution of the spectra is achieved when
random local magnetic fields due to susceptibility effects are partially averaged out by the
application of magic angle spinning techniques (MAS). High resolution (HR) MAS NMR
spectroscopy is used to investigate the dynalnics of olefinic hydrocarbons in NaX zeolites
and their conformational changes in the adsorbed state. The high precision arrived in the
measurements allows the determination of chemical shifts and coupling constants of IH
NMR spectra in dependence on loading and temperature, which are used in order to get
deeper insights into the dynamics of the adsorption process. In this paper for the first time
conformational changes of adsorbed molecules could be investigated by means of HR MAS
NMR spectroscopy.

2. Introduction

NMR spectroscopy is a versatile tool for the investigation of structure and dynamics of
molecules in interaction with internal surfaces of porous solids like zeolites. In many
systems the applicability of one and two dimensional NMR techniques is limited due to the
poor resolution of the IH NMR spectra because of the presence of inhomogeneous local
magnetic fields in the polycrystalline materials [1-4]. As described in a previous paper [3],
the line shape of the spectra of molecules adsorbed in zeolites is strongly influenced by the
distribution of the magnetic fields resulting from the polarization of the grains in the zeolite
powder by the strong external magnetic fields. The line shape may be modulated by the
thermal motion, too. The static magnetic field distribution can be reduced in the NMR
spectra by using magic angle spinning (MAS) techniques with moderate spinning
frequencies of 3 to 6 kHz [4]. The combination of standard high resolution NMR
spectroscopy with MAS will be referred to as HR MAS NMR. Complex Gaussian and
Lorentzian line shapes with spinning site bands arise at low spinning rates (several 100 Hz).
Already at moderate spinning frequencies well resolved spectra with Lorentzian line shapes
appear the residual line widths of which show a quadratic dependence on the flux density B
of the external magnetic field. This dependence is in competition with the linear increase of
83
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 83-95.
© 2002 Kluwer Academic Publishers.
84

the chemical shift resolution with increasing magnetic fields as demonstrated recently [4]
for measurements with B values from 2.35 to 17.6 T. In the present systems, the line
narrowing by means of MAS, however, cannot be treated independently of the influence of
the thermal motion of the molecules over distances where the local field inhomogeneities
may be randomly modulated.
The present work is devoted to the investigation of the adsorption of simple olefms
sorbed in zeolites, like I-butene molecules in NaX. Standard ID and lH_ \3C HSQC
measurements will be combined with IH NOESY measurements. A good resolution in the
IH MAS NMR spectra for these systems has been found, both with respect to chemical
shifts and indirect spin-spin couplings, when a proton Larmor frequency of 600 MHz was
used. Based on these improvements in spectral resolution, it is the aim of the paper to study
conformational changes, the dynamics of adsorption, and the interaction of the molecules
with the Na+ cations in the zeolite framework. Since it is already known from previous
NMR studies [5] that changes of the l3C chemical shifts at certain molecular groups of the
adsorbed molecule can be related with small changes of the electronic density and, thus,
with specific interactions of the 1t electrons of the molecule with the Na cations, in
particular the question arises whether the interaction of the adsorbed molecules with
adsorption centers will also lead to structural changes.
It is the aim of the paper to combine chemical shifts measurements for varying
temperature and pore filling factor in order to investigate not only the change in the electron
density but also to check whether interatomic distances in an adsorbed olefinic molecule are
changed in the course of adsorption. First a detailed analysis of the line shape for the
different molecular groups is carried out which allows to determine chemical shift values
for the adsorbed state with high accuracy. In the subsequent part IH NOESY NMR
measurements are run in order to derive distances between certain intramolecular proton
spin pairs. On this experimental basis the state of the adsorbed molecules will be
investigated.

3. Experimental

All NMR measurements were run at a DRX 600 NMR spectrometer (Bruker Analytik
GmbH) using a high resolution MAS probe head with a r. f. gradient coil for 4 mm rotors
and a diameter of the cylindrical sample tubes of 3 mm at a MAS frequency of 5 kHz. The
measuring conditions enabled a uniform excitation of the NMR spectra investigated. Single
pulse IH NMR experiments at room temperature were performed at nearly the same
experimental conditions (32 scans, repetition time 10 s at a proton relaxation time of
Tl = 2 s, and a 1t/2 pulse length between 9 and 11 I1s). The measurements were realized for
six different pore filling degrees from 0.13 up to 4 molecules per supercage in a NaX
zeolite.
Temperature dependent measurements (between ca. 270 K and 340 K) were performed
for the I-butenelNaX system with a loading of 1 molecule per supercage using the same
parameter settings as already mentioned. Measurements at higher temperatures were not
run because at higher temperatures above ca. 420 K appreciable catalytic transformations
occur, like the double bond isomerization from I-butene to 2-butene or the skeletal
isomerization to isobutene [6].
85
HSQC experiments for a comparison between the temperature dependence of the proton
and the carbon chemical shifts, especially of the olefmic =CH- group of I-butene, were run
at different temperatures (295, 311, 323 and 343 K) using the HR MAS NMR probe head
described above.
For the comparative HR NMR experiments on a liquid sample, I-butene was dissolved
in CDCh and a small amount of about 1% TMS was admitted for the calibration of the
spectra. To get an optimum resolution in the MAS NMR spec~a of the adsorbed molecules,
special attention had to be paid to the preparation of the zeolites. Especially, a low
concentration of paramagnetic impurities had to be guaranteed in order to reduce the
influence of paramagnetic species on the proton spin relaxation rates. Zeolite powder with a
relatively large mean crystallite size of at least 5J.1.m was necessary to ensure a high
efficiency of the MAS averaging. These requirements were best fulfilled in the case of
zeolites which were prepared by S. P. Shdanov et al. (St. Petersburg) [7] as used in our
experiments. After heating the samples at 670 K in vacuo, a known amount of molecules
for adsorption is given to the cylindrical sample tubes by distillation. Subsequently the
tubes were sealed off in a way to enable magic angle sample spinning [8].

4. Results and Discussion

4.1. ANALYSIS OF THE LINE SHAPE OF THE DIFFERENT GROUPS OF LINES

As mentioned, this work strongly benefits from the appreciable increase of resolution in
the IH NMR spectra of molecules adsorbed in porous media if MAS techniques are applied.
The better spectral resolution achieved for I-butene molecules sorbed in NaX is
demonstrated in fig 1 in comparison to the spectra for I-butene molecules dissolved in
liquid CDCh. The typical proton resonance lines for I-butene molecules can be resolved in
the adsorbed state but only a few J couplings are visible. The conclusions from these
measurements are mainly derived by comparing the experimental spectra with simulations
which are rather simple and which, thus, are not shown here. The spectra can be explained
by using the chemical shifts in the adsorbed state, the different J-coupling constants and the
residual line widths in the case of MAS measurements. The same J-coupling constants as
found for the molecules in the solution are taken [4]. The chemical equivalence of the two-
CH 2- and the three -CH3 group protons in the adsorbed species can be inferred from the
spectra. Two chemically non-equivalent protons in the =CH 2 group appear as expected
from the SP2 hybridization in the olefmic part of the molecule. The asymmetry in the proton
positions is also reflected in the difference of the constants of the indirect spin-spin
couplings with the proton in the olefmic =CH- group. J-coupling constants greater than
about 6 Hz can be resolved. In particular, the 3JHH coupling constants between the protons
in the =CH 2 and the =CH- group can be measured. Compared to the liquid state, the proton
NMR lines for the groups =CH 2 and =CH- of the molecules in the adsorbed state are
shifted to lower magnetic fields, because of changes of the electronic densities due the
interactions of the adsorbed molecules with the exchangeable Na+ cations as adsorption
sites [5]. In the temperature range from 270 K to 340 K there are practically no changes in
the residual line width of ca. 5 Hz for all pore filling factors. This shows that homogeneous
contributions to the line widths due to spin-spin interactions are not dominant which are
86
expected to change from group to group. This allowed us to use the same inhomogeneous
line shape function for each line in the simulations. A good agreement with the
experimental spectra is achieved for the aliphatic groups -CH2-, -CH3 and for the olefmic
=CH- group.

I
--
i

(e)

I
I
I ~ U L I
t
,
~
1
I ' ,
6
~
.
·1
4
~r~~

3
1 ' 1
0
. -~-~

pprr
(bf (~) ( d)
(a)

~~'~r~·~·'~ U
(f)
~·- '1-'~~· ~·nT.-~r~·'- ~' ~r-"~~·'·
J\ Jl
p~· T- r ·· r· · ~' · ·· r-rr ·· ~' ·· ' - r - ' · - r · "-T - ~ ·'- ' --'- -r --

6 5 4 3 2 1 0 pp'm

(a)
(b) ~ 1

~ H~
~V~
Iv NI

1
jJVJJV\J.\JV\
1 1 1
- JU"'"--- 1 1 1
5.90 5. 85 ppm 5 . 00 4.95 ppm

--~~"----- 1 1 1 1
~-- 1 ; 1
6 .50 6.4 5 6 . 40 ppm 5.10 5 . 05 ppm

(c ) (d)
i
Ai~ .~ ~ /1 A
_.-J\) \
I
I

I _... - - __ _ -'J v v 'IA.._ _... I


"
\ ) "--____ _
I ~-. -- --

-· '~- '-··r ···.···. ·- -'- ·--r- · ·' -r ' ---~·--' --------r- -·' -·'-' --' · -T~-~~-----r-· · . r
- -- - -
------·-r··-- ·· -'-----· ··- · ·-·--· ' --- -r ~- ·' ·········r -·····_·r······_·_·

2.1 2.0 ppm 1. 00 ppm

---~ ~-- _J~______


I
I
A

1
~ ~ ~~I_ - -.----.---r -

IL -_2.25
~~-l
I

2.15
2 . 20 ______
_ ______ ppm 1. 00 ppm
--- - ----- --- -

Figure 1. Comparison of lH NMR spectra of I-butene (T = 298K,


VH = 600.13 MHz, VMAS = 5 kHz) dissolved in CDCh with 1% ofTMS (top)
and sorbed at NaX (vrot = 5 kHz, second spectrum from top). The inserts
show the line shape for different molecular groups [(a) = CH, (b) = CH 2, (c)
= CH 2 and (d) = CH3] in adsorbed and in solution state. Peak (e) is due to
TMS which was used for calibration of I-butene in solution. In the spectrum
of the adsorbed molecules the CH3 group chemical shift was calibrated to 1
ppm. Peak (f) belongs to the remaining protons in D20, which was used for
locking but could not be taken for the calibration of the spectra.
87
There is a distinct change, however, in the line shape for the molecular group =CH 2- when
the pore filling factor is increased from ca. 0.l3 to 4 molecules per supercage (fig 2). The
particular situation observed for the =CH 2 group (b) can be understood in the following
way: Since the two protons in this group are chemically non-equivalent as long as the
molecules are in an isotropic medium like in solution state, in first order of the perturbation
theory each of the proton NMR lines of a =CH 2- group gives rise to a doublet. Here the
splitting is caused by the 3JHH-coupling to the proton in the =CH-group which differs for
both =CHz- protons, viz. 10.23 Hz and 17.32 Hz for the cis- and trans-configuration,
respectively. Obviously, in case of adsorption the difference between the proton chemical
shifts of the both =CH 2- group protons is decreased more and more if the pore filling factor
is reduced (see fig 2). Moreover, this difference decreases appreciably with decreasing

I 'b ) ' 0) , OJ I

I~
[ ......... j......

. :-. - ' .
6 5 60
I

55
;~ ~~- .l
1I-.. .
........:;::i'~;:;:::;:::;:::;:::;:~:~:;::::r .. .......,:::;... j.....

5. 0 4.5
~..
4. 0 3.5
i~-
3.0 2.5
1
.. ...... j........ ........-.....~
2. 0
...-::;::..,
}.5
jI
-......-.....-......................... ..

-_. . . . . _ . . . _) ppm

_~f\J I~""~;v\~~ rL.~ ·~-P .p ~\J


6.5 6.4

(\> 1101.!
pp~ 5.1 pp.
...- / 2'T.:..........
.. 1. 05 1

,1
0 PP~
!
...-1 U L- - ~~~""r""""- -~ .. I
"~~JA
I ~:J~·1 ·,T~~.,.i
................... -~..,... •... - r·· .. ···~.............·...
M;i11
! 1
. - ......... - .......

6.5 ~'. ~ 5 .1 1:
1.
' 2.~ pp . 1.05 ~l.O
O PP· ,
1

l--r~--r~-1
~-"',! f-...--J \..- -J i
/~M~il pp LA~
- ; !---,.......- ........--.I
6.5 6 .4 PPrTt

~
Figure 2. I-butene sorbed at NaX (T = 298K, VH = 600.l3 MHz,
VMAS = 5 kHz) for different loadings. The upper part shows the complete IH_
HR-MAS-NMR spectrum with a loading of 4 molecules per supercage. In the
inserts the dependence of line shape and line position for each molecular
group «a) = CH, (b) = CH 2, (c) = CH 2 and (d) = CH3) is illustrated for
various loadings, viz. 4, 2, 1, 0.5, 0.2 and 0.1 molecules per supercage. The
line splitting due to the coupling constants is partially resolved.
88

temperatures (fig 3) and we arrive at a (partial) superposition of the one part of the trans-
doublet with the complete cis-doublet. The changes in the spectra may be shown by a
simulation where we have assumed the same coupling constants for adsorbed I-butene
molecules as in solution (viz. 10 and 17 Hz). A rather good agreement was achieved for the
different loadings used, clearly showing the peculiarities in the behavior of the protons in
the =CHr group (fig 2 (b)). This means especially that for the adsorbed molecules the
measurable difference in the chemical shifts between the both chemically non-equivalent
=CH 2 protons is considerable less than the value (48 Hz) for the I-butene molecules
embedded in CDCh solution. For the adsorbed species this difference reduces from ca. 21
Hz at 340 K about linearly to 10Hz at 290 K and it is expected that the difference is further
reduced for decreasing temperature and fmally seems to become zero.

3870
• =CH-group
3860 • -C~-group

3850
~

:st::::::::=:
~3840
ro

i 1305.0
1302.5
1300.0+-~--.---~--.-~--.-~-.-~-,-
20 30 40 50 60 70
temperatue rcl

3025

20 30 40 50 60 70
teJll)erab..re I.'Cl

Figure 3. Chemical shifts (in Hz) for I-butene sorbed in NaX in dependence
on temperature for the =CH- and the -CH2- group (a) and for the two
protons of the =CH 2 group (b) of the molecule. The arithmetic average [tJ. in
(b)] of the cis- [e in (b)] and trans-proton [_ in (b)] of the =CH 2 group is
independent of temperature in the investigated range.
89
For a better understanding of these effects in the proton spectra we have to discuss also the
change of the chemical shifts with temperature in a range from 290 to 340 K and a pore
filling factor between ca. 0.13 and 4 molecules per supercage. The dependencies are
illustrated in figs. 2and 3. The =CH- group and the aliphatic groups show a shift to higher
fields with increasing temperature and loading: In this context we have also to take into
account that there is a dynamic equilibrium between the molecules interacting with the
adsorption sites and the remaining ones sorbed in the zeolitic cages. In case of a fast
exchange between the molecules at the adsorption sites (fraction PA) and the remaining
ones (fraction PM) an averaged chemical shift, 0 = PAOA + PMOM, is measured for the various
protons which is simply due to the change of the fractions of both types of molecules. In a
small interval there is approximately a linear variation with temperature. It is interesting to
note that the linear variation with temperature is also reflected in the chemical shift of each
proton in the =CH 2- group but the mean value of both shifts shows only a very weak
variation with temperature (see the triangles in fig 3). Moreover, if there is a so-called weak
adsorption complex between the molecule and the active surface site, i.e. if the product of
the equilibrium constant k of the complex formed and the number of adsorption sites NA is
small, kN A < I, then we expect also a small and nearly linear dependence of the measured
chemical shift on the total number of sorbed molecules [9, 10]. This assumption is valid in
our case as may be inferred from the chemical shifts of the \3C NMR line in the HSQC
spectrum for the =CH group for adsorbed I-butene molecules [5] which will not be further
discussed here. In conclusion to this point, the assumption of a fast molecular exchange
between the types A and M of adsorbed species seems to be well established.
The dependence of the chemical shift difference on the pore filling factor (see the
distinct changes in the line shape in fig 2b) as well as the reduction of this difference with
decreasing temperature (fig 3b) lead to the conclusion that molecules in interaction with the
adsorption sites have about the same resonance positions for both =CH 2 protons. The latter
could be related with a model (see below) where either a conformational change of the 1-
butene molecule associated with an adsorption site or a fast exchange of the protons
between the two positions in the =CH2- group are taken into account. A further
interpretation of this situation, however, requires additional evidence, e.g. from a study of
the distances between the =CH 2- and the =CH- protons by using lH NOESY NMR
measurements. This study is treated in the following part.

4.2. 2D NOESY NMR

It was already explained in our previous work [11], that the better spectral resolution
realized in our lH-MAS NMR measurements enabled two-dimensional lH NOESY NMR
experiments of I-butene sorbed in a NaX zeolite. Standard pulse sequences are used in
these studies and the cross-relaxation rates between the different proton pairs in the
adsorbed molecule are calculated from the cross-peak intensities in the 2D NOESY spectra
taken with different mixing times.
The cross-relaxation rates aij between two protons i and j are given by

(1)
90
with the proton Larrnor frequency 0) and the gyromagnetic ratio y of the protons. In the
simplest case of a fIxed distance rij between the nuclei and of an isotropic reorientation with
one correlation time 't, the reduced spectral density Jij(0) is given by

(2)

For the further discussion, however, the shape of the reduced correlation function is not
of interest. The only requirement is that it should be the same for all proton spin pairs in the
molecule. Hence, fast internal motions on the time scale of the correlations times for the
reorientation of the whole molecule are not considered. In particular, cross-peaks with
methyl protons will not be taken into account because of the well-known fast anisotropic
rotation of the -CH3 group. Consequently, we can only concentrate on the dependence of
the quantities aij on the factor lIr{
If we consider the cross-relaxation rates between two different spin pairs i, j and k, I we
have to expect a ratio

(3)

Let us discuss, for instance, the ratio of the cross-relaxation rates aI2/a\3 between the
=CH- proton (denoted in the following by (I), see fIg 4) and the different =CH2- protons
(2) and (3). In case of the standard geometry (e.g. for I-butenes dissolved or for I-butenes
in liquid state) we would fmd a value of 3.7. Unfortunately, the experimental estimation of
this ratio for adsorbed molecules is more complicated because the respective cross-peaks
between proton (1) and the two (only slightly different) protons (2) and (3) overlap. Hence,
the cross-peak intensity originates from the cross-relaxation of proton (I) with the two
different protons. Only an average cross-relaxation rate at can be determined which allows
to estimate an average distance r:

(4)

In our situation, we may assume identical line shapes in the two frequency dimensions.
Then it is possible to determine the ratio of cross-relaxation rates by integrating the
intensities of the cross-peaks at a certain mixing time. This ratio can be obtained by
deconvoh1tion of a (representative) cross-peak slice. Fortunately, the spectral resolution in
the 2D NOESY NMR spectra is still suffIcient (fIg 5) and hence it is possible to check
91

model A model B

Figure 4. Models for conformational changes of I-butene sorbed at an


adsorption site. The upper part shows the conformation of the molecule in
liquid state with different distances (r\2 and rl3, due to the SP2 hybridization)
between proton (1) and both protons of the =CH 2- group. For the adsorbed
state two models (A and B) are proposed which could lead to equal distances
rl2 and r13: Model A suggests a 90° twist of the =CH r group protons around
the C-C double bond. In model B an exchange of these protons is assumed
which results in an averaging of both proton positions to a position 'D' .

,
5.05 pp

Figure 5. Cross section in F2 direction of a cross-peak in a IH-NOESY


spectrum (T = 343 K, VH = 600.13 MHz, VMAS = 5 kHz) between the =CH-
group and the =CH 2- group consisting of 2 line pairs that belong to the two
protons of the =CH 2- group (see text). The splitting of each line appears due
to the 3JHwcoupiing with the =CH- group proton. A deconvolution was
performed for the two line pairs to compare the integrals. The dashed line
shows the sum of the four lines demonstrating the quite good agreement for
the simulated peaks. The integration leads to a ratio of about 1 (0.9±0.09).
92
whether cross-peaks with strongly different cross-relaxation rates al2 and al3 are
superimposed. In the present case it can be surely excluded that the both cross-relaxation
rates differ by a factor of 3.7 which has to be expected (see above) if the standard geometry
for the liquid state would be also valid for adsorbed molecules. Instead, the calculations
point clearly to a situation with about the same average distances, i.e. r ::::: rl2 ::::: rn ::::: r .
As already mentioned two models are feasible to explain these apparently equal
distances (fig. 4). For model A, a rotation of the =CH2 group by an angle of 90° around the
carbon-carbon double bond axis is supposed which seems to be not very probable.
Conformational change would lead here to the same distances between each of the two
=CH 2- protons and the =CH- proton. In model B, a fast exchange between the two protons
of the olefmic =CH2- group or, alternatively, a 1800 jump-like rotation around the carbon-
carbon double bond is considered. This process would lead to about the same average
distance (rID) between each of the two olefmic =CH2- protons and the =CH- proton (see
fig. 4). The accuracy achieved so fare, however, is not sufficient for a clear differentiation
between the models A and B. Therefore, we use the cross-relaxation rates for various cross-
peaks in order to prove various intramolecular distances by using eq. (3), viz. the distances
r12, rl3, r34, and r24' For this "internal calibration" we can assume that the distance rl4
remains the same, independent of the structure models discussed and we specify the
following cross-relaxation path ways:

• The relaxation concerning al4 takes place between the proton at position (1) and the
two equidistant protons of the aliphatic -CHz- group (4). In this case eq. (4) should
lead to the correct distance if the influence of the thermal motion is properly taken
into account. 'Here we use this quantity in order to derive a value of
F( 0) = 4.13 X 1010 m6 which will not be discussed further in this paper.
• For al2 and al3 we have complications because the cross-peaks are overlapping. But,
the protons (2) and (3) can be treated to be equidistant to proton (1) in a good
approximation as already discussed above.
• A more complicated situation occurs for the (overlapping) cross~relaxation rates a24
and a34 because each of the protons (2) and (3) interacts with two, in dependence on
the model non-equidistant protons in position (4). As a result we have to build average
distances for the cross-relaxation rates a24 and a34 similar to eq. (4).
• In the other cases the average distances match to the distances between the proton
pairs. The determined theoretical proton distances and the derived average distances
for the assumed geometry of both models are shown in Table 1.

Based on eq. (3) we are able to calculate the average proton distances (;;D and ;:;4) in
dependence on one of the distances from the experimentally determined cross-relaxation
rates (see table 2). Therefore we use ;;4 for calibration because it is the only distance that
rests equal in both models presented and which results in:
93

Weare now able to compare the results from experimental data to the distances which are
determined from the model geometry (table 1). Therefore we have to refer the latter values
as well to the calibration distance ;;4 and we get for the different models:

Model A: ~2 = ~3 = 0.93 ~4 and ;4 = 1.07 ~4


ModelB: ~D =0.87 ~4 and ;4 =1.05 ~4 .
The comparison, especially for the distances (1,2) shows that the exchange model B of
the two protons is the more favorable one. Moreover, since the measured cross-relaxation
rates oij are proportional to the sixth power of rij, they are quite sensitive to control small
differences of inter-proton distances. In particular, for the ratio 012/014 a value of 1.55
should be found to fulfill the conditions of model A which can be excluded with a high
probability by the measurements that lead to a value of2.47.

Table 1: Calculated distances for the olefmic group of I-butene for the
models discussed. The distances (for indices, see fig 4) are derived from the
geometry and are averaged (marked by a bar) by using eq. (4). For
equidistant proton pair distances, there is no difference between the average
and the simple distance. The distance between proton (2) and both protons at
position (4) differs only for model A (see fig 4). rl4 does not change in both
models and therefore is used as a reference to which the other distances are
compared.

model A: model B:

TI. = ii, = 2.97 A

TI2 = TI3 = ii, =2.78 A TID = iiD = 2.59 A

T24b = 3.5 A, T2 •• = 3.0 A, ;;;. = 3.19 A TD4. = TD4b = ;;;. = 3. \3 A

Table 2: Cross-relaxation rates for the olefmic part. of I-butene sorbed in


NaX zeolites (1 molecule per supercage, T = 295 K) determined by fitting
cross-peak intensities from IH NOESY NMR spectra (v = 600.13 MHz) in
dependence on the mixing time. OlD denotes the sum of 012 and 013 which
cannot be separated due to the overlap of the cross-peaks between the proton
pairs (1,2) and (1,3).

(TID (= (T12 + (Tn) (TI' (T,.

0.079 0.031 0.021


94
5. Conclusion

It could be shown that the advantage of high spectral resolution achieved by means high-
resolution (HR) IH MAS-NMR spectroscopy allows a detailed study of the proton chemical
shifts for adsorbed simple olefms also for very low pore filling factors. The most interesting
fmding in these measurements is an appreciable change in the chemical shifts for the
olefmic =CH 2- group for adsorbed I-butene molecules when measuring temperature and
pore filling factor are varied. The experimental results point to a situation where the two
olefmic protons become chemically nearly equivalent for adsorbed molecules in contrast to
the behavior well-known for molecules in solution or in liquid state. The measurements
suggest either a model in which there is a fast proton· delocalization within the =CHr
group or a model with conformational changes of the molecule due to adsorption.
To elucidate this situation in more detail, 2D IN NOESY MAS NMR measurements
were carried out. The problem was here to arrive at an accuracy in the estimation of
intramolecular distances which allows further conclusions about conformational changes,
i.e. the physical nature of the changes in the chemical shifts as mentioned. For this reason
the influence of the thermal motion has been taken into account for the treatment of the
cross-relaxation data. Thus, it was possible to derive various rather precise distance values
from NMR measurements. The data support the conclusions derived from the chemical
shifts and allow the more detailed discussion of two models. For model A, a rotation of the
=CH 2 group by an angle of 900 around the carbon-carbon double bond axis is supposed.
This conformational change would lead here to the same distances between each of the two
=CH2 protons and the =CH- proton. In model B, a fast exchange between the two protons
of the olefmic =CH 2 group or, alternatively, a 1800 jump-like rotation around the carbon-
carbon double bond is considered leading as well to an equal average distance (rID) between
each of the two olefmic =CH 2 protons and the =CH- proton which differs from the distance
in Model A more than 10%. A main advantage of the analysis was the possibility to use
cross-relaxation rates for various cross-peaks in order to check the accuracy in determining
different intramolecular distances. For this "internal calibration" we can assume that at least
one distance is independent of the models discussed and, hence, it is equal to the standard
value. This situation gives the possibility to correct the influence of thermal mobility on the
cross-relaxation rates. As the consequence the accuracy is sufficiently high in order to
conclude that model B is the most probable one. It describes a proton exchange within the
molecular group =CH r -'- after adsorption.
The conclusions derived form the spectroscopic data are very interesting for a deeper
understanding of the "activation" of adsorbed olefmic molecules in the course of catalytic
transformations. In fact, we have not only found the typical chemical shift changes of
certain molecular groups due to the interaction with adsorption sites which are already well-
known [5] but also detailed changes in the bonding of the protons in the terminal group
=CH 2-. The interpretation that this change is due to a proton exchange within the terminal
olefmic group should be still checked also in comparison with the behavior of higher
olefins (being in progress). But is important to note that the following main statement can
be drawn independent of the fact that the experimental cannot strictly exclude that also a
conformational change might occur: Both interpretations point to the fact that there is a
large polarization of the molecule at the fmal olefmic position which has much similarity
with a model in which the formation of a carbenium type-like structure is suggested as a
95

transition state for olefm transformations. Even this point is an important question in the
further work of2D 'H NOESY NMR on higher olefms.

6. Acknowledgement

The authors are very indebted Dr. W. Bohlmann for suggestions and comments. The
support of the Deutsche Forschungsgemeinschaft within the SFB 294 is gratefully
acknowledged. The authors would like to thank also Mr. L. Moschkowitz for the
preparation of the samples used in the experiments.

7. References

1. VanderHart, D.L. (1981) Resolution in 13C NMR of organic solids using high-power proton decoupling
and magic-angle sample spinning, J. Magn. Reson. 44, 361-401 .
2. VanderHart, D.L. (1996) Magnetic susceptibility & high resolution NMR of liquids and solids, John
Wiley & Sons, New York.
3. Schwerk, U., Michel, D. and Pruski, M. (1996) Local magnetic field distribution in a polycrystalline
sample exposed to a strong magnetic field, J. Magn. Reson. 119,157-164.
4. Roland, J. and Michel, D. (2000) High-resolution IH NMR spectroscopy of absorbed molecules in the
presence of strong magnetic fields, Magn. Reson. Chem. 38,587.
5. Engelhardt, G. and Michel, D. (1987) High-resolution solid-states NMR of silicates and zeolites, John
Wiley & Sons, Chichester.
6. Rutenbeck, D., Papp, H., Freude, D. and Schwieger, W. (2001) Investigations on the reaction
mechanism of the skeletal isomerization of n-butenes to isobutene Part I. Reaction mechanism on H-
ZSM-5 zeolites, Applied Catalysis A-General 206, 57-66.
7. Shdanov, S.P., Kovoshchov, S.S. and Samuelevich, N.N. (1981) Synthetic Zeolites, Khimiya, Moscow.
8. Freude, D., Unger, M. and Pfeifer, H. (1982) Study of Bronsted acidity of zeolites using high-resolution
proton magnetic resonance with magic-angle spinning Chem. Phys. Lett. 91,307.
9. Bohlmann, W., Michel, D. and Roland, J. (1999) IH NMR studies of the mobility of olefins adsorbed in
zeolites of type NaX, Magn. Reson. Chem. 37, 126-134.
10. Michel, D., Germanus, A. and Pfeifer, H. (1982) Nitrogen-IS nuclear magnetic resonance spectroscopy
of adsorbed molecules, J. Chem. Soc., Faraday Trans. 78,237-254.
II . Schwerk, U. and Michel, D. (1996) Methods for the determination of molecular mobility of adsorbed
molecules based on high resolution NMR, Colloids and Surfaces A: Physicochemical and Engineering
Aspects 115,267-276. .
"LIGHTING UP" NMR AND MRI IN COLLOIDAL AND INTERFACIAL SYSTEMS

A. PINES, J.W. LOGAN, M.M. SPENCE


Materials Sciences Division, Lawrence Berkeley National Laboratory, and
Department of Chemistry, University of California at Berkeley,
Berkeley, CA 94720
USA

1. ABSTRACT

By means of optical pumping with laser light, the nuclear spin polarization of gaseous
xenon can be enhanced by many orders of magnitude. The enhanced polarization has
allowed an extension of the pioneering experiments of Fraissard and coworkers to novel
applications of NMR and MRI in chemistry, materials science and biomedicine.
Examples are presented of developments and applications of laser-polarized xenon
NMR and MRI on distance scales from nanometers to meters. The size of the xenon
atom is similar to that of small organic molecules, such as methane, yet the nuclear
magnetic resonance (NMR) signal from xenon proves a more sensitive probe for the
local environment. Laser-polarized xenon NMR has been used, in collaboration with
Sozzani and coworkers, to investigate the interactions present in an effectively one-
dimensional gas phase inside nanochannels. Small changes in channel size and/or
structure lead to very different modes of diffusion. Optically pumped Xe NMR can
distinguish between these different diffusion modes out to unparalleled time scales
(several tens of seconds). These studies are particularly useful for gaining a fundamental
understanding of tile laws that govern heterogenous mass transport such as gas transport
into porous catalysts or molecular sieves, or liquid transport through pore-forming
transmembrane proteins in biological systems. The understanding of mass transport
inside microporous materials is crucial for many industrial and commercial processes.
Recent experiments will also be described in which xenon has been used to investigate
the cavities of biological nanosystems and in which polarization has been transferred to
molecules on surfaces and in solution. As an example, in collaboration with Wemmer
and coworkers, xenon has been used as a molecular probe to investigate the
hydrophobic surfaces and interiors of macrocyclic molecules and proteins; recent results
show evidence for binding of xenon to the outside of a protein, a proposed cause of the
anesthetic mechanism of xenon. Indeed, localized injection of polarized xenon solutions
into human blood has provided observations of the real-time process of xenon
penetrating red blood cells. The injection technique also makes it possible to provide
enhanced magnetic resonance images of localized areas in living organisms.
Furthermore, the use of laser-polarized xenon also opens an exciting new frontier in the
possibility of "functionalized xenon" as a biosensor of analytes and metabolites in
97
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 97-106.
© 2002 Kluwer Academic Publishers.
98
chemistry, materials science and biomedicine. The novel biosensor offers advantages of
multiplexing capabilities and the possibility of detection in-vivo.

2. FIGURES

293 K
/ /\
273 K
./ J\
243K
A
223 K

203 K

120 100 80 ppm


Figure 1. 129Xe line shapes for xenon gas inside one-dimensional channels of Tris(o-
phenylenedioxy)cyc\otriphosphazene (TPP) at different temperatures. Although the xenon is in the gas phase,
the interactions of xenon with the channel wall or other xenon atoms in the same channel distort the xenon
chemical shift anisotropy (CSA) tensor, resulting in an orientation dependent chemical shift. Increasing the
density of xenon in the channels, either by lowering the temperature, or increasing the mole fraction in the gas
mixture (data not shown for the latter case), induces a change in the sign of the (CSA). The progression from
an oblate to a prolate symmetry of the CSA is smooth, passing through a region where the line shape is
isotropic. The line shapes indicate the xenon atom have cylindrical symmetry imposed upon them from the
interplay of the xenon atoms with the waIls of the channels (forces perpendicular to the channel axis) and the
adjacent xenon atoms in the channels (forces parallel to the channel axis). Figure adapted with permission
from reference 16.
99

120 110 100 90 80


ppm

120 110 100 90 80


ppm

II; o
o
Figure 2. (A) At high temperatures or low pressures of xenon in the channels, the xenon atoms wi ll be far
apart from one another. Under these circumstances, the bonding interaction between the xenon atom and the
1t-eleclrons in the walls of the channels yield a CSA tensor with an effective oblate symmetry. resulting in a
positive CSA. (8) At low temperatures or high pressures, the xenon atoms will be close to each other. In this
case, the xenon- xenon Van der Waals interaction dominates the bonding interactions wi th the walls. and
yielding a prolate symmetry in the CSA tensor and a negative CSA value.
100

time

Diffusion

time

time

Figure 3. Gas transpon caused by flow may have one of three different displacement profiles. The first is a
linear displacement in time due to coherent motion (top). The second involves displacement that scales as the
square rOOt of time due to [normal] diffusion (even in the case of unidirectional diffusion) (middle). The last
case is single-file diffusion (bottom). Single-file diffusion has a displacement proponional to the founh root
of time (for shon times) . The slower displacement is due to the inability of the atoms to "jump over", or
switch places, with adjacent atoms. In small channels. such as those in TPP, single-file diffusion is observed.
Figure adapted with permission from reference 6.
101

,........,
.....
J!3
§

~
.....
of
......
~

a
-,
'r;}
=
"E
G.)

0
0 40 80 120 160
time [s]
Figure 4. (A) The structure of the one-dimensional nanochannels in TPP (top view). (B) The circles indicate
the intensity of the xenon resonance from the xenon gas inside the TPP channels as a function of build-up
time. The solid line is a nonlinear least-squares regression of the data to a single-file diffusion model. The
dashed line is the nonlinear least-squares regression of the data to a normal diffusion model. It is clear that the
xenon atoms inside the TPP channels exhibit single-file diffusion behavior. Prior to the start of the build-up
period. a series of 7tl2 saturation pulses were applied to the sample to destroy all of the xenon polarization.
Next, optically polarized xenon gas flows into the sample and begins to diffuse into the TPP channels. The
signal intensity as a function of build-up time increases for short times, and then levels off. as the rate of
polarized xenon entering the channels equals the rate of polarization loss due to relaxation in the channels.
The methods used in this work allow one to determine the mode of diffusion up to tens of seconds. whereas
conventional pulsed-field gradient methods can only be used up to -1 s. Figure adapted with permission from
reference 6.
102

Xegas

Xegas

60 o -60 -120 -180


Chemical Shift (ppm)
Figure 5. Typical l~e NMR spectra of thennally polarized xenon gas dissolved (A) ~O and (B) a 10 mM
metrnyoglobin solution under -5 atrn of xenon overpressure are shown. The upfield peak corresponds to
xenon gas filling a capillary tube placed within the samples for in situ referencing. The addition of protein
both shifts and significantly broadens the dissolved xenon resonance, indicating fast exchange between protein
and solvent environments. Xenon has been shown to bind to small hydrophobic pockets ubiquitous in
proteins, and has been used as a probe of these internal binding sites. figure adapted with permission from
reference 12.
103

!
A
10

8
'f
E0.
E: 6 !
"0

.
OJ
~
VI
.n
4
~


0
c.o
2


o0• 5 10 15
Concentration Metmyoglobin (mM)

8
4
I
Ea. 3
E:
"0
. •
.
1: 2
VI
.n

0
c.o

2 3 4

Concentration
Denatured Metmyoglobin (mM)

Figure 6. Both specific and non-specific interactions between xenon and myoglobin are noted. To observe the
effect of myoglobin on the xenon chemical shift. the xenon resonance is measured in a series of solutions of
varying protein concentration (A). The chemical shift moves downfield from that in water with increasing
protein concentration but seems to have an upfield component that becomes significant at higher
concentrations. The initial downfield shift reflects non-specific surface interactions between free xenon and
the myoglobin. The upfield component appears as more protein is added and more xenon samples the internal
binding site. In denatured myoglobin. the specific binding site is removed and only non-specific interactions
exist. The titration of denatured myoglobin (B) shows a linear downfield trend over the accessible protein
concentrations. Figure adapted with permission from reference 12.
104

linker
ligand

cage
target

Figure 7, To make xenon detect only specific binding events. rather than al l xenon-protein interactions in the
solution. xenon must be targeted to a particular protein by functionalizing the xenon , This is accomplished by
trapping xenon in a cage and using a linker to connect the xenon cage to a ligand that binds strongly to the
target. A a prototype biosensor. a water-soluble cryptophane-A cage was attached to biotin via a linker,
Biotin is a smal l Vitamin which binds to the protein avidin with one of the largest protein-ligand binding
constants known (K-I 0'5 M·'). Figure adapted with permission from reference 17 .
105

I I I I I
75 74 73 72 71
Figure 8. Spectrum A shows the functionalized xenon without avidin. with the more intense peak
corresponding to functionalized xenon and the smaller peak corresponding to xenon in the bare cage. serving
as both a chemical shift and signal intensity reference . Spectrum B shows the spectrum upon the addition of
-80 nmol of avidin monomer. A third peak. corresponding to functionalized xenon bound to avidin. has
appeared and the unbound functionalized xenon peak has decreased in intensity. Figure adapted with
permission from reference 17.
106

3. References
I. Albert, M. S., Cates, G. D., Driehuys, B., Happer, W., Saam, B., Springer, C. S. and Wishnia, A.
(1994) Biological Magnetic Resonance Imaging using laser polarized Xe-129, Nature. 370. 199-
201.
2. Eckenhoff, R.G. and Johansson, 1.S. (1997) Molecular interactions between inhaled anesthetics
and proteins, Pharmacol. Rev.• 49. 343-367.
3. Fraissard, 1. and Ito, T. (1988) I~e N.M.R. study of adsorbed xenon: A new method for studying
zeolites and metal-zeolites, Zeolites, 76, 350-361.
4. Hahn, K.; Karger J., and Kukla, V. (1996) Single-file diffusion observation, Phys. Rev. Lett., 76,
2762-2765.
5. Klirger J., Petzold, M., Pfeifer, H., Ernst, S., and Weitekarnp, 1. (1992) Single-file Diffusion and
reaction in Zeolites, J. Catal., 136, 283-299.
6. Meersmann, T., Logan, J.W., Simonutti, R, Caldarelli, S. , Comotti, A., Sozzani, P., Kaiser, L.G.,
and Pines, A. (2000) Exploring Single-File Diffusion in One-Dimensional Nanochannels by Laser-
Polarized I~e NMR Spectroscopy, J. Phys. Chem. A, 104, 11665-11670.
7. Prange, T., Schiltz, M., Pemot, L., Colloc'h, N., Longhi, S., Bourget, W. and Fourme, R. (1998)
Exploring hydrophobic binding sites in proteins with xenon or krypton. Struct. Funct. Genet.. 30,
61-73.
8. Raftery, D., Long. H., Meersmann, T., Grandinetti, PJ., Reven, R., and Pines, A. (1991) High-field
NMR of Adsorbed Xenon Polarized by Laser Pumping, Phys. Rev. Lett., 66, 584-587.
9. Ripmeester, J.A. (1982) Nuclear Shielding of Trapped Xenon Obtained by Proton-Enhanced,
Magic-Angle Spinning Xe-129 NMR Spectroscopy, J. Am. Chem. Soc., 104, 289-290.
10. ROdenbeck, C., Klirger J., and Hahn, K. (1997) Exact analytical deScription of tracer exchange and
particle conversion in single-file systems, Phys. Rev. E, 79, 5697-5712.
II. Rubin S.M., Spence M.M., Dimitrov I.E., Ruiz EJ., Pines A., and Wemmer D.E. (2001) Detection
of a conformational change in maltose binding protein by I~e NMR. J. Am. Chem. Soc., 123.
8616-8617.
12. Rubin, S.M., Spence, M.M., Goodson, B.M., Wemmer, D.E., and Pines, A. (2000) Evidence of
nonspecific surface interactions between laser-polarized xenon and myoglobin in solution. Proc.
Nat. Acad. Sci., 17, 9472-9475.
13. Schoenborn, B.P., Watson, H.C. and Kendrew, J.C. (1965) Binding of xenon to spenn whale
myoglobin. Nature. 1965,207,28-30.
14. Sholl, D.S., and Lee, C.K. (2000) Influences of concerted cluster diffusion on single-file diffusion
ofCF. in A1P04-5 and Xe in AIP04-31, J.Chem. Pbys., 112, 817-824.
15. Song, Y.Q., Goodson, B.M., and Pines, A. (1999) NMR and MRI using laser-polarized xenon,
Spectroscopy, 14. 26-33. '
16. Sozzani, P., Comotti, A., Simonutti, R, Meersmann, T., Logan, J.W., and Pines, A. (2000) A
Porous Crystalline Molecular Solid Explored by Hyperpolarized Xenon, Angew. Chem. In!. Ed.,
39, 2695-2698.
17. Spence, M.M .• Rubin, S.M., Dimitrov, I.E., Ruiz,'EJ., Wemmer, D.E., Pines, A., Yao. S.Q., Feng,
T., and Schultz, P.G. (2001) Functionalized xenon as a biosensor. Proc. Nat, Acad. Sci., 98, 10654-
10657.
PROGRESS IN mGH RESOLUTION SOLID STAIE NMR OF
QUADRUPOLAR NUCLEI: APPLICATIONS TO POROUS MATERIALS
AND CATALYSTS
M. PRUSKIa , J. P. AMOUREUX b and C. FERNANDEZc
DAmes Laboratory, Iowa State University, Ames, lA, 50011, USA
bUniversite de Lil/e, LDSMM, F-59655 Villeneuve d'Ascq Cedex, France
cUniversite de Caen, Laboratoire Catalyse & Spectrochimie,14050 Caen Cedex, France

Several new experiments are reviewed that combine multiple quantum magic angle spinning
(MQMAS) NMR with cross polarization (CP) or heteronuclear recoupling via REDOR. The
advantage of these techniques is that they provide direct inference of the connectivities between spin-
112 and half-integer quadrupolar nuclei in the isotropic spectra of the latter. Additionally, the MQ-
REDOR methods yield the internuclear distances in these systems. We have demonstrated the
capabilities of these techniques in the studies of IHYAI and 19F_27AI spin pairs in aluminophosphate
molecular sieves.

KEY WORDS: NMR, quadrupolar nuclei, connectivities, MQMAS, HETCOR.

INTRODUCTION

The development of double rotation (DOR) [1], dynamic angle spinning (DAS) [2]
and multiple quantum magic angle spinning (MQMAS) [3] allowed for measurements of
high-resolution spectra of half-integer quadrupolar nuclei in solids. These spectra are isotropic
in the sense that the second order anisotropic quadrupolar contribution is averaged out.
Several methods that measure the interatomic connectivities between these spins (e.g., 23Na,
27AI) and spin-112 nuclei (e.g., IH, 19F) under high-resolution followed the inception of these
techniques. Most of the methods utilize indirect excitation via cross polarization (CP) for
spectral editing of the DOR [4] and MQMAS [5-9] spectra or for the DAS- [10] and
MQMAS-based [11,12] heteronuclear correlation (HETCOR) experiments. As the spin
dynamics of the CP transfer that involves quadrupolar nuclei is very complex [13,14], the CP-
based methods suffer from severe quantitative uncertainties. However, demonstration of
internuclear connectivities between spin-1/2 and quadrupolar nuclei can also be made by
reintroducing the heteronuclear dipolar dephasing between these spins using RF pulses that
are synchronized with the MAS rotor [15]. For example, rotational echo double resonance
(REDOR), was combined with the high-resolution capabilities of MQMAS to provide a new
spectroscopic tool for measurement of dipolar interactions and internuclear distances [16]. In
this paper, we briefly review these experiments and their applications to porous materials and
catalysts.

MQMAS WITH CROSS POLARIZATION

The dipolar coupling between two spins can be used to indirectly produce transverse
nuclear magnetization of the observed spins (S) via cross polarization transfer from another
107
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 107-113.
© 2002 Kluwer Academic Publishers.
108

spin system (1). In the standard spin-II2


~ spin-lI2 CP experiments, polarization (a)
is usually transferred from the abundant to
rare spins in order to enhance their
nuclear magnetization [17]. The spin
dynamics that is involved in polarization
S
transfer in such systems is well
understood, and obtaining quantitative
spectra is relatively easy. When
appropriate line narrowing methods are
applied to both spins, this polarization can
be further employed to obtain high-
resolution HETCOR spectra. Such
experiments are routinely used to probe (b)
the proximity of different spin-I/2 nuclei III~II
S .!:!:~~~=--{CP]
in solids [18].
For reasons described elsewhere
[4,13 ,14,19], the CP between spin-112 and
quadrupolar nuclei does not generally
[!!"]"a 12

lead to enhanced sensitivity. However, it


remains a useful method to study the
correlations between coupled nuclei. The
fIrst high-resolution HETCOR NMR
spectroscopy involving quadrupolar
nuclei used DAS to average the Figure 1: Pulse sequences and coherence
anisotropic second-order quadrupolar transfer pathways for I ~ S (a) and S ~ I (b)
interaction [10] . The development of I QCP-3QMAS methods. The I ~ S MQCP·
MQMAS spectroscopy was followed by MQMAS experiment (not shown) uses direct
efforts to couple this method with polarization of MQ coherences (Le. the tl
polarization transfer from spin-II2 nuclei domain follows immediately the CP period)
into the single-quantum (1 QCP-MQMAS) [7-9]. The ratio R defines the time of formation
[5,6] or multiple-quantum (MQCP- of an isotropic echo in the t2 domain [3].
MQMAS) [7-9] coherences. The pulse
sequences and coherence transfer pathways used in some of the implementations of these
techniques are shown in Figure 1. The pulse sequence of Figure la uses the spin-112 nuclei as
the original source of magnetization (l ~ S), whereas the experiment shown in Figure 1b
applies the reverse process (S ~ I).
It was shown that the IQCP-3QMAS experiment is possible between 19F and 27Al
nuclei in fluorinated triclinic chabazite-like AlP04 aluminophosphate (Figures 2a,b). This
material contains one octahedral site Alt coordinated to four oxygen and two fluorine atoms,
and two sites labeled Ah and Ab representing aluminum in AI04 tetrahedra. The aluminum
site that is bonded to fluorine was easily distinguished in the 1QCP-3QMAS spectrum, see
Figure 2b [5]. A similar experiment was applied for probing the interaction of IH nuclei in
water with the framework aluminum of fully rehydrated AlP04-11 aluminophosphate [6]. As
Figures 2c,d demonstrate, lH~27AI IQCP-3QMAS allowed for distinguishing the Al sites
that are most and least susceptible to hydration. Such distinction would not be possible
without the separation of 27Al resonances under high-resolution conditions. Examples of
3QCP-3QMAS, 5QCP-5QMAS and even 7QCP-7QMAS spectra have been demonstrated by
Ashbrook et al. [7], Rovnyak et al. [8] and Lim et al. [9].
109

la ) (b )

"l
(e) (d)

" ~"-JA_O>_.__ ...........-I

Figure 2: (a), (b) - 2D NMR spectra of 27AI in AIP04-CHA and obtained using 3QMAS and 19p_
27 AI I QCP-3QMAS, respectively. The octahedral site All is directly coordinated to fluorine. (c),
(d) - similar 2D NMR spectra of 27 AI obtained in AIP04-11 using 3QMAS and IH~27 AI 1QCP-
3QMAS. The four- and six-coordinated AI species that exhibit the strongest interaction with
adsorbed water are marked with asterisks in spectrum (c). Spectra (a), (b) were taken at 9.4 T on a
Chemagnetics Infinity spectrometer using a 3.2 mm CPMAS probehead. The Hartman-Hahn
condition was established with an RF field of -5 kHz under MAS speed of 20 kHz. 19F
decoupling was applied during evolution and acquisition times using an RF field of -80 kHz.
Spectra (c), (d) were acquired with a Bruker ASX-400 spectrometer and a 4 mm CPMAS
probehead under similar conditions, except for the spinning speed, which was 12 kHz.

We note that the experimental scheme of Figure la can be easily modified to achieve a
three-dimensional high-resolution HETCOR experiment by inserting an additional evolution
period between the initial (1t/2)\ pulse and the CP transfer. However, due to unfavorable
relaxation rates such experiment would require a very long acquisition time. Therefore, it is
advantageous for HETCOR NMR to use the two-dimensional (2D) scheme shown in Figure
1b, which utilizes the S -+ I polarization transfer from fast relaxing quadrupolar nuclei. For
example, a 2D HETCOR MQMAS NMR spectrum of 23Na and 31 p in Na3P309 was obtained
using such a scheme by Wang et aI. [11]. A similar spectrum was obtained earlier by Jarvie et
al. using the 2D DAS-CPMAS experiment [10]. An extension of Wang's method to nuclei
with spin value larger than 3/2 has been recently reported [12]. The pulse sequence is similar
to that of Figure Ib, but the coherence pathway must be modified to 0 -+ +3 (-5) -+ -1 -+ CP
for the triple (quintuple) version of the experiment.
Application of the above methods to half-integer quadrupolar nuclei is often difficult
due to the convoluted spin dynamics involved in both the spin locking and the cross
polarization processes under static and MAS conditions. These dynamics are strongly
anisotropic with respect to crystallite orientation and depend on the relative size of the
quadrupole frequency, the amplitudes of the RF field applied to the I and S spins, the spinning
speed and the resonance offsets [4,13,14].
110

MQMAS WITH DIPOLAR RECOUPLING VIA REDOR

The quantitative distortions that


resuh from the CP transfer can be avoided (a)
by combining the high-resolution
capabilities of MQMAS with the direct
I III1I1
determination of hetero-nuclear dipolar
couplings via REDOR. Figures 3a and 3b
show the schematic diagrams of two MQ- o (n+I)1, 2(n+I);:' f'
+3
REDOR experiments, referred to as MQ-
t2-REDOR and MQ-tl-REDOR, .
.,.--~"-""-" " '-"'-"-~"""""""""'"

respectively. The MQ-trREDOR


technique [20], employs two strong and
two selective RF pulses at the Larmor
frequency of the S spins with phases
cycled to select the 0 ~ ±p ~ 0 ~ +1 ~ - III III (b)

I I I
1 coherence pathway. When spaced by an I
integer number of rotor periods, the ---~I""(--""''t-)IL-.-I'(..--.,)r"'------
selective pulses create two windows in tl n n RII
which the sequences of 7t pulses are S ---. r;0T~-. 12
applied to the spin-112 (I) nuclei. Similar to "--.....---1-1_ _ '-'~I"-_--,~~~+-....<!~ntHn~"o....,
o (n+I)1, 2(n+I)I, •H
the standard REDOR experiment [16], the
role of the 7t pulses is to prevent MAS
from refocusing the dephasing due to the
dipolar interaction between spins I and S.
The spectral editing capabilities of
the MQ-t2-REDOR methods are illustrated
1/21 l!2l ,l21 1!2
in Figure 4, which shows several 27 Al (c)
t
NMR spectra of the sample of AlP04 -CHA
aluminophosphate [20). The two I( n.1 1
tetrahedral sites are not resolved by MAS
alone, but are easily distinguished by using
the MQMAS method, compare spectra (a)
and (b). The so-called REDOR difference,
obtained by subtracting the 3Q-t2-MQMAS
spectra obtained without and with the
dephasing pulses is shown in Figures 4c-d.
As expected, only the resonance from
fluorinated site (All) remains in the Figure 3: Pulse sequences used in the 3Q-t2-
spectrum of Figure 4c. However, the REDOR (a), 3Q-t l -REDOR (b) and DD-MQMAS
applications of the MQ-trREDOR method experiments. In these figures, n is an integer
extend beyond spectral editing. By number and tr denotes the rotor period.
changing the number of rotor cycles, one
can measure the REDOR curves [16] (not shown) for all three Al species and estimate their
distances from the nearest fluorine [20]. The obtained results, rAl,_F = 1.88 (±O.02) A, rAI2 - F =
4.1 (±O.2) A and rAlJ - F = 4.7 (±O.2) A, are in good agreement with the synchrotron radiation
data (1.88 A, 3.94 Aand 5.72 A, respectively).
III

The MQ-tl-REDOR experiment (see


Figure 3b), employs the REDOR sequence
during multiple rather than single quantum
evolution [21]. Because the dipolar effect is
enhanced by a factor of p during the p-quantum (a)
evolution, this technique has potential for
improved sensitivity toward weak dipolar
interactions and therefore long internuc1ei
distances. Indeed, the increased recoupling
power of MQ-tl-REDOR is well demonstrated
by comparing spectra (d) and (e) of Figure 4,
which were taken with the same number of
recoupling pulses. By measuring a complete
MQ-tl-REDOR curve, it was determined that the
apparent size of 19FYAl coupling in AIP04-
CRA was higher by a factor of 3 when compared
with the data obtained with 3Q-trREDOR [21].
Although the concept of dipolar
recoupling during MQ evolution is attractive, the (d)
3Q-tl-REDOR experiment has potential
limitations, which include loss of sensitivity due
to the more complex coherence pathway
employed by this method. In case of 27 Al in
AIP04 -CRA, the SIN was lowered by a factor of (e)

-5 upon introducing the additional ±3 --t +3 step


in the coherence pathway. Another limitation
may arise from non-uniform response of the
quadrupolar spins to the RF pulses used in the
MQMAS experiment. One may expect an
~~~~~~~~~~~

enhanced effect of the orientational anisotropy of 50 25 o


MQ excitation on the MQ-tl-REDOR curve, ppm from A1(t-W)e'+
especially in samples with significant CQ values.
However. both experimental and theoretical Figure 4: 27 AI NMR spectra of tluorinated
results show that the experiment is relatively AIP04-CHA aluminophosphate: (a) MAS;
robust. The numerical simulations indicate that (b) 3QMAS; (c) 3Q-h-REDOR difference
fitting the 3Q-tl-REDOR data with the standard for n = I; (d) 3Q-h-REDOR difference for
formula used for spin-1/2 pairs seems justified in n= 12; (e) 3Q-tl-REDOR difference for n=
a surprisingly wide range of experimental 12 and (f) DD-MQMAS. In figures (b)-(t),
conditions [21]. only the isotropic projections of sheared 2D
Finally, we have proposed another spectra are shown. The spectra were
experiment, referred to as MQMAS with dipolar acquired at 9.4 T using a Chemagnetics 3.2-
dephasing (DD-MQMAS) [19]. It uses the mm MAS probehead and a MAS speed of
standard z-filtered MQMAS scheme for spins S, 20 kHz [19-21].
while the dipolar recoupling with spins I is
accomplished using a series of 1t pulses applied synchronously with the rotor frequency to the
I spins during the multiple quantum evolution (Figure 3c). The simpler 0 --t ±p --t 0 --t -I
coherence pathway that is used here results in line broadening, but allows for fast, yet only
qualitative, determination of the internuclear connectivities in well crystallized compounds.
Again, the same AIP04 -CHA sample was used to test this method (Figure 4f). Clearly, the All
112

site resonance is broadened beyond observation in this spectrum, thus the connectIvity
between this site and fluorine is demonstrated. It is noted that dipolar recoupling also affects
the isotropic linewidths of the remaining two sites, leading to a single broadened line, which
is the shortcoming of this scheme.

CONCLUSION

The capabilities of MQMAS can be extended by combining it with cross polarization


(CP) and with rotational-echo double resonance (REDOR). These techniques provide direct
information about the connectivities and distances between spin-II2 and quadrupolar nuclei in
solids in the isotropic spectra of quadrupolar nuclei, i.e. without the limitations in resolution
imposed by MAS. Additionally, the MQ-REDOR methods yielded the distances between
interacting spins. We have demonstrated the applications ofCP-MQMAS and MQ-REDOR in
the studies of IH)1AI and 19F)7Al spin pairs in aluminophosphate molecular sieves. Future
applications will include determination of the arrangement of molecules and atoms in
adsorbent/adsorbate systems.

Acknowledgment. This research was supported at Ames Laboratory by the U.S. Department of
Energy, Office of Basic Energy Sciences, Division of Chemical Sciences, under Contract W-7405-
Eng-82 and at the University ofLille by the Region Nord-Pas de Calais.

REFERENCES

(l) Samoson, A., Lippmaa, E., and Pines A.. (1988) High-resolution solid-state NMR averaging of 2nd_
order effects by means of a double-rotor. Mol. Phys., 65, 1013.
(2) Mueller, K.T., Sun, B.Q., Chingas, G.C., Zwanziger, J.W., Terao, T., and Pines, A. (1990)
Dynamic-angle spinning of quadrupolar nuclei. 1. Magn. Reson., 86, 470.
(3) Frydman, L. and Harwood, I.S. (1995) High-resolution NMR spectra of quadrupolar nuclei. 1. Am.
Chern. Soc., 117, 5367.
(4) Sun, S., Stephen, J.T., Porter, L.D., and Wu. Y. (1995) Rotation-induced resonance and second-
order quadrupolar effects on spin locking of half-integer quadrupolar nuclei. 1. Magn. Reson.,
A1l6,181.
(5) Pruski, M., Lang, D.P., Fernandez, C., and Amoureux, J.-P. (1997) Multiple-quantum MAS NMR
with cross-polarization: spectral editing of high-resolution spectra of quadrupolar nuclei. Solid
State Nucl. Magn. Reson. , 7, 327.
(6) Fernandez, c., Delevoye, L., Amoureux, J.-P., Lang, D.P., and Pruski, M. (1997) 27AL-IH cross-
polarization triple-quantum MAS NMR. 1. Am. Chern. Soc., 119, 685.
(7) Ashbrook, S.E. and Wimperis, S. (2000) Single- and multiple-quantum cross-polarization in NMR
of quadrupolar nuclei in static samples. Mol. Phys., 98, I.
(8) Rovnyak, D., Baldus, M., and Griffin, R.G. (2000) Multiple-quantum cross polarization in
quadrupolar spin systems during magic-angle spinning. 1. Magn. Reson., 142, 145.
(9) Lim, K.H. and Grey, C.P. (2000) 19f/~a multiple quantum cross polarization in solids. 1. Chern.
Phys., 112, 7490.
(10) Jarvie, T.P., Wens low, R.M., and Mueller, K.T. (1995) High-resolution Solid-State heteronuclear
correlation NMR for quadrupolar and spin-!It nuclei. 1. Am: Chern. Soc., 117,570.
(11) Wang, S.H., De Paul, S.M., and Bull, L.M. (1997) High-resolution heteronuclear correlation
between quadrupolar and spin !It nuclei using MQMAS. 1. Magn. Reson., 125, 364.
(12) Fernandez, C., Morais, C., Rocha, J., and Pruski, M. (2001) High-resolution heteronuclear
correlation spectra between 31p and 27Al in microporous aluminophosphates. Solid State Nucl.
Magn. Reson., submitted.
(13) Vega, A.I. (1992) MAS NMR spin-locking of half-integer quadrupolar nuclei. 1. Magn. Reson.,
96, 50.
1I3

(14) Vega, A.J.(1992) CPIMAS ofquadrupolar S=312 nuclei. Solid State Nucl Magn. Reson., 1, 17.
(IS) Schaefer, 1. (1996). in 'Encyclopedia of Magnetic Resonance', eds. D.M. Grant and R.K. Harris,
John Wiley & Sons, Chichester, Vol. 6, p. 3977.
(16) Gullion, T. and Schaefer, J. (1989) Rotational-Echo Double Resonance NMR. J. Magn. Reson.,
81,196.
(17) Pines A., Gibby, M.G., and Waugh, ].S. (1973) Proton-enhanced NMR of dilute spins in Solids.
J. Chem. Phys., 59, 569.
(18) Burum D.P. and Bielecki, A. (1991) An improved experiment for heteronuclear Correlation 20
NMR in Solids. J. Magn. Reson., 94, 645.
(19) Pruski M., Fernandez, C., Lang, D.P., and Amoureux, J.-P. (J999) Measurement of interatomic
connectivities in molecular sieves using MQMAS-based methods. Catalysis Today, 49, 401 .
(20) Fernandez, C., Lang, D.P., Amoureux, J.-P., and Pruski, M. (1998) Measurement of heteronuclear
dipolar interactions between quadrupolar and spin 112 nuclei in solids by MQ-REDOR NMR. J.
Am. Chem. Soc., 120, 2672.
(21) Pruski M., Bailly, A., Lang, D.P., Amoureux, J.-P., and Fernandez, C. (1999) Measurement of
distances between spin 1/2 and quadrupolar nuclei by using REDOR during multiple-quantum
evolution. Chem. Phys. Lell., 307, 35.
APPLICATIONS OF HYPERPOLARIZED 1l'xE NMR SPECTROSCOPY TO
THE STUDY OF MATERIALS

I. L. MOUDRAKOVSKI, A. V. NOSSOV, V. V. lERSKIKH, S. LANG, E.


B. BROUWER, D. V. SOLDATOV, C. I. RATCLIFFE AND J. A.
RIPMEESTER
Steacie Institute for Molecular Sciences, National Research Council of
Canada, Ottawa, Ontario, Canada KiA OR6

1. Introduction

The use of 129Xe NMR spectroscopy in the study of porous materials is by now a well-
established method that has been reviewed many times [1-3]. The main idea is that the
chemical shift tensor of Xe nuclides is extremely sensitive to local surroundings,
primarily because of the large number of polarizable electrons [4, 5]. For small pore
systems this is best summarized by stating that the isotropic shift is dependent on the size
of the void space available to a xenon atom, whereas the anisotropic shift is sensitive to
shape [4, 5]. These simple ideas are often modified by the need to take into account
dynamics as well as Xe-Xe interactions (especially in large-pore materials), and it must
be considered that if an isotropic line is obselVed in non-cubic solids, this can only be
ascribed to the presence of fast dynamic processes [6-8]. As such, one must take into
account the pore volume sampled by a xenon atom on the appropriate time scale in order
to understand chemical shifts and relaxation parameters [6-8]. Dynamic processes also
lend themselves to study, but only as equilibrium processes, eg by 2D EXSY or
polarization transfer experiments [9]. As in the case of many NMR applications, the
utility of the technique is limited by the need to acquire sufficient signal in a time that is
short compared to the time scale of the process itself. The enhancement of the 129Xe
NMR signal by the optical pumping of Rb, followed by spin exchange with Xe, has
allowed an entire new approach to the study of materials [10-14]. Some of the main
advantages include the rapid acquisition of spectra, which allows the study of non-
equilibrium processes, and the acquisition of spectra at very low concentrations of Xe,
which allow the efficient acquisition of "zero-loading" spectra and the use of Xe as a
tracer to follow adsorption-desorption processes, phase transitions, etc.

2. Experimental

NMR experiments were carried out on Bruker spectrometers (MSL200, AMX300,


DSX400). Xe gas with the natural isotope distribution was polarized either in batch
mode or in a flow system by spin exchange with optically pumped Rb (30W diode laser-
115
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 115-122.
© 2002 Kluwer Academic Publishers.
116

Opto-Power Co) [15-17]. In batch mode experiments, polarization levels of 6-7% were
achieved, and for these experiments the hypeIpolarized (HP) xenon gas was transferred to
a glass vacuum line assembled in the bore of the magnet for transfer to the sample in the
probe. For flow experiments, equipment similar to that descnbed in ref. [17] was used,
with the pumping cell in the fringe field of the magnet. A gas mixture consisting of 1%
N2, Io/oXe and 9S% He was used, typically at a flow rate of 100 - 500 scclmin and at
polarization levels of I-6%. The gas was delivered at atmospheric pressure to the NMR
or micro- imaging probes. The MAS probes (Bruker, Chemagnetics) were modified to
allow gas delivery to the bottom of the spinner [IS]. Generally, for the NMR
experiments, short rf pulses « 1tI20) were used to monitor the polarization. For the
imaging experiments a birdcage resonator was used and a water-cooled gradient system
capable of producing a maximum gradient of 1TIm. The pulse sequence used for
chemical shift imaging consisted of an excitation pulse followed by a free precession
delay during which phase encoding was applied.

3. Results and discussion

3.1 BATCH MODE EXPERIMENTS

Batch mode experiments are carried out by producing a volume of hypeIpolarized xenon
in a pumping cell [10]. The HP Xe gas is then separated from the quenching gas and
allowed to contact the sample. If such experiments are carried out to study processes, the
time scale of the process to be studied is limited to no more than several times the
relaxation time of Xe.

3.1.1 Hydrate kinetics and metastable states


Gas hydrates are . materials where small molecules ranging in size from Ar to
methylcyclohexane are trapped in solid lattices consisting of hydrogen bonded water
molecules (porous ices) [19]. 129Xe NMR has been used since the early SO's to provide
information on structure and, since the hydrates are non-stoichiometric, also the
distribution of the guests over the different sites in hydrates [4, 5, 20). Three structural
families are known (str. I, II. H) with the type determined in the main by·the largest guest
in the hydrate [19). Xe forms str. I hydrate, as do methane, CO2, and guests up to the size
of methyl bromide. The use of Xe as anything more than a structural probe is hampered
by the long Xe relaxation times, or the need to use cross-polarization techniques [20].
However, HP Xe can be used to good advantage to study the formation of Xe hydrate
from ice and Xe gas (Fig. 1) [21, 22]. Initial experiments some years ago [21] were
followed by more extensive experiments that illustrate several features of the hydrate
formation process [22]. For instance, as in bulk gas uptake studies, the experiments
indicate that there is an induction time between initial exposure to gas and the rapid
growth of hydrate. The Xe experiments show that the distribution of Xe between small
and large cavities changes rapidly with time during the induction period, indicative of a
precursor phase that is distinctly different from the fmal product (Fig. 2). This is the first
molecular-scale technique to indicate that precursor structures other than the equilibrium
phase are important. Xe experiments also have shown there to be a surface memory
117

Large

400 300 200 100 o -100 -200


(pm)

Figure 1. Time development of the xenon spectrum for the reaction ofHP Xe with powdered ice (f=243K, Pxe
= 580 bars). The signals are assigned to Xe in the gas (.-.()ppm), and in the large cage (-150 ppm) and the small
cage (-240ppm) ofstr. I. Short RF pulses (IllS, -40) were used to minimize the effect of depolarization by the
pulse. The elapsed time is shown on the right side of the spectra. [22]

300
• Smal\cage
Cii4 B

,
A •
r
Large cage E
200 .. Gasphase ~
~
c• •••••• ........ 1\1
e>
...J
D
.!l Z2
.5100 -0
c
1\1
Q.
••••••••••••••• ::l

0 ........ 0
0

0
fP
I 1
0 200 400 600 800 1000 1200 0 200 400 600
Time, S Time, S

Figure 2 A). Integrated intensity of the peaks shown in fig. I, showing the induction period; B) The ratio of
the integral intensities of signals from xenon in the large and small cages for the run shown in Fig. 1 and
2A[22]
118

effect: after a cycle of exposure to xenon followed by desorption, the re-fonnation of Xe


hydrate takes place without an induction time.

3.1.2 Diffusion into hard porous materials


Fig. 3A shows the time dependence of the Xe spectrum when a piece of porous Vycor
glass is exposed to lIP Xe [23]. The growth is due to ingression of Xe into the pores, and
the decay is due to the spin-lattice relaxation and the cumulative effect of the rf pulses.
Fitting the growtb-decay cwves (Fig. 3B) in tenns of an appropriate model yields the
diffusion constant This can also be done by measuring the Xe profIle in the Vycor by
ID imaging techniques. In either case, the diffusion constants can be measured in a very
short time compared to pulsed field gradient methods.

T = 283K
B

o 20 40 60 80
Time, S
Figure 3 A) 12~e NMR spectra of HP Xe diffusing into a 7.35 mm od cylinder of porous Vycor at T=233K, in a
batch mode experiment with p..... = 988mbar, POOl. = 942mbar. The line at -100 ppm arises from Xe in Vycor
glass. B) 12~e NMR integrated intensities of the line arising from adsorbed xenon as a function of time after the
start of adsorption. The solid line represents a fit used to derive a diffusion constant (23)

3.2 FLOW SYSTEM EXPERIMENTS

Flow systems that produce a continuous supply of HP Xe at a 1-6% polarization level


bypass the limits imposed by spin-lattice relaxation in the batch mode approach, as the
polarization is replenished on a continuing basis. This allows the execution of more
complex experiments such as micro-imaging, or the following of processes on a time
scale much longer than the Xe T \.

3.2.1 "Zero-loading" Xe NMR spectra


The so-called "zero-loading" value of the chemical shift normally is obtained by
extrapolation from a shift versus loading plot, thus requiring a serious investment of time
as the low Xe pressures require significant signal averaging in order to obtain sufficient
SIN [1-3]. However, these zero-loading values are important, as they are characteristic of
119

the framework in the absence of Xe-Xe interactions. We have shown that effectively
zero-pressure values can be obtained by using the low partial pressure of Xe (~7 torr)
produced in a continuous flow system [24 J. The presence of 1% N2 as quencher and the
98% He have little effect on the measured chemical shift, eg, for Xe in the channels of
the porous aluminophosphate ALPO-II (fig. 4). Of course, one should always be aware
of the fact that N2 could compete with Xe for adsmption sites, thus affecting the chemical
shifts, and requiring a careful analysis of the sotption characteristics for Xe and N2 in the
material of interest. For Xe in the siliceous channel compound SSZ-24, some unique
features are seen that are not evident at higher loadings of thermally polarized xenon: a
small peak at~ 125 ppm attributed to defects, and a peak near the gas line attributed to
intetparticle Xe.

A c
Inter-crystalline
space

D
, , i i»
* iii i i i , i ,

200 150 100 50 0 -50 200 150 100 50 -50


Chemical Shift, ppm

Figure 4. 12SXe NMR spectra of HP Xe adsoroed in powder samples of ALPO-II (A, static; B, MAS) and
SSZ-24 (C, static; D, MAS). A gas mixture of I%Xe, 1%N2 and 98% He was used at a pressure of 1020-1050
mbar and a flow rate of 300 sec/sec. (* indicate spinning sidebands) [24].

3.2.2 Materials Imaging


The imaging of porous materials with Xe is an attractive possibility, as again use can be
made of the favourable chemical shift sensitivity to obtain spatially resolved chemical
shift data. However, the use of thermally polarized xenon is again a severely limiting
factor, requiring long acquisition times and high xenon pressures to obtain even partial
chemical shift resolution [25J.

A first example of the imaging of materials with full chemical shift resolution using HP
Xe produced in a continuous flow system is shown in fig. 5 [26]. The phantom is a
hollow, porous Vycor tube filled with zeolite (fig. 5A). The chemical shift spectrum is
shown in the inset on the figure, and the images at each chemical shift value are shown in
fig.5B.
120

A c
C .S. , ppm
HPX~
flow 0.6

60 .1

76.1
B
D
~
I
120
I i i
100 so 60
i i i
40 20 0
i
-20
C_IShitt. ppm

Figure 5. UP Xe chemical shift images of a phantom (A) consisting of a 6.8 mm porous Vycor tube packed
with NaY zeolite powder and placed within an open 9mm 10, 2.5 em long glass tube. A 3D image with one
spectroscopic and two spatial dimensions is shown (8) together with sections taken at the indicated chemical
shifts (C). Dashed lines indicate the walls of the tube. A 10 12~e NMR spectrum of the phantom taken in a
single scan is shown in (0). The CS image was acquired in a matrix of 40x40x128 points and a square FOV of
2Omm. At the maximum, phase gradients Ox and Oy reached 420/cm each. Two scans per each encoding step
were made. The flow rate of gas was 200 scc/min. and the total experimental time was 30 min. The 4mm thick
slice was approximately in the middle of the cylinder. The slice selection gradient Oz was 340. During the
Processing the original matrix was zero-filled to 128x128x156. The 76ppm chemical shift region corresponds
to Xe sorbed in porous Vycor, the region at 60 ppm to NaY and that at 0.6 ppm to the interstitial gas. [26]

~ Den,."~

~ ~ B
trans-isomer (50%) + cis-isomer (50%)

~-fO'm)
A
~ CH,CI. flow
CH.CI2 !Iow on
only trans-isomer (100%)
220200 ''''' '60 140120 100 80 so 40 20 0 .2O(Wm)

Figure 6 (left) Isomer distribution in the dense and open forms of [CuLz). (right,A) 12~e NMR spectra that
show the transformation of the dense form of [CuLz) to the open form. The UP Xe is added as a tracer to the
gas stream containing the transforming substance, methylene chloride, in this case. As the flow is started there
is no signal, as [CuLz] is in the dense form. After the substance is transformed to the open form by methylene
chloride, the newly created pore space becomes visible as the methylene chloride is stripped out of the channels.
Application of heat (70°C) transforms the open form back to the dense form. (28)
121

3.2.3 Adsorption/desorption and phase transitions


One goal in the design and characterization of new materials is to produce "smart"
materials. This means that a material's functional behaviour can be controlled in some
way by the application of an external stimulus. For instance, a "smart" adsorbent,
capable of existing in dense and open states might not recognize certain adsorbates and
remain in a dense form [27]. On the other hand, for adsorbates that are recognized, the
material switches to the open form, and adsorbs the material rnpidlyand efficiently. One
material recently recognized as being capable of such behaviour is the metal coordination
polymer [CuL2] where L is the is the p-diketonate ligand CF3COCHCOC(OCH3)(CH3h
This material is a robust zeolite mimic that exhibits a dense form as well as an open
form with channels capable of sorbing many small molecules. In addition, certain guests
also are able to transform the dense form to the open form upon contact, a process that
depends on switching the conformation of a ligand L in half of the CU[L2] species. (Fig.
6, left). Thus the transformation from closed to open form involves a molecular-level
switch, a feature that likely contributes to the robustness of the phases and the
responsiveness of the transformation to specific stimuli. The entire process of activation
and deactivation is illustrated in fig. 6, right When the dense form is present, only the
gas phase line at Oppm is seen. Contact with methylene chloride vapoUr transforms the
dense to the open form. This only becomes visible once the flowing gas containing the
HP Xe strips out the methylene chloride, as Xe cannot compete with methylene chloride
as sorbent. The increasing amount of void space is evident from both the increase in
signal intensity as well as the change in chemical shift. Upon application of a heat pulse,
the material transforms back to the dense form, thus expelling all HP Xe and leaving the
gas phase line.

4. Conclusions

It is clear from the applications presented in this paper that working with hyperpolarized
xenon has much to offer those who are interested in characterizing the void space in
solids or its time dependence as the result of a process. The increase in sensitivity allows
many experiments that otherwise would be far too time consuming, and it also gives Xe
NMR spectroscopy the unique capability of following non-equilibrium processes over
times scales from fractions of a second to hours.

5. References

1. Raftery, D, B. F. Chmelka, (1994) NMR Basic Princ. Progr., 30, III


2. Ratcliffe, C. I. (1998) Ann. Rep. Nmr Spectr. 36, 123
3. Bonardet, 1.-L., 1. Fraissard, A. Gedeon, M. A., Springuel-Huet, (1999) Catal. Rev.-
Sci. Eng. , 41, 115.
4. Ripmeester,1. A. (1982).1. Am. Chem. Soc. 104,289
5. Ripmeester,1. A., C. I. Ratcliffe, 1. S. Tse, (1988) . .1. Chem. Soc. Faraday Trans. I, 84, 373l.
6. Ripmeester,1. A., C. I. Ratcliffe, (1990).1. Phys. Chem. 94, 7652
7. Ripmeester,1. A., C. I. Ratcliffe, (1993) Anal. Chim. Act. 283, 1103
8. Breeze, S. R., S. 1. Lang, A. V. Nossov, A. Sanchez, I. L. Moudrakovski, C. I. Ratcliffe and 1.
A. Ripmeester, (2000) Nanoporous Materials IT, Elsevier, Amsterdam, 491
122
9. Moudrakovski, I. L., C. I. Ratcliffe, J. A Ripmeester, (1995) .In Zeolites: A Refined Tool for
Designing Catalytic Sites; BOlUleviot, L., Kaliaguine, S., Eds.; Elsevier Science: New York,
10. Chmelka, B. F.D. Raftery, A V. McCormick, L. C. De Menorval, R. D., Levine, A Pines, A
(1991) Phys. Rev. Lett. 66,580
11. D. Raftery, L. Reven, H. Long, A Pines, P. Tang, J. A Reimer, (1993).1. Phys. Chem. 97,
1649
12. MacNamara, E. C. V., Rice, J. Smith, L. J. Smith, D. Raftery, (2000) Chem. Phys. Lett. 317,
165;
13. Pietrass, T., A Bifone, A Pines, (1995) Surf Sci. 334, L730;
14. Seydoux, R, A Pines, M Haake, J. A Reimer, (1999).1. Phys. Chem. B 103,4629.
15. Grover, B.C. Phys. Rev. Lett. 1978,40,39
16. Happer, W., E. Miron, S. Schaefer, D. Schreiber, W. A van Wijngaarden, X Zeng, (1984)
Phys. Rev. A 29, 3092
17. Driehuys, B.; G. D. Cates, E. Miron, K. Sauer, D. K. Walter, W. Happer, (1996) Appl. Phys.
Lett. 69, 1668.
18. Hunger, M, T. Horvath, (1995).1. Chem. Soc. Chem. Commun. 1423
19. Jeffrey, G. A (1996) in Comprehensive Supramolecular Chemistry, ed. J. L. Atwood, J. E.
D. Davies, D. D. MacNicol and F. Vogtle, PergamonlElsevier, Oxford, Vol. 6,757
20. Davidson, D.W.; Y. P. Handa, 1. A Ripmeester, (1986).1. Phys. Chem. 90,6549
21. PietIass, T., H. C. Gaede, A Bifone, A Pines, A, J. A Ripmeester, (1995).1. Am. Chem.
Soc. 117,7520
22. Moudrakovski, I. L, A A Sanchez, C. I. Ratcliffe and J. A Ripmeester, (200 1).1. Phys.
Chem. B (submitted)
23. Moudrakovski, I. L., A A Sanchez, C. I. Ratcliffe, 1. A Ripmeester, (2000).1. Phys. Chem. B
104, 7306
24. Moudrakovski, I. L., A Nossov, S. Lang., S. R Breeze, C. I. Ratcliffe, B. Simard, G. Santyr,
J. A Ripmeester (2000) Chem. Mater. 12, 1181
25. Gregory, D. M, R E. Gerald II, R E. Botto, (1998).1. Magn. Reson. 131,327
26. Moudrakovski, I. L., S. Lang, C. I. Ratcliffe, B. Simard, G. Santyr and J. A Ripmeester,
(2000).1. Magn. Reson. 144, 372
27. Soldatov, D. V. J. A Ripmeester, S. I. Shergina, I. E. Sokolov, A S. Zanina. S. A Gromilov,
Yu. A Dyadin, (1999).1. Am. Chem. Soc. 121,4179
28. Nossov, A v., D. V. Soldatov, J. A Ripmeester, (2001).1. Am. Chem. Soc. 123,3563
NMR IN COLLOID SCIENCE WITH SPECIAL EMPHASIS ON SELF-
AGGREGATING SYSTEMS

OLLE SODERMAN, CARIN MELANDER, MAGNUS NYDEN AND


DANIEL TOPGAARD
Physical Chemistry 1
Chemical Center, Lund University, P.o.Box 124, S-221 00 Lund, Sweden
and (Magnus Nyden)
Applied Surface Chemistry
Chalmers University of Technology, SE-412 96 GOteborg, Sweden
e-mail: 01le.soderman@fkem1 .lu.se

1. Introduction

Fluids used in many processes of industrial or biological relevance often have a complex
structure on the colloidal length scale. They may contain polymers, which may interact to
form gels, self-assembled surfactant aggregates or two separate fluid phases such as in an
emulsion [I] . The structure on the colloidal length scale is often decisive in determining
how a certain fluid performs in a specific application. As a consequence, much effort has
been devoted to developing methods whereby structural studies of such complex fluids can
be investigated.
In the field of colloidal science, such fluids go under a number of different names; e.g.
complex fluids, soft-matter, and microstructured fluids. NMR has been, together with
different scattering methods, perhaps the most useful tool in structural studies of such
complex fluids [2]. In particular, two methods have been used: NMR relaxometry and
diffusometry. While the former method relies on rather complicated modeling to extract the
relevant information from the measured relaxation times, the latter method conveys easily
interpretable information. It should be noted that both relaxometry and diffusometry
measure dynamical quantities (relaxation rates and diffusion coefficients). As a
123
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 123-138.
© 2002 Kluwer Academic Publishers.
124

consequence, the extraction of structural information from the observable quantities relies
on modeling. In fact, there is often an intricate coupling between structure and dynamics in
complex fluids, the details of which are often decisive in technical applications.
Dissolution of solid polymers and drying of paint are two examples. This contribution will
deal with the application of NMR diffusometry, and we will attempt to show how this
method can be used in studies of complex fluids.
With NMR diffusometry, one can determine the root mean squared (rms) displacement
of the various components making up the system. The time over which the molecules are
allowed to diffuse can be varied from a ms to several seconds (although with modulated
gradients the time can be shortened to a few 100 IlS [3]). The relevant structural information
is then obtained from simple considerations about the distances the molecules diffuse
during the diffusion time and how the value of this distance is compatible with the possible
structures of the system under study. It should be remarked that molecular transport over
colloidal dimensions (such as in a biological cell) occurs through self-diffusion, a fact
which constitutes further motivation to perform NMR diffusion studies.
The NMR method for determining diffusion and flow has a number of advantages
over other methods: it is component resolved, it does not require isotopic labeling, it is
capable of determining self-diffusion coefficients (henceforth referred to as SDC) over a
wide range (from those for simple liquids such as water to very slow diffusion as is found
for large polymers in solution and gels), it is sensitive to hindrance for the diffusion and,
finally, it is a rapid method [4, 5].
Within the field of colloidal science, NMR diffusometry has been useful in the
characterization of surfactant systems, in particular those containing waterl"oil"/surfactants
[2]. Other areas pertain to porous systems such as foodstuffs, emulsions and rocks [6].
Finally, polymers and gels constitute another area where the method has been successfully
applied [7].
Here we will report examples from our work on surfactant and polymer systems.

2. Experimental and theoretical details

2.1 EXPERIMENTAL DETAILS

There are numerous accounts concerned with the technical aspects of NMR diffusometry
[2, 4, 5], and we shall not discuss technical aspects here, save for the problems connected
with temperature gradients, since in our experience, this is often a serious problem in
125

studies at temperatures above room-temperature, in particular for studies of low viscosity


liquids. Most NMR probes are temperature-regulated by passing a stream of temperate air
of specified temperature along the sample. The air stream usually encounters the bottom of
the tube first, and if experiments are performed above room temperature, a negative
temperature gradient along the NMR sample tube may develop, the presence of which may
give rise to convective flow. Such flow will contribute to <Z2> and not <Z> (since there
will be a counter-flow such that there is no net-flow through any cross-sectional area in the
sample) and therefore one measures a value for the SDC which is too high. The standard
way to test for the presence of convective flow is to measure the self-diffusion as a function
of the diffusion time. For simple fluids, the self-diffusion coefficient should not depend on
the diffusion time in the absence of convection.
We note that pulse sequences have been developed which compensate for convective
flow [8]. An example is given in Figure 1, where we display data for surfactant diffusion in
a binary water/surfactant solution. Clearly, the influence of convection can be quite large
but the effect can also be minimized by the use of appropriate pulse sequences.

14
12
,. til
10
N

E 8
0

0 6
Q 4
2
0
290 300 310 320 330 340
T/K
Figure 1. Experimentally obtained diffusion coefficients for a surfactant of
concentration 1 wt% in water. Results presented are from a Hahn-echo based diffusion
experiment, using two different diffusion times (the two upper lines). The observation
of a time dependent SDC is the sign of the presence of convective flow in the sample.
Also shown are the results of a pulse sequence, designed to compensate for convection
(lower line). With the latter sequence there is no dependence of the SDC on the
diffusion time.
126

2.2 THEORETICAL DETAILS

We will not discuss in detail the theoretical background ofNMR diffusometry, but refer the
reader to a number of books and reviews [2, 4-6, 9]. Here we will limit ourselves to a
definition of the basic relations pertaining to NMR diffusometry.
The experiment is based on spin-echoes, into which pulsed field gradients are inserted.
In the simplest case, a Hahn-echo, with a gradient pulse inserted on either side of the 1t-
pulse, is used. For the case of Gaussian diffusion, the normalized amplitude of the spin echo
is then given by:

where ~ is the distance, from leading edge to leading edge, of the gradient pulses with
length 8 and amplitude G. D is the SDC of the investigated component. A semi-logarithmic
plot of the echo-intensity vs. the parameter k yields a straight line, from which the value of
the SDC may be obtained. Such a plot is often referred to as a Stejskal-Tanner plot.
For the case of non-Gaussian diffusion, it is customary to introduce the variable q,
defined as q = yG8 /(21t). Under some rather general conditions, the mean-squared
displacement can be obtained from the initial slope of the echo-decay plotted vs. q2,
irrespective of whether the diffusion process is Gaussian or not [6).

3. The microstructure of surfactant solutions

One of the main applications of surfactants is to make it possible to mix oil (with oil we
mean non-polar substance such as "simple" hydrocarbons) and water into a stable one-
phase solution. Such solutions are usually termed microemulsions, and it is the ability of the
surfactant to form a film, which locally separates water from oil, which makes it possible to
form microemulsions. Such films can have many different structures and it is in the
determination of these structures that NMR diffusometry has been very useful [10). It has
turned out to be fruitful to classify the structures according to whether the film encloses a
certain volume as in discrete particles (normal and reversed micelles or vesicles are
examples) or whether the structure is connected.
127

NMR diffusometry allows one to chose between these possibilities in a simple and
straightforward fashion. Consider the case of a normal oil-swollen micelle. Here, the
surfactant forms a film that encloses the oil-component and the resulting aggregate is
dispersed in a continuous water domain. As a consequence, the oil- and surfactant diffusion
coefficients are equal and given"by the hydrodynamic radius of the micelle (the value of
which can be obtained from the value of the oil or surfactant SDC). The water SDC is
slightly reduced from its bulk value, on account of obstruction effects exerted by the
micelles and "binding" of water to the surfactant film. In the opposite case of a reversed
micelle, the hydrocarbon forms the continuous domain and its diffusion is close to that of
the pure hydrocarbon component while the surfactant and water diffusion is equal and again
given by the hydrodynamic radius of the micelle. For bicontinuous structures (defined as
structures where there are sample spanning diffusion paths for all components) the situation
is different. The water and oil component diffusions are now essentially occurring in a two-
dimensional situation (the surfactant film blocks one dimension) and thus their diffusion
coefficients are reduced by about 30 to 50 % from the value in bulk (everything else being
equal). The surfactant diffusion is now lateral diffusion along a surfactant film.
This brings us to our first example of the application of NMR diffusometry to
complex fluids. Consider the phase diagram in Figure 2.
128

Oil

I,
0

V
0

Hp 50 DDAS

Figure 2. Isothermal ternary phase diagram for the DDASlhexadecane/water system.


The region of interest in the context of the present discussion is the microemulsion
region marked L, that extends out from the water comer. The three dotted lines indicate
investigated samples (see text for details).

The surfactant is the cationic didodecyltrimethylammonium with sulfate as counter-ion


(DDAS), while the oil-component is hexadecane. The region marked L\ shows the
extension of the microemulsion region. Thus in the entire L\ region, one-phase solutions are
found containing various amounts of a hydrocarbon and water. To unravel the local
organization of the solutions, we present in Figure 3 values for the diffusion coefficients of
the surfactant and oil-components on lines connecting the upper boundary of the
microemulsion with the lower boundary. The two connected points correspond to the
maximum and minimum oiVwater ratio that a given water to surfactant ratio may solubilize.
129

a 8

t!I

D
o~ 2

t!I
B

wt % 10wer line
3.5 r-------.---T---.----.-----,
a o0
I!I

;~ 2.5 o

.
o
g 2
q D
0"1.5
o
o
o
o
0.5 L....~.o_'_~~-'-~~-'-'~~.lL c....,
" _.......
o w w w w ~

wt % lower line
4.5 . - - _ - - - - - . _ - - . . - - - . - - - - . - - - - , .

"~ 3.5
~'e 3

g 2.5
q
ct 2
a
1.5 t!I

Figure 3. Self-diffusion coefficients of oil (0) and surfactant (0) for samples on three
different lines connecting the upper and lower boundary of existence of the
microemulsion (see the phase diagram in Figure 2). The samples have been made by
mixing appropriate amounts from a sample from the lower line with a sample from the
upper line. The x-axis is expressed in terms of wt % of the lower sample in the
investigated sample.

The three lines correspond to, from top to bottom, increasing surfactant concentration. As
can be seen, for the case of the highest surfactant dilution (top panel in Figure 3), the oil
130

and surfactant SDC are equal for all samples. This signals the presence of normal oil-
swollen micelles. The decrease in the value of the SDC that occurs as the oil-to-water ratio
decreases indicates that the micelles grow in size. In fact, in the region close to the upper
line, with almost constant oil and surfactant diffusion, the aggregates are spherical. The
point where the SDC starts to decrease indicates that a structural transition of the micelles
occurs.
At intermediate surfactant concentration (middle panel in Figure 3), the surfactant and
oil diffusion SDC is first equal and then starts to differ as one approaches the lower
boundary of microemulsion existence. The point where the oil and surfactant SDC starts to
diverge indicates a structural transition into a bicontinuous structure. The mechanism
whereby the oil and water diffuses is no longer solely the diffusion of discrete surfactant
aggregates.
Finally, at the highest surfactant concentration (bottom panel in Figure 3), the oil and
surfactant SDC start to differ roughly halfway into the region. Again, this indicates that a
structural transition takes place, where the discrete aggregates are transformed into a
bicontinuous structure. While the values of the SDC do not indicate the detailed appearance
of the surfactant film in the bicontinuous solutions, they do certainly indicate where the
transition takes place. In the description of bicontinuous microemulsions, models based on
interconnected rods or minimal surfaces are often used. We end this section by showing in
Figure 4 a diagram in which the structures found in the LI phase of the DDAS/oil Iwater
system as based on detailed NMR diffusometry is outlined. The bicontinuous structure is
indicated as one based on interconnected surfactant rods [11]. In fact, the picture in Figure
4 represents a structure with long-range ordering. In the microemulsion solution, there can
be no long-range order, so the structure in this case corresponds to a "melted" version of the
one in Figure 4, where the long-range order is lost, but locally the structure resembles the
one in the Figure.
131

Figure 4. A schematic diagram showing the evolution of the microstructure in the LI


region of the DDASlhexadecane/water system. The y-axis is expressed in terms of the
ratio of the oil and surfactant volume fractions, while the x-axis is given by the volume
fraction of surfactant plus oil in the solutions. Also given is a representation of the
interconnected rod structure, in which oil-swollen surfactant rods are interconnected and
form a three-dimensional network.

4. Concentrated emulsions - a "soft" porous system

Emulsions consist of two separate fluid phases where one is in the form of droplets of
colloidal dimensions dispersed throughout the other fluid. Since the two fluids are not
miscible, an emulsion is unstable and will eventually separate. In order to stabilize the
droplets, an emulsifier is added. Often these are surfactants, but could also be polymers.
132
One important characteristic of emulsions is the volume fraction of dispersed phase. In
concentrated emulsions (sometimes called gel-emulsions or high internal phase ratio
emulsions) this fraction can be as large as 0.99. At these high volume fractions of dispersed
phase, the droplets are deformed, and the system can be described as domains of a liquid
separated by a thin film of another liquid. In the case of a w/o concentrated emulsion, the
domains consist of water separated by a thin film of hydrocarbon. Concentrated emulsions
are used in a number of technical applications, for instance in cosmetic and pharmaceutical
formulations. In such applications, it is often important to be able to characterize the
diffusion of the various components of the emulsion.
Consider the diffusion of water in such systems. It is influenced by the presence of the
barrier the oil-film exerts. For relatively short diffusion times, a water molecule does not
reach the droplet boundaries and therefore the observed water SDC is close to that of pure
water. For long diffusion times, the diffusion will be given by a random walk between the
water domains, and thus the lifetime of a water molecule in the water domains and the
dimensions of the droplets give the diffusion coefficient. At intermediate diffusion times,
such that the water molecules only jump between a few numbers of droplets, one may
observe "diffusion diffractograms" [12]. An example is given in Figure 5.
109

108 •'1
I
Co i
'~107 i~

\.
B
c

106 II ~

105
0 2 lOs 4 lOs 6 lOs 8 lOs
·1
q/m
Figure 5. "Diffusion diffractogram" for water in a concentrated W/O emulsion. The
position of the peak is related to the droplet size. The emulsion has been investigated
over a period of 24 hours (the different symbols correspond to data taken at regular
intervals during this time). Clearly, the emulsion undergoes no change during the
investigated period.
133

The origin of the peak can be understood as follows. The gradient pulses induce a
phase shift along the sample with a certain wavelength, the value of which depends on the
gradient strength. For the case where the molecules jump well-defined distances, diffraction
like peaks will occur at multiples of wavelengths along the samples that match the jumping
distances. Thus the value of the peak gives direct information about the droplet size in the
case of concentrated emulsions. Moreover, as the method is non-evasive, one may study the
long-time stability and the evolution of the droplet size in concentrated emulsions.
As mentioned above, one application of concentrated emulsions is in the field of
pharmaceutical formulations. Often an active substance is included, and it is of importa)1ce
to be able to follow the diffusion of the active component. As an example of how this can
be done, we present measurements, in a double-logarithmic plot, of the mean squared
displacement of the acetate ion vs. diffusion time for acetic acid/sodium acetate mixtures in
Figure 6.

10-9

HO
2
10- 10
5195
M

E
JI
N
V 10- 11

10- 12
10 100 1000
ill ms

Figure 6_ Mean squared displacements of acetate in a concentrated W/O emulsion,


plotted VS. the diffusion time. The water solutions used to prepare the emulsions are
mixtures of acetic acidlsodium acetate with molar ratios HAcINaAc as indicated in
the Figure (the total concentration is 0.25 M). Also included is the mean squared
displacement of water in the emulsions_

The fact that the data for falls on straight lines indicates that <Z2> scales linearly with the
diffusion time and thus the intercept with the y-axis yields the (long-time) SDC of the
components. For water (also included in Figure 6) the long-time SDC is of the order 2xlO· 10
m2 S-I, and thus reduced from the bulk value by approximately one order of magnitude. For
134

the acetate ion, the system containing 95 % undissociated acid has the highest value of the
long-time SDC, while the system with 5 % undissociated acid has the lowest value of the
SDC. This means that the resistance of the film for diffusion of our model active substance
governs the long-time diffusion. Recall that the film is composed of a hydrocarbon, in
which the ionic acetate ion has very low solubility, whereas the uncharged acetic acid
molecule has a higher solubility. The flow of any component across the film is given by the
product of the SDC in the oil and the solubility in the oil of the component. This thus
explains the observed behavior. In conclusion, NMR diffusometry is capable of providing
information about the emulsion structure (droplet size), long-time emulsion stability and
transport of active components solubilized in the emulsions.

5. Diffusion in polymer solutions and gels

As a final example in this expose over applications ofNMR diffusometry, we shall describe
its application to the problem of solution and gel-structure in ethyl(hydroxyethyl)cellulose
(EHEC)/water systems. EHEC is a cellulose derivative that is widely used as thickener in
paints and related products. EHEC solutions display a complex phase behavior with
separation into two solutions, one rich and one poor in EHEC content, upon increase in
temperature.
Thus water becomes a poorer solvent for EHEC as the temperature increases and one
natural question is how this effect manifests itself in terms of the solution structure of
EHEC solutions. To shed some light on this question, we have studied the diffusion of
probe molecules of varying size through the EHEC-matrix, both in solutions and in
chemically cross-linked gels ofEHEC.
As probe molecules, we have chosen monodisperse polyethyleneoxide PEO polymers
of varying molecular weights. PEO was chosen since it would appear that it does not
interact with EHEC, which is a necessary condition as specific interactions between the
EHEC and the probe molecules would complicate the interpretation.
The concentration of the EHEC was chosen to be 1 wt%. With a molecular weight of
200 kD, this corresponds to an average distance between two polymers of approximately
300 A, which can be taken as the size of the mesh of the polymer matrix provided that the
polymer is evenly distributed in the solution. The concentration of the probe molecule was
0.01 wt% and their molecular weights ranged from 10 kD to 963 kD. The hydrodynamic
radius of the 10 kD PEO is roughly 30 A, thus much smaller than the mesh size of the
EHEC matrix given above.
135

Echo-decays for 4 different PEO molecules of increasing molecular weights in EHEC


solutions are presented in Figure 7. Two features are immediately apparent. First, the echo-
intensities fall on a straight line in the Stejskal-Tanner plots. Secondly, the SDC of the
probe-molecules is considerably reduced as compared to the case of probe molecules in
water in the absence of EHEC (compare the slopes of lines in Figure 7). The last statement
is true also for the probe-molecule with the lowest molecular weight.
In addition, the SDC obtained is independent of the diffusion time in the investigated
range from 100 ms to 600 rns.

10kD
SOkD
..,§
Ii
~
c
.c
~

4 lOW 6 10w 8 IOID 1.2 lO" 1.6 1011


q'/m" q' I nr'
c)
d)
288 kD 963 kD

§
."g
~ 0.1
,g
Jl

q'~ ~~
Figure 7. Echo intensity plots for PEO molecules of varying molecular weights in a I
wt% EHEC solution. Also included are the results for the PEO molecules in pure
water (thinner lines).

What information about the EHEC solution can we derive from these observations?
The obvious explanation for the much reduced value of the SDC is that the EHEC matrix is
inhomogeneous. In other words, there are denser and less dense regions of EHEC. Across
the dense regions, the flux of probe molecules is low on account of the fact that the
136
diffusion of the probe molecules is low in the dense regions and the concentration of probe
molecules in these regions is reduced. As a consequence, the longtime probe SDC in the
polymer solution is reduced. However, the probe molecules sample both the dense and less
dense regions on the experimental time scale, since the decay is linear in the representation
of Figure 7. There are two conceivable mechanisms for this averaging process. In terms of
solution structure, the correlation length of the spatial concentration variation of EHEC is
less than a few !lm. This number can be derived from the mean squared displacement of the
probe molecules.
Alternatively, in terms of EHEC solution dynamics, the fluctuation in EHEC
concentration may average out the EHEC matrix inhomogeneities on the relevant length-
scale.
Let us now consider what happens if we chemically cross-link the EHEC solution.
This can be done by addition of suitable cross-linking agents. Apart from the cross-linking,
all other conditions (EHEC and probe concentration, temperature etc.) remain the same as
in the EHEC solution experiments.
The results are presented in Figure 8. Now the decays an, no longer straight in the
Stejskal-Tanner plots. This indicates that the length-scale of the inhomogeneities have
increased such that there is slow exchange of the probe molecules between dense and less
dense regions of EHEC.
137

b)
SOleD

0. 1
"

t Jd ' 0.01 OL.........L.I~IO~"~2~10~"~3-1~0"~-4~IO'-'.............


5 Id1
<i 1m"
oj

.......
d)
288 leD
963 leD
,
,
.

q 2 Im'~ q~ Im'~

Figure 8. Echo intensity plots for PEO molecules of varying molecular weights in a I
wt% EHEC chemically cross-linked gel. Also included are the results for the PEO
molecules in pure water.

In terms of the mechanisms for averaging the EHEC matrix inhomogeneities


discussed above; the dynamics of the matrix are considerable slower in the chemically
cross-linked gel than in the EHEC solution. Thus in the EHEC gel the spatial length scale
of the matrix inhomogeneities must be larger than a few f..lm.

6. Concluding remarks

The purpose of this paper has been to show how NMR diffusometry can be used to convey
information about structure and dynamics of complex fluids. The information can be
obtained without the need to invoke complicated models. In fact, simple reasoning
pertaining to how a molecule may diffuse through the studied systems allows one to chose
138

between different conceivable solution structures and to obtain non-trivial information


about solution structure.

7. References

I. Evans, D.F. and H. WennerstrOm, The colloidal domain where physics, chemistry, and biology meet. 1999,
New York: Wiley-VCH.
2. SOderman, O. and P. Stilbs. (\ 994) NMR studies of complex surfactant solutions, Prog. NMR Spectroscopy
26,445-482
3. Callaghan, P.T. and 1. Stepisnik. (\995) Frequency-domain analysis of spin motion using modulated-
gradient NMR, 1. Magn. Reson. A 117, 118-122.
4. Stilbs, P. (1987) Fourier transform pulsed-gradient spin-echo studies of molecular diffusion, Prog. NMR
Spectroscopy 19,1-45.
5. Callaghan, P.T., Principles of nuclear magnetic resonance microscopy. 1991, Oxford: Clarendon Press.
6. Callaghan, P.T. and A. Coy, PGSE NMR and molecular translational motion in porous media, in NMR
Probes and Molecular Dynamics, P. Tycko, Editor. 1994, Kluwer: Dordrecht. p. 489-523.
7. Nyden, M., NMR diffusion studies of microheterogeneous systems. Surfactant systems, polymers solutions
alld gels. PhD thesis, 1998. Lund University: Lund.
8. Jerschow, A. and N. MUller. (\997) Suppression of convection artifacts in stimulated-echo diffusion
experiments. Double-stimulated-echo experiments, 1. Magn. Reson. 125,372-375.
9. Kimmich, R., NMR: tomography, diffusometry, relaxometry. 1997, Berlin: Springer Verlag.
10. Olsson, U. and H. WennerstrOm. (1994) Globular and bicontinuous phases of nonionic surfactant films,
Adv. Colloid Interface Sci. 49, 1\3-146.
I \. Mariani, P., V. Luzzati, and H. Delacroix. (\ 988) Cubic phases of lipid-containing systems. Structure and
biological implications, 1. Mol. Bioi. 204, 165-189.
12. Hdkansson, B., O. SOderman, and R. Pons. (1998) Structure determination of a highly concentrated WIO
emulsion using PFG-SE NMR diffusion-diffractograms, Langmuir 15, 988-991.
SOLID STATE NMR CHARACTERIZATION OF POLYMERS AND
SURFACTANT MOLECULES AS CONFINED TO POROUS SILICA MATERIALS

ROBERTO SIMONUTTI, ANGIOLINA COMOTTI, SILVIA BRACCO,


PIERO SOZZANI
Department ofMaterials Science, University of Milan-Bicocca,
via R. Cozzi 53, 1-20125 Milan, Italy. E-mail: Piero.Sozzani@unimib.it

1. Introduction

The present chapter is aimed at presenting the characterization of nanocomposites formed


by mesoporous silica and surfactants or polymers. The hybrid materials were prepared in
the former case by condensation of silica about micellar aggregations of surfactants [1,2];
in the latter case novel polymer nanocomposites were formed by direct polymerization or
by absorption of polymers in the pores of the mesostructured silica. The materials provide
intimate interactions between the nanophases, and thus a challenging field for NMR
characterization of the extended interfaces. The steric limits imposed to the chains by the
rigid siliceous structure permit the controlled reduction of the degrees of conformational
and translational freedom as compared to the bulk phases. Concerning polymers, unusual
states of aggregation can be observed massively in these nanocomposites and basic
phenomena, like glass transition or crystallization, are to be re-explored in the confmed
state. Preliminary NMR characterization of nanophase properties like confined
crystallization of alkyl chains, or monomer reactivity in the inclusion state are described
hereafter.
Solid State NMR has been widely applied for the characterization of silica and silica
composites. \3C MAS NMR provides carbon chemical shift (CS) and carbon relaxation
times that can describe the conformation and motional behavior of the confined polymer or
surfactant [3]. In 29Si MAS NMR spectra the intensity ratio between the peaks and their
line width can describe accurately the silica particles. IH_ 29Si Cross Polarization (CP)
dynamics are a valuable tool for studying the proximity of protonated systems, like
polymers, to silica surfaces. We have shown this effect for a polyisoprene/silica system [4]
where the characterization of the cross polarization dynamics allowed us to demonstrate the
proximity between the polymer and the silica surface at the molecular level
Polymer/silica interfaces in standard composites are very diluted and poorly defmed.
MCM-41 silica systems with extended and structured interphases, therefore, can provide a
detailed model for polymer-silica composites and for the organization of surfactants on
silica surface [3].
In the following the description of the organization at low temperature and reactivity of
cetyltrimethylammonium inside the nanopores is reported. Furthermore, once removed the
templating agent, the mesopores are accessible to organic molecules. Some monomers
(methyl methacrylate, caprolactame) have been absorbed in MCM-41 and polymerized
inside the pores, also poly(ethylene oxide) has been included in the nanochannels:
139
1. Fraissard and O. Lapina (eds.J, Magnetic Resonance in Colloid and Interface Science, 139-153
© 2002 Kluwer Academic Publishers.
140

confonnations and mobility of the confined polymers and silica-polymer interactions will
be discussed.

2. Materials

MCM-41 samples with variable amount of aluminum were prepared following the
procedure reported by Stucky et al [5]. The molar composition of the reaction mixture
consists of 1 NaAI0 2, 5.3cetyltrimethylammonium chloride (C I6TMACI), x tetraethyl
orthosilicate (TEOS) and 1450 H20. The x orthosilicate amount was varied giving rise to
four samples with SiiAI ratios of 16, 12, 10, and 8. A sample with no sodium aluminate was
also prepared. The reactions were carried out under refluxing for 5 h. The products were
washed with deionized water to remove any surfactant excess. All the samples contain
more than 45% of the templating agent, as measured using thennogravimetric analysis. The
samples were subjected to following thennal treatments: a) at 150°C for 1.5 h under
nitrogen atmosphere; b) by calcination in a muffle furnace at 630°C for 5 h in order to
remove the template. In the latter case all the samples present a surface area higher than 800
m2/g as measured from nitrogen adsorption.
The MCM-411poly(methyl methacrilate) (PMMA) nanocomposite (with SiiAI ratio
equal to 16) was obtained by free radical polymerization using the following procedure.
Initiator 2,2'-azobisisobutyronitrile (AIBN) was dissolved in distilled methyl methacrylate
(MMA), 0.005 AIBNIMMA molar ratio with moderate stirring under N2 atmosphere for 0.5
h. Then the thennally degassed MCM-41 powder was poured into the mixture and stirred
for another 0.5 h. Finally, the mixture was subjected to several freeze-to-thaw cycles and
kept at 4°C overnight. The MCM-411MMA powder was isolated by filtration. In order to
remove the MMA present outside the MCM-41 channels, the powder was kept under
vacuum for a few minutes. The MCM-41IPMMA nanocomposite was obtained by heating
the powder at 100°C for 48 h.

3. Surfactant Crystallization and Reactivity in Nanopores

I3C Single Pulse Experiment (SPE) MAS NMR spectrum of the as-synthesized mesoporous
silica MCM-41 with a SiiAI ratio of 10, is shown in Figure 1. The CS assignments are made
by comparison to the solution spectrum and are reported in the figure [6]. The surfactant
molecules within the mesopores of MCM-41 present, at room temperature, a high content
of gauche confonnations as determined by 13C chemical shift analysis. The carbQns along
the chain (from Cs to C 14) resonate at about 31 ppm, this value is close to that measured for
long-chain alkanes in the melt that contain a large fraction of gauche confonnations [7]. In
linear alkanes the largest confonnation effect on the chemical shift (CS) is called y-gauche
interaction, producing an upfield shift of about 5 ppm for a gauche arrangement [8].
Packing and susceptibility effects can also playa non negligible role. In the case of MCM-
41, the aliphatic chains interact mainly one another and the magnetic susceptibility of the
inner methylenes is similar to that of the bulk phase, thus a comparison between chains
confmed to mesoporous and in the bulk is valuable. In the spectrum we can also recognize
broad peaks associated with the carbons close to the polar groups (from C2 to C4) and the
methylene carbons directly bonded to nitrogens (C 1). All of them experience restricted
mobility because the ammonium group strongly interacts with the silica surface and the
aluminum sites. The broadening of the C 1 resonance is also due to the I3C_ 14N dipolar
coupling that is strongly affected by the 14N quadrupole interaction [9]. 13C TIS are short
141

and increase from the polar head towards the chain end, nevertheless proton to carbon cross
polarization is efficient for all carbons (not shown) demonstrating that the surfactants
confmed to the mesopores exhibit gel-like mobility with anisotropic motions progressively
reduced from the chain end towards the polar head [6].

1CH
I 3
3 5 7 9 11 13 15 17
1 CH -+N
3 I 2 4 6 8 10 12 14 16
1CH
3

C5- C 14

iii •••• 'i' iii I"" i Ii' ',"". Ii ii" 'i. Ii iii'.""" "

70 60 50 40 30 20
(ppm)
Figure 1. 75.5.MHz \3C SPE MAS spectrum recorded at room temperature of MCM-41,
with Sil Al ratio of 10, containing the surfactant molecules.

Magnetization transfer from the surfactant protons to the silicons on the pore surface is
found to be efficient, showing the proximity of the ammonium group of the surfactant to
the silica surface.
The mild thermal treatment at 150°C of MCM-41 sample with SilAI=10 causes
Hofmann elimination [10] of some surfactant chains with the formation of trimethylamine
and hexadecene. At room temperature the spectrum is similar to that of the untreated
sample except for two new signals at 139.3 and 114.8 ppm, due to the olefmic carbons
(Figure 2a). From the analysis of the spectrum, nearly 25% of C l6TMA chains transform
into I-hexadecene. As a consequence a fraction of the chains are detached from the surface
without any constraint. The Hofmann elimination has already been proposed [1,11,12] as
the main mechanism for template decomposition during calcination or thermogravimetry
but the I-hexadecene molecules have never been quantified by \3C solid state NMR after
142

controlled treatments and studied in situ. After the mild thermal treatment at 150°C, the
MCM-41 samples with decreasing content of aluminum, show the formation of decreasing
amounts of hexadecene: the MCM-41 sample with no aluminum shows, after the thermal
treatment, the formation of only 10% of hexadecene. This result suggests the active role
played by aluminum sites to the based-catalyzed Hofmann elimination.

4 6 8 10 12 14 16

3 5 7 9 11 13 15

a)

298K

b)

287K

c)

279K

d)

271K
e)

250K
i I •

140 120 100 80 60 40 20


(ppm)

Figure 2. 75.5 MHz 13C SPE MAS NMR spectra recorded at variable temperatures of
MCM-41 with SiiAI of 10 treated at 150°C for 1.5 h:
a) room temperature; b) 287 K; c) 279 K; d) 271 K and e) 250 K.
143

As a consequence of the Hofmann elimination a fraction of the chains do not present


anymore the polar heads and are not anchored to the surface. At room temperature the
mobile cetyl chains and hexadecene molecules confmed to mesopores present sharp
resonances and only near the ammonium group the mobility is restricted. At lower
temperatures, the spectra present a considerable broadening of the peaks and a dramatic
change in the region of the inner methylenes. Transition from gauche conformations to
nearly all-trans conformations along the main chain is detected by the appearance of a new
peak at 33.8 ppm for the midchain carbons (CS-C IS , Figure 2 c-e). At 279 K the coexistence
of both conformational states is observed by the presence of two peaks; at lower
temperature the upfield peak decreases and disappears at 250 K. At 250 K only the
downfield peak at 33.8 ppm survives.

a)

287K

b)
279K

c)
271K

d)
250K
60 50 40 30 20
(ppm)

Figure 3. 75.5 MHz l3e CP MAS NMR spectra recorded at variable temperatures of
MCM-41 with SiiAI ratio of 10 treated at 150°C for 1.5 h;
a) 287 K; b) 279 K; c) 271 K and d) 250 K.
144

The l3C chemical shift dependence of inner methylene carbons on conformations has
already been described in the literature for both polyethylene and n-alkanes [8].
Polyethylene in the orthorhombic crystalline phase shows a signal at 32.8 ppm [13].
Crystallization in other forms gives rise to a further downfield shift, for example, the
monoclinic phase shows the signal at 34.3 ppm [14]. These values for the crystalline phases
are due to bonds arranged mainly in trans conformations along the main chain. The
chemical shift of alkanes stretched in nanochannels, where the chains assume all-trans
conformations, resonates at 33.0 ppm (15,16]. Even polymethylene sequences in ethr,lene-
rich copolymers show a chemical shift of about 34 ppm [14]. For comparison the I C CP
MAS NMR spectrum of the pure cetyltrimethylammonium chloride was also recorded (not
shown), the CS of the inner methylenes in the crystalline phase resonates at 33.9 ppm.
Therefore, we can assign the new signal at 34 ppm to the cetyl chains crystallized inside the
mesoporous inorganic material in trans conformations.

277K

b)
253K

c)
222K
'i""I"";' • i' , 'i'iliS' Ii. , " i •••

60 50 40 30 20
(ppm)
Figure 4. 75.5 MHz l3C CPMAS NMR spectra recorded at variable temperatures of
MCM-41 with no aluminum treated at 150°C for 1.5 h:
a) 277 K; b) 253 K and c) 222 K.
145

We have also perfonned variable-temperature CPMAS NMR experiments with short


contact times in order to enhance the rigid and crystalline phases (Figure 3). The transition
is also observed under CPMAS conditions but the peak at 34 ppm already prevails at 279K.
In the sample with no aluminum the \3C SPE MAS and CPMAS NMR experiments,
perfonned at variable temperatures, show the same trend as reported above for sample with
SiJAI ratio of 10 (Figure 4), but the transition is smoother and occurs at lower temperatures.
The analysis of mesoporous silica MCM-41 synthesized in acidic conditions (S+X'I+
mesostructure) shows that the thermal treatment at 150°C does not produce any hexadecene
(conftnning that the Hofmann reaction requires a basic catalysis); low temperature \3C
MAS NMR spectra of this sample are characterized by broad signals and the transition
from gauche to trans confonnations is detectable only at lower temperatures [3]. This
suggests the contribution of free hexadecene to the crystallization of the alkyl chains when
conftned to the mesopores. In addition, the removal of the trimethylamine during the
elimination reaction generates some space available for the cetyl chains to reorganize and
crystallize.

4. Inclusion Polyaddition: Poly(methyl methacrilate)

As reported by Aida et al. [17] it is possible to have large adsorption of AIBNIMMA


mixture in MCM-41 calcined systems: the authors reported also the characterisation of the
polymer obtained after heating at 100°C and extraction with tetrahydrofurane.
In Figure 5b the \3C Single Pulse Experiment MAS spectrum of MCM-41IMMA is
depicted (4 seconds of recycle delay): the peak at 168.4 ppm is due to carbonyl, the signals
at 137.7 and 126.0 ppm correspond to the oleftnic carbons, the methoxy group resonates at
52.3 ppm and the methyl group on the double bond shows a peak at 18.9 ppm. All the
signals are sharp indicating that inside the channels the main part of MMA behaves like a
liquid. After heating at 100°C for 48 h the MCM-41IMMA system, we obtain a MCM-
41IPMMA nanocomposite. In fact the spectrum changes drastically and cross-polarization
experiments can also be perfonned (Figure Sa). New peaks appear in the spectrum, at 178.0
ppm, 56.1 ppm, 52.4 ppm, 45.5 ppm and a very broad peak centred around 20 ppm. All
these signals correspond to the reported values for an atactic poly(methyl methacrylate)
[18]. Minor peaks due to residue monomer are present in the CP spectrum. In order to
evaluate the fraction of monomer still present in the nanocomposite, a quantitative SPE
MAS spectrum with recycle delay of 100s was perfonned. From the quantitative analysis of
the spectrum we calculated that nearly 70% of the monomer polymerized. Even if the
monomer still present is highly mobile, it is entrapped in the channels, in fact after a
treatment of 30 min in vacuum at 10-3 torr, no change of MMA concentration is evident.
MCM-41IPMMA nanocomposite is also characterized using Differential Scanning
Calorimetry. The calorimetric trace does not show any glass transition between 25°C and
250°C (glass transition of bulk PMMA is around 110°C).
The composite has been characterized also by 29Si Solid State NMR, in fact 29Si
chemical shifts are very sensitive to the structure of the silica p,articles. Silicon atoms
bearing one hydroxyl group (Si-O)3Si-OH (single silanols called Q ) show a chemical shift
of about -99 ppm, geminal silanols (Si-OhSi-(OHh (called Q2) resonate at -89 ppm,
instead a resonance at -109 ppm is due to silicon atoms without hydroxyl groups Si(Si-O)4
(called Q4) [19]. On the other hand it is possible to correlate the line width of the peaks with
the distortion of the siloxanic tetrahedra [20]. Furthennore the time constant (TsiH) of the
build-up of silicon magnetization in the CP experiment is proportional to the sixth power of
146

the proton to rare nucleus distance [21].

a)

b)

I i

200 160 120 80 40 (ppm)

Figure 5. a) 7S.5.MHz \3C CPMAS NMR spectrum of MCM-4IIPMMA nanocomposite;


b) 7S.S.MHz \3C SPE MAS NMR spectrum MCM-41IMMA sample.

In pure siliceus systems Q3 groups, having a proton two bonds apart, show short T SiH.
On the contrary the Q4 silicons have rather long TSiH because they are not in close contact
with protons and can receive magnetization only from the protons of single or geminal
silanol groups which are at least four bonds apart [22]. In the case of an extended interface
between silica and a protonated polymer, if the polymeric protons come in close contact
with the silica surface, Q4 siliconscan receive magnetization also from these protons and
show a shorter time constant (TSiH) of the build-up of silicon magnetization. In particular, a
quite short T SiH for Q4 silicons has been measured when a bidentate molecule,
bis(triethoxysilylpropyl)tetrasulfane at 3% concentration in a polyisoprene/silica
compound, was used. The bidentate molecule, reacting with the polymer by sulfur bridges
and with the silica surface by condensation, increases the closeness of the polymer chains to
the hydrophylic inorganic surface [4].
147

""-""
,,''''', "',,I' \
,
= _ _ _~-'= _____ .'::"""."""''''-''' __ _:"':.a _______:".:i,:;.;."........,::-...._

-80 -90 -100 -110 -120


(ppm)

Figure 6. 59.6 MHz 29Si CPMAS NMR spectrum of MCM-41IPMMA nanocomposite.


The spectrum was fully deconvoluted by Gaussian line shapes and the singles
dashed lines are reported for Q2 at -89 ppm, Q3 at -99 ppm and Q4 at -109 ppm.

In Figure 6 29Si CPMAS NMR spectrum of MCM-41IPMMA nanocomposite with 3.5


ms of contact time is reported. The signals due to Q2, Q3 and Q4 groups are easily
recognizable and the line widths are comparable with the published spectra of amorphous
silica [4].

~ : 18ms

: :
16ms

~ 14 ms
12 ms

: ~ :=
8ms

~ : 7ms
6ms

==
- -==
5ms

~:
~
:::: 4ms
3ms
2ms

~ 1 ms

~ 0.8ms

~ 0.6ms
~ 0.4 ms
, , ,
-60 -80 -100 -120 -140
(ppm)

Figure 7. 59.6 MHz 29Si CPMAS NMR spectra ofMCM-41IPMMA nanocomposite at the
reported contact times.
148

70
60
50
I (A.U.)
40
30
• °3
20
• °4
10

o 4 8 12 16 20
Contact times (ms)

Figure 8. Evolution of 29Si magnetisation in the variable contact-time cross-polarisation


experiment and non-linear least-squares fitting. Q3 filled dots; Q4 filled triangles.

Proton to silicon cross polarization dynamics were exploited using variable contact time
experiments on the MCM-41IPMMA nanocomposite, the spectra together with the fitted
intensities are reported in Figure 7 and 8. Short TSiH (3 ms) is found for Q4 species, as
reported for the MCM-41 sample containing the surfactant, and a rather long TllH) is
measured [3]. This highlights a strong interaction between the polymer chains and the
surface of the channels. The reduced mobility of the polymer chains causes the vanishing of
the glass transition.

5. Inclusion Polycondensation: Nylon-6

A class ·of polymerisation less studied in confined systems is polycondensation. In this type
of polymerisation small molecules, usually water, are obtained as by-products; in
nanochannels slightly larger than the monomeric unit like those made by urea,
perhydrotriphenylene (PHTP) and tris-o-phenylenedioxycyclophosphazene (TPP) the by-
products cannot escape and therefore the polymerisation is unfeasible. On the other hand,
MCM-41 channels having a larger pore dimensions (20-30 A) are well suited for
polycondensation. Once the monomers have reacted, there is enough space for the
rearrangement of the macromolecular chain and fmally for the removal of by-products.
In the field of polycondensation, we have studied the polymerisation of e:-caprolactame
(CPL). Usually CPL is polymerised starting from an aqueous solution of CPL (60-80%) in
the presence of small amount of e:-amminocapronic acid as catalyst. The reaction mixture is
brought to ca. 200°C under pressure, and then, keeping the temperature nearly constant, the
pressure is decreased to ambient pressure or vacuum in order to get high molecular weights.
The polymerisation of CPL confined to mesoporous has been carried out at ambient
pressure and no catalyst has been used because we have exploited the catalytic properties of
the silica surface. In this peculiar environment made by MCM-41 we do not need a solvent
as water in order to avoid the precipitation of the monomer and oligomers. The
polymerisation process was developed as follows: a) confinement of the monomer inside
the MCM-41 by absorption in the pores and b) polymerisation of the monomer at 180°C.
149

CPL is a crystalline solid with melting point around 70 °c. In Figure 9 the SPE spectrum of
CPL confined to MCM-41 channels is shown. The spectrum is pretty much featureless,
only two signals are recognisable: one at about 180 ppm (the lactamic carbon), and the
other broad signal from 30 to 50 ppm due to the overlapping of the methylenic carbon
atoms. This spectrum is quite informative, in fact the absence of significant spinning side
bands for the lactarnic carbon of CPL and an overall broadening of the methylenic carbons
account for high mobility of the caprolactame molecules inside the channels. The restricted
geometry produced by the mesoporous structure prevents the crystallisation of the
molecules but, since the cross-section is of about 20-30 A, a wide reorientation of the
caprolactame molecules is permitted. The low polarisation transfer in the CPMAS spectrum
also confirms the mobility of the caprolactame molecules.

I i i I i i i i i I i i i i i i i

240 200 160 120 80 40 (ppm)


Figure 9. 75.5.MHz DC SPE MAS NMR spectrum MCM-411CPL sample.

The polymerisation of CPL confined to the mesopores of MCM-41 has been carried out
heating the MCM-41/CPL sample at 180 °c for 16 hours under N2 atmosphere. A blank
Nylon-6 sample has been obtained by polymerisation of a mixture CPLIe-aminocapronic
acid (3% e-arninocapronic acid by weight) at 210 °C for 16 hours under N2 atmosphere.
Nylon-6 is a sernicrystalline polymer, with essentially three crystalline forms: form ex, p
and y. Forms ex and Phave the polymeric chains in an all-trans conformation, and hydrogen
bonds are formed between antiparallel chains [23]. The two forms differ only by the
shearing of the sheets along the c-axis: in the case of ex form the shearing is recuperative
leading to a monoclinic cell, instead in P form the c-shear is progressive and leads to a
triclinic cell. The y form shows the same basal plane of the monoclinic ex form, but
hydrogen bonds are formed between parallel chains. The amorphous phase is not well
characterised, but some authors suggest that the amorphous chains prefer an extended
conformation about the amide moiety [24]. In Figure lOa the DC CP MAS spectrum of the
blank Nylon-6 sample synthesised by us is shown and the chemical shifts are reported in
Table 1. The crystalline form of this nylon-6 is the ex form, consistent with the literature
[25]. In Figure lOb the 13C SPE MAS spectrum MCM-41INylon-6 composite is also shown,
150

the chemical shifts are reported in Table 1.

a)

• i' 'iii i ' 'i. 'i i • i •

240 200 160 120 80 40 (ppm)

Figure 10. a) 75.5.MHz 13C CP MAS NMR spectrum ofNylon-6;


b) 75.5.MHz I3C SPE MAS NMR spectrum MCM-41INylon-6 sample.

The line width at half height is considerable large with respect to that of the crystalline
phase and is of the same order of the amorphous phase. The calorimetric trace of the
nanocomposite shows no glass transition or melting point that are typical transitions of the
pure polymer. The polymer confined to the channels has therefore a quite peculiar thermal
behavior: the non-specificity of the local environments explored by the polymer chains
cause a variety of interactions together with a broad distribution of correlation times. The
distribution is not much affected by the change of temperature. The chemical shift of the
polyamide carbonyl inside the channels is close to that of Nylon-6 dissolved in
trifluoroethanol, indicating the presence of a strong hydrogen bonding between the amidic
group and the surface of the MCM-41 channels.
151

TABLE 1. I3C chemical shifts of Nylon 6 and MCM411Nylon-6 nanocomposite.

Nylon-6 Nylon-6 Nylon -6


.
Carbon Species

0 Crystalline Amorphous in MCM-41 Solution


II (ppm) (ppm) (ppm) (ppm)

r "NH~n
C 1 3 5

C=O 173.4 174.1 182.6,177.0 177.8

C1 43.4 40.3 40.3 40.7

C2 30.4 30.2 27.3 29.5

C3 30.4 27.4 27.3 27.1

C4 26.7 26.2 27.3 26.4

C5 36.6 36.9 37.2 37.2


* Diluted tnfluoroethanol solution.
6. Nanocomposite Based on Polyethyleneoxide

The last example of polymer included in the channels of MCM-41 is polyethyleneoxide


(PEO). The polymer is included by absorption from aqueous solution, being PEO soluble in
water.
In Figure lIb the 13C SPE MAS spectrum of pure PEO is presented. As reported by
Schantz [26], PEO spectrum shows a peak due to the overlapping of two resonances, one
narrow (71.8 ppm, 31 Hz line width) due to the amorphous phase and the other broader due
to the crystalline phase (71.7 ppm, 830 Hz line width). The line width of the crystalline
component is attributed to motional broadening. The I3C SPE MAS spectrum of MCM-
41IPEO nanocomposite sample is reported in Figure lla. The PEO confmed to the MCM-
41 channels shows only one peak with the line width between amorphous and crystalline
peak line widths. Also in this case the nanocomposite does not show any calorimetric
transition. The chemical shift of the included macromolecules is substantially coincident
with that of the bulk amorphous phase. This indicates that the confmement prevents the
formation of crystalline aggregates, on the other hand the amorphous phase behavior is
locally not much affected by the confinement. The cross section of MCM-41 channels is
large enough to contain more than one chain one next to the other and in the case of the
flexible polymer the mobility in the amorphous phase is not much restricted. This
phenomenon can be compared to the confinement in narrow nanochannels of 5 A diameter
constituted by tris(o-phenylenedioxy)cyclotriphosphazene that compel the PEO chains to
stay in the extended chain conformation [27].
152

In conclusion MCM-41 silica systems, like the MCM-41IPMMA, MCM-41INylon-6


and MCM-41IPEO nanocomposites, with extended and structured interfaces provide an
intriguing model for more common polymer-silica composites.

a)

b)

,," 4 . ,'"'' 0;: , .. , "I' 'h' iii "'" '" iii " .. I.",," 'i .. ' ...... I' """'I"""'"
100 80 60 40
(ppm)
Figure 11. a) 75.5.MHz 13C SPE MAS NMR spectrum MCM-41IPEO nanocomposite. b)
75.5.MHz DC SPE MAS NMR spectrum of PEO. The spectrum was fully
deconvoluted by Lorentzian line shapes and the singles lines are reported for
crystalline and amorphous phases.

7. Acknowledgement

This work has been partially supported by CNR-Agenzia 2000, MIUR and MSTA II
projects

8. References

1. Beck, J.S., Vartuli, lC., Roth, W.J., Leonowicz, M.E., Kresge, C.T., Schmidt, K.D.,
Chu, C.T.-W., Olson, D.H., Steppard, E.W., McCullen, S.B., Higgins, J.B. and Schlenker,
lL. (1992) J. Am. Chem. Soc. 114, 10834.
2. Corma, A. (1997) Chem. Rev. 97,2373.
3. Simonutti, R., Comotti, A., Bracco, S. and Sozzani, P. (2001) Chem. Mater. 13,771.
4. Simonutti, R., Comotti, A., Negroni, F. and Sozzani, P. (1999) Chem. Mater. 11, 822
5. Huo, Q., Margolese, DJ., Ciesla, U., Demuth, D.G., Feng, P., Gier, T.E., Sieger, P.,
Firouzi, A., Chrnelka, B.F., Schuth, F. and Stucky, G.D. (1994) Chem. Mater. 6, 1176.
6. Bovey, F.A. and Mirau, P.A. (1996) NMR o/Polymers, Academic Press, San Diego.
153
7. Moller, M., Cantow, H.J., Drotloff, H., Emeis, D., Lee, K.S. and Wegner, G. (1986)
Makromol. Chem. 187, 1237.
8. Tonelli, A. (1988) NMR Spectroscopy and Polymer Microstructure: The Conformational
Connection; VCH Publishers, Deerfield Beach, FL, 1988.
9. Naito, A., Ganapathy, S. and McDowell, C.A. (1981) J. Chem. Phys. 74,5393.
10. Hitz, S. and Prins, R. (1997)J. Catal.168, 194.
11. Keene, M.TJ., Gougeon, R.D.M., Denoyel, R., Harris, R.K., Rouquerol, J. and
Llewellyn, P.L. (1999) J. Mater. Chem. 9,2843.
12. Chen, C.-Y., Li, H.-X. and Davis, M. E. (1993) Microporous Mater. 2, 17.
13. VanderHart, D.L.(l986) J. Chem. Phys. 84, 1196.
14. Bracco, S., Comotti, A., Simonutti, R., Camurati, I. and Sozzani, P. (2002)
Macromolecules 35, 1677.
15. Sozzani, P., Bovey, F.A. and Schilling, F.C. (1991) Macromolecules 24, 6764.
16. Comotti, A., Simonutti, R., Catel, G. and Sozzani, P. (1999) Chem. Mater. 11, 1476.
17. Kageyama, K., Ng, S.M., Ichikawa, H. and Aida, T. (2000) Macromol. Symp. 157, 137
18. Wilhelm, M., Neidhofer, M., Spiegel, S. and Spiess, H. W. (1999) Macromol. Chem.
Phys. 200,2205.
19. a) Sindorf, D. W. and Maciel, G. E. (1981)J. Am. Chem. Soc. 103,4263. b) Sindorf, D.
W. and Maciel, G. E. (1982) J. Phys. Chem. 86, 5208. c) Leonardelli, S., Facchini, L.,
Fretigny, C., Tougne, P. and Legrand, A. P. (1992) J. Am. Chem. Soc. 114,6412.
20. Liu, C.C. and Maciel, G.E. (1996) J. Am. Chem. Soc. 118,5103.
21. Veeman, W.S. and Mass, W.EJ.R. (1994) NMR Basic Principles and Progress,
Springer, Berlin, Vol. 32, p. 129.
22. Chuang, I-S and Maciel, G. E. (1996)J. Am. Chem. Soc. 118,401.
23. Leon, S., Aleman, C. and Munoz-Guerra, S. (2000) Macromolecules 33, 5754.
24. Hatfield, G.R., Glans, J.H. and Hammond, W.B. (1990) Macromolecules 23, 1654.
25. Schreiber, R., Veeman, W.S., Gabrielse, W. and Amauts, 1. (1999) Macromolecules 32,
4647.
26. Schantz, S. (1997) Macromolecules 30, 1419.
27. Simonutti, R., Sozzani, P., Bracco, S. and Comotti, A. (2000) Polym. Mater, Sci. Eng.
82, 161.
Characterisation of Porous Materials by NMR

J H Strange, L.Betteridge and M J D Mallett


School of Physical Sciences,
University of Kent, Canterbury, Kent, UK.

Introduction

Porous media occur widely in nature and in technologically significant materials. They
are important, for example, in chemical engineering as catalyst supports and in reaction
chambers. Biological materials are frequently porous and this property is critical in
determining the biological functions. Performance of construction materials such as
polymer, cement and clay products is directly related to pore characteristics. Geo-
materials are also porous and fluid flow has importance from agriculture to reservoir
behaviour. Porous materials are thus of great interest for environmental and food
sciences. Two of their most important properties are porosity and pore size distribution.
Others, such as connectivity, tortuosity, wetability and surface morphology, further
define these materials and their characterisation in terms of these properties is
fundamental. Characterisation of pore sizes is frequently accomplished by gas
adsorption/desorption or mercury intrusion. NMR is gaining importance and acceptance
as a very powerful tool for non-destructive determination of many pore characteristics.
Techniques and applications that particularly inform about surface interactions and pore
morphology are reviewed and discussed below with some relevant examples of their
application.

1. Pore characterisation

1.1 PORE SIZE DISTRIBUTIONS

We have recently introduced a method of pore size measurement using NMR


cryoporometry [1] . The technique relies on the quantitative nature of NMR and the
nwlting point depression from the normal value, To, of materials in small pores,
(typically in the range of 3 to lOOnm). The melting point of a solid of dimension r is
related to the size of the solid according to the Gibbs-Thomson equation [1]:

/).T = 2yTo cosO (1)


pM-lfr
155
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 155-169.
© 2002 Kluwer Academic Publishers.
156

Where, ~T, is the freezing point depression, y, interfacial energy between the solid
phase and its liquid, p, density, H" latent heat of fusion, r, pore diameter and the a,
contact angle between the solid and its liquid phase.

Or more simply, flT oc Yr


For solids confined within pores the linear relationship between pore size and melting
point depression has been confirmed experimentally for a variety of organic materials
(Jackson and McKenna [2]) and also water. NMR provides a very convenient way of
quantifying the amount of melted solid within pores as a function of temperature and
hence offers a quantitative and non-destructive method for the determination of pore
size distributions. In NMR cryoporometry the porous material is saturated with a
suitable liquid, such as water or cyclohexane, which is then cooled until the absorbed
liquid is all frozen. The sample is then warmed and the volume of liquid recorded as a
function of rising temperature using the NMR signal from the liquid phase. This can be
differentiated from the solid using the characteristically much narrower NMR line from
the liquid phase. The signal intensity provides a measure of the pore volume associated
with pores smaller than the size given by equation (1) at each measurement temperature.
From the. temperature dependence of the signal amplitude the pore size distribution can
be established [1] . An example using the melting characteristics of frozen water in a
porous silica is shown in figure 1. Measurements have been shown to give good
agreement with the gas adsorption technique using well-defined porous silicas. In
addition to measurements on these well-characterised porous silicas, examples such as
polymers, rocks and minerals have been investigated [3,4,5]
Another NMR method that is used to assess pore dimensions relies on
relaxation time analysis of a liquid that fills the pores and the enhanced relaxation that
occurs in a liquid at the solid/liquid interface. Assuming rapid exchange of molecules
between surface and bulk fluid a measure of the surface-to-volume ratio of the confined
liquid may be obtained and this can be interpreted in terms of pore diameter. Relaxation
can be related to pore surface/volume ratios of the porous medium using the relationship
of equation (2) (Brownstein and Tarr) [6] .

lIS
-=--+p - (2)
T TBu1k V
Where, p is the surface relaxivity, S, the pore surface area, V, the pore volume.

To estimate the pore size distribution a multi-exponential relaxation decay must be


analysed into its component relaxation time distribution. To obtain pore dimensions the
pores are usually assumed cylindrical, which is often a rather sweeping assumption, and
the surface relaxivity must be obtained!
157

1.2 PARTIAL SATURATION

In order to estimate surface relaxivity, equation (2), relaxation studies as a function of


pore filling are often undertaken. To derive a value for this quantity and hence scale the
pore sizes, it is usually assumed that the filling occurs by a build up of fluid on the pore
surface layer by layer (Halperin et al) [7,8]. As we shall see, this is not necessarily
justified. Also, pores are usually assumed to be smooth and cylindrical in shape. Careful
studies by Hills et al [9,10] have also investigated partial filling of silica systems, but for
much larger pore systems of order 0.1 mm. They showed that water tended to collect in
the crevices formed when large solid silica particles met.

1.3 CRYOPOROMETRY

A series of experiments on partially filled silicas using NMR cryoporometry provides


interesting evidence for the distribution of fluids within pores. The melting point of the
small crystallites that form within the interstices of the pores (when partially filled) is
greatly reduced according to their size. Using NMR to measure the freezing
characteristics, clear quantitative evidence of the size distribution of the puddles of
water within the hydrophilic silica pores was obtained (Allen [II)). For example,
crystallite sizes of mean diameter 2nm appear when the water in 6nm porous silica
containing 20% (pore) volume of water was frozen. The mean size when the pores were
full corresponded to S.Snm.
This is in contrast to the behaviour of cyclohexane confined within the same type of
silica. In this case the size of crystallites appears within the range found for the fully
saturated silica. At the lowest saturation level studied (10%) the apparent size of pores
(i.e. the crystallite size) was at the smallest extreme of the range obtained from the 100%
filled sample. This indicated that only the smallest pores had been filled. As the
saturation level was increased the range of filled pores increased from the smaller pores
until at 100% filling all pores were seen to be full and the cryoporometry pore size
distribution curve was again produced. This indicates that the smallest pores fill first and
completely with cyclohexane, and larger pores fill as the saturation level increases.

1.8-r----------------------------------------------------1
1.5- [ Bulk liQuid ~ ·----1
1.2- ~
1.0-
0.8- [ p,rehq~
0.5-
0.2-
O.O-\::;:;:;;:;:;:;:;::;::;:;::;:y:;::;::;:::;:;:::;::;:::;:;::;::;::;::;:;:::;:;::;:;;::;;..,...,.,..,.................~...,...,...,.....,_n-rr.,....,....,........,...,......,..,..,...........-.--.._r
-15.0 -12.0 ·10.0 -8.0 -6.0 -4.0 0.0 2.0 4.0
Temperature (Celsius) ..
Figure I(a) Spin echo amplitude (arbitrary units) as a function of temperature gives the melting curve for
water in l2SA porous silica.
158
1.60-r---------------~-------...,

1.40-
1.20- Average pore size = 12SA
1.00-
O.SO-
0.60-
0.40-
0.20-
O.OO-\--:!:=.~-..,__-..,_-=:::;:===r==r===r===r==;===1
o 50 100 150 200 250 300 350 400 450 500
Pore size (Angstroms)

Figure I (b) Pore size distribution of the silica calculated from data of figure I (a).

1.4 RELAXATION

We have recently extended the relaxation technique [12] to study partially filled pores in
an attempt to further investigate the morphology of porous media, the interaction of
liquids with the solid surface and nucleation of the solid formed on freezing the liquid
within the pores. The filling process of water in porous silicas with average pore
diameters of 4, 6 and 11 nm was studied via the spin-spin relaxation time T2 as a
function of saturation. At low saturation levels two-component transverse relaxation was
observed and interpreted in terms of a surface layer of relatively immobile water (short
Tz ) with formation of small puddles (longer T2) of water in the silica interstices. The
latter grew in signal intensity and in relaxation time as more water was added until the
pores were fully saturated with water. At this point a single relaxation time only was
evident as surface water was in fast exchange with the other pore water. An example is
shown for 60A (nominal pore diameter) silica in Figure 2.

It proved possible to analyse the density of 'puddles' and their size usi~g the
relationship:

[3]

Where: p is the 'puddle' size, T2s and T 2b refer to surface layer and bulk respectively .


159

The formation, number and size of the puddles indicate the uneven nature of
the internal pore surface and provide a measure of the pore morphology. These results
contrast with experiments performed using cyclohexane instead of water in the same
silicas. In this case only one relaxation time was observed over virtually the entire range
of filling factor (2% to 100%) employed and the value of Tz was almost independent of
filling factor [12]. A similar behaviour was observed with water in silica that had been
specially prepared to provide a hydrophobic pore surface. This behaviour can be
explained on the basis that the filling process involves completely filling the smaller
pores first with others remaining empty. The relaxation time measurements and their
interpretation support and extend the interpretation of the cryoporometry results.

(a)
40 H20 in 60 Asilica .
.. . •
• •
<I
30
.
-
.-
~

••
.. .
"'-'

~ 20 ..
.,.
~

...00·

-
~
10
.. •• . ,
• 0 0 0'00 - ;
•••
... .. +- ... +, ...............
...... .......
..;. .
+ •

0
0 0.2 0.4 0.6 0.&
0
Figure 2. Transverse relaxation time as a function of filling fraction, e, of H20 in 60 A silica.

1.5 SUSCEPTIBILITY

More detailed information about the distribution and relative mobility of the absorbed
liquid can be obtained by a study of self-diffusion within the fluid in the partially filled
pores. This produces attenuation of a single spin-echo due to the strong internal field
gradients. We have modelled the gradients for various internal geometries and, as
expected, the greatest magnetic field gradients occur near the sharpest regions of
internal pore surface curvature, see figure 5, (where we predict that the small puddles of
water form at low filling factors) .
160

Both detailed resonance frequency measurements and single-echo decay time studies are
consistent with the previous results. For example, water at 5 and 10% filling shows
much enhanced decay rates compared to samples that are 50% or more filled, figure 3
(b). Cyclohexane shows little variation with filling factor and the decay rate is very
similar to the over 50% water-filled samples, figure 3 (a). The latter observation (Allen
[\3]) is consistent with the two liquids having similar diffusion coefficients and
experiencing similar internal gradients.

o~. ~Q ~ o S<nullticu"·5%
0.8 ~o (a) • SaiUllltiCXF'IO'Io
-g 07.
OJ ' 1Il~ ;. SaitDllticxFSO%
::::
0. 06
. o SaitDmioo= 100'10
~ OJ
~ 0.4
:i,
:r. 0.3
0.2
0.1
9
9
o g

o 5 10 15 20
Echo tirre (11"6)

o.~~g o SaturatiOlM%
0.8 0\ 2R (b) • SatufariOl1= I0'/0
10\ 22
D..
-$.5 0.7 l!. Satural iOl1- 0'/0

11 06 20 o Satw-ariOiFIOCWo
0. • l:J.
~ 0.5 D ••
~ 0.4
cO
Vi 0.3 o ••
0.2 •
0• ·.
0 •

0.1 o 9
o
o ODD ~ e • Q

o 5
ho tirre (rrJ g 15 20

Fi gu re 3 SPin CdlO amplitude a a function 01'% saturation for (a) Cyclohcxan~ in 200A silica alld (b) Wala
III200A Silica
161

The NMR spectrum from a fluid confined within pores is substantially different from
that of the bulk liquid. In particular the shift of the peak position in the NMR spectrum
and the change in linewidth is indicative of the magnetic environment of the liquid in
the pores.

i-. ...... . .
1 .,.
1
;
-0- Water in eoA Silica

0.& +11' . .•. Cydollexane in 60A Silica


' . • • • •• •
.

06

I
*= 0.4
.,
:!
.It

i 0.2

Saturation level
(a)

2.6

e
!
i'i
2

1.8 ""I
1.6 t ...•.....
.......•..
......... ,
...e"

§ + .......... .
1······ ···
\.4

1.2:
0
I___ + ___+-__-+_-o-
__Cydanexane
..•. in eoA. Silica
ws-+t_er_ln_15_o_A_s-jiUca

I I I
o W ~ ~ W 100
(bl Saturation 'evel

Figure 4. Results of partially tilling 60A porous silica (a) change of peak position and (b) variation of
linewidth, both as a function of saturation level for water and cyclohexane.

The variation in peak shift and linewidth as a function of saturation level, figure 4 (a)
and (b), can be used to determine the filling characteristics of the liquid in the pore
162

system. Water and cyclohexane are believed to inhabit different parts of the pore system
at low levels of saturation. Water in particular migrates to the 'nooks and crevices'
within the pore and can be distinguished by its relatively broad line and large peak shift
at very low saturation levels. Cyclohexane migrates to the smallest pores first,
completely filling them, so presenting a more homogenous magnetic medium at low
saturation levels.

The internal magnetic field gradients of a porous system can be modelled using a finite
element approach. The susceptibility differences between the three phases of a partially
filled system (solid, liquid and vapour) can create large variations of the static magnetic
field. This effect is most pronounced when there is a large material variation over very
small distances. This situation is most prevalent when the pore surface is very rough and
where, in the case of spherical silica particles, solid particles touch forming nooks and
crevices. When these small crevices are filled with liquid there can be very large field
gradients over the length scale of the crevice. This can cause a broad NMR line shape
and a reduction in the transverse relaxation time if the effect is not averaged out by fast
diffusion.

a) Cornn filling. b) Surfac(' coating.

Flu)!: 08ns.Jty
lO R! \ll"1J~~

0 ..88

0.76

0.40
0.28

0.'"
0.04
.0.08
...
--~ ~.:- - ~ - .0.20
.0.32

Figure 5. Computer modelling of the magnetic fie ld gradient variation for different pore filling regimes. (a)
fill ing of crev ices and comers, (b) fi ll ing by surface coverage.

1.6 IMAGING

The combination of cryoporometry with MRI techniques (figure 6) has already been
shown to permit visualisation Of the spatial variation of pore size distribution [14,15). In
addition 'it is now possible, using partially filled samples, to map wetability
(hydrophilicity) and surface roughness. Such morphological mapping is likely to be of
interest, for example, for petrology, catalysis and constructional materials technology.
The connectivity and pore length characteristics may be assessed by a combination of
the above techniques with NMR diffusion measurement. When a liquid-filled porous
structure is cooled the larger pores freeze first. If these are connected by smaller pores,
163

whose content remains unfrozen, the fluid so enclosed will exhibit restricted diffusion.
Indeed, if the connectivity is such that they are totally isolated then diffusion will be
bounded and the extent of the boundaries estimated. We have carried out such
measurements in the mesopores of porous silica and by holding the temperature of the
sample constant at ten degrees below the normal freezing point we were able to show
the connectivity of (5nm) pores extended for about 1 micron.

Figure 5 (a) Figure 5 (b)

Figure 6. 3 D M R cryoporometry images of oil reservoir rock cores, taken at (a) 268K and (b) 295K (colour
scale : dark grey (high porosity, medium grey (med ium poros ity), light grey (low porosity)

2. Surface interactions

We now describe two examples ofNMR measurements that illustrate applications to the
quantitative measurement of surface interaction. The first employs cryoporometry to
study the effect of surfactant grafted on ultra high purity (UHPV) silica as used in
chemical reactors. The second example explores the ~in-side-out' porosity equivalent,
namely particulate interactions with solvent as might arise, for example, with pigment
preparations and employs relaxation time methods.

2.1 UHPV MODIFIED SILICA

Nuclear Magnetic Resonance (NMR) Cryoporometry has been used to measure the
apparent pore size distributions of four ultra high pore volume (UHPV) silica samples,
A, B2, C, and D. Three have been modified with surfactants. The samples were: A -
unmodified silica, nominal pore diameter 150A, B2 - silica modified with standard
surfactant, C - silica modified with fluorinated surfactant (1), D - silica modified with
fluorinated surfactant (2). The latter fluorinated coatings were thought to be
hydrophobic.
164

Figure 7 (a)

/\
f\/
!
/\ \

Figure 7 (b)
~70'-r--'~~~--------------------~~-,r-·
OAO~

o,ao ~

OM'
aao·

_ _ _ _ _ 410_
Figure 7 (c)

030 T~-·---·

~$"i I

(UU"I
0," -
0;'10·

Q,·"'LL=::::::::::==:;=;:::::;p.~ore~
0,(10·
o 100 100_ ..
. ,.,~~~
..
'~~(.~P:~..~>
D . _.
. .• 100' 1_
Figure 7 (d)

Figure 7, Cryoporometry results for UHPV silica samples, showing the effect of surface treatment. (a)
untreated, (b) treated with standard surfactant at 100% filling, (c) as (b) but with 50% filling, (d) treated with
fluorinated surfactant (1)
165

Measurement of the pore size distribution for each sample when saturated with water
and when partial filled by 50% were undertaken. Results for samples A, B2 and Care
shown in figure 7.
The addition of a standard surfactant, sample B2, introduced smaller pores into the pore
size distribution, possibly by exaggerating constrictions in the pore morphology. The
other possibility is that the surfactant does not produce uniform coverage and produces
small pores.

Sample B2 (figure 7 (b)), modified with a standard surfactant showed a comparable pore
size distribution to the unmodified silica but is shifted to a smaller average pore size of
l25A. The surfactant seems to reduce the overall dimension of the pores by an average
of about 25A. This is interpreted as the reduction of the pore volume due to the thin
surface layer formed on the pore walls by the addition of a surface layer of thickness
about lOA. The other possibility is that the surfactant does not give uniform coverage
where the layers are thick, so smaller constrictions are formed. The amplitude of the
signal from the water contained within the pores of the modified silica (B2) was reduced
by 8% compared to the signal from the water in the unmodified silica (A) (figure 7 (a»
which is consistent with the pore size reduction.

The effect of adding the fluorinated surfactant was quite dramatic. Both samples C and
D appeared hydrophobic. The melting point curve for sample C shows a depression of
the melting point for a small volume of water indicating some pores had been filled. The
signal amplitude of this confined water was roughly 17% of the signal from the confined
water in samples A and B. This reflects the reduction in volume of water that sample C
held. If this water did in fact fill the pores present then the cryoporometry suggests that
the fluorinated surfactant in sample C reduced the pore dimensions drastically, to about
41 A. Furthermore the distribution (figure 7 (d» indicates that the reduction in pore
volume due to this surfactant may not be uniform. The pore size distribution is no longer
a Gaussian shape but has a more complex structure ranging from 25A to about 500A.
Caution is required, however, since the amplitude of the overall signal is small and
signal to noise is poor. Sample D, treated with a different fluorinated surfactant (2)
appeared to take up no water into the pores. The melting point graph showed signal
coming only from bulk water (no pore size distribution curve was obtained).

The apparent pore size distributions of the under filled sample B2(50%) (figure 7 (c))
was the same as the previous overfilled and 100% filled samples. This was to be
expected from the work on partial filling with little differences appearing until filling
fractions below 30%. The under-filled surfactant-treated sample, B2, did exhibit some
stmcture to the pore size distribution. A small shoulder appeared between 70A and 85A
which would correspond to a number of small isolated puddles of water present. This
addition to the pore size distribution may be caused by the surfactant polymer strands
extending from the silica surface, rather like the bristles of a bmsh. The surface layer
may accentuate constrictions in the pore morphology sufficiently to produce distinct,
166

almost isolated, regions within the pores. Another possibility is that the surfactant
strands are trapping small puddles of water. It is clear from the similarities in the pore
size distribution of the modified and unmodified samples that the surfactant does not
stop water filling the silica pores. Although the difference in the pore size distribution
curves is small, it is reproducible.
The unmodified silica (sample A) was measured to have a pore size distribution about
an average pore size of 150A. The addition of the standard surfactant (sample B2) seems
to reduce the overall dimension of the pores by, on average 25k Sample C, fluorinated
surfactant (1), absorbed approximately only 17% of the water samples A and B2
absorbed. If the pores were filled, the pore size was measured to be about 4oA, a drastic
reduction in size to samples A and B2. Further the results show a distribution of pore
sizes being introduced up to 50oA. Sample D, fluorinated surfactant (2), did not appear
to have any confined water present. Partial filling measurements highlight the
introduction of smaller pores present in the modified silica B2. These smaller pores are
possibly formed by the addition of the surfactant that accentuates the constrictions in the
pore morphology and allows them to be distinguished in the pore size distribution.

3. Liquid interactions at solid surfaces

Another technologically significant system consists of solid particles held in suspension


within a liquid matrix. Typical examples of this sort of system include adhesives (e.g.
gap filling glues, cement, concrete), pigments (e.g. paint, ink) and foodstuffs (e.g.
custard, sauces, spreads) among others. One of the most important aspects of this type of
system is the ability of the liquid matrix to 'wet' the solid particles (very often the
definition of 'solid' in such a system is purely in the sense that there is preferential
agglomeration of one of the phases). The wetting properties can be modified by the
addition of surfactants.

A study has been undertaken using NMR relaxation techniques on a simple


solvent/surfactant/solid system. Two different solid phases were used, one of them was
an organic solid, the other had a more polar nature. An organic solvent (Maranil ®) was
used. The expected behaviour was to observe an increased 'wetting' effect of the organic
solid surface with the organic solvent compared with the 'wetting' of the solvent on the
surface of the polar solid. The effect of adding increasing amounts of surfactant to the
system was monitored. The solid particulate phase had a very broad proton NMR
linewidth and was not measurable in these experiments. The NMR signal from the
surfactant was <2% of the total and so has a negligible effect on the measurement.
Therefore the change in the NMR relaxation behaviour is, to a first order, a measure of
the effect of increasing the surfactant concentration on the solid/solvent interaction. The
longitudinal and transverse relaxation times were measured as a function of surfactant
concentration. The T2 values are much less than those for T] (figures 8 and 9). The two
systems displayed different relaxation behaviour. The longitudinal relaxation time can
be considered to be a measure of the surface interaction between the solid particulate
167

and the liquid matrix, whereas the transverse relaxation time provides information on
the mobility of the liquid matrix.

3.1 TI MEASUREMENTS

Polar system: At zero percentage surfactant concentration the polar system has a
relatively long T 1• This would indicate that the solvent is not efficiently wetting the
surface of the solid and so does not experience an enhanced, surface induced, TI decay.
As the surfactant concentration increases there is more efficient coupling between .the
solid surface and the solvent leading to an enhanced TI decay.

Organic system: In contrast to the polar system, at zero surfactant concentration the
solvent couples efficiently to the organic surface leading to an enhanced TI relaxation of
the solvent. As the surfactant concentration increases there is a reduced coupling
efficiency between the solvent and the solid surface, this causes TI to increase.

Tt relaxation times as a function of surfactant concentration

3000

~Organ i c solid ,
l-- - - - - -I
2500
Polar solid
2000
~

'" 1500
.§.
!o-
1000

500

0
0 5 10 15 20
% surfactant

Figure 8. Longitudinal relaxation time as a function of added surfactant.

In the two systems studied the surfactant preferentially binds to the surface of the
particulate solid and at high surfactant concentrations the solid surface is effectively
shielded from the solvent by the surfactant layer. This leads to a convergence of the TI
relaxation times for both the polar and organic solid system as the solvent 'sees'
agglomerations of surfactant molecules rather than the solid particulates.
168

3.2 T2 MEASUREMENTS

Polar system: the transverse relaxation time decreased monotonically with increased
surfactant concentration. This is interpreted as a decrease in mobility of the liquid matrix
as the number of, relatively large, surfactant molecules tethered to the solid surface
increases.

Organic system: similarly to the polar system, the transverse relaxation time was seen to
decrease as a function of increased surfactant concentration. This can again be attributed
to a decrease in the solvent mobility as the number of surfactant molecules increases.

T2 relaxation time as a function of surfactant concentration

e0 r-------------------________________________ ~

60

50

'"
~ "0
....
I-
30

20
--+-Organic solid
'0 - . - Polar solid

0
0 '0 '2 ,. ,. " 20

OJ,, surractant

Figure 9. Transverse relaxation time as a function of added surfactant.

4. Conclusion

It is clear that NMR cryoporometry and relaxation measurements compliment each other
in the study of porous media and provide a valuable non-destructive suite of techniques
for the characterisation of porous materials. In particular much can be learned about
surface interaction within pores by studies as a function of saturation. With the addition
of MRI methods microscopic properties can be mapped on a macroscopic scale.

The experiments described in section 3 illustrate the potential for surface interaction
studies. The addition of a standard surfactant to porous silica appeared to reduce the
overall pore dimensions. The absorption of hydrophobic surfactants was seen to exclude
most of the water but the trapping of fluid by the surfactant and the uniformity of
surface coating is amenable to study. The effect of an added surfactant on a solid
169

particulate / liquid matrix system can be monitored by the change in the NMR relaxation
time behaviour of the system. The addition of a surfactant phase to a polar solid /
organic solvent system was shown to create a closer coupling of the solvent to the solid
surface than if the surfactant were omitted. The opposite was observed when the
surfactant was added to an organic solid / organic solvent system. In this situation the
surfactant binds to the solid surface in preference to the solvent. Without the surfactant
the coupling between the solvent and the solid surface is efficient. When the surfactant
is added the solvent is 'pushed' further away from the solid surface and so the coupling
between the two is less efficient. This is observed as an increase in the T, relaxation time
of the solvent in the presence of the solid particulates.

Acknowledgements

The authors would like to thank EPSRC for supporting much of this work. L.Betteridge
would like to thank EPSRC and Lafarge-Braas for the award of a CASE studentship,
M.J.D.Mallett would like to thank CMPT for funding the rock core imaging study. The
authors also acknowledge the help of Dr.E.Sang at the University of Reading for
supplying the UHPV silica samples.

References

I J.H.Strange, M.Rahman and E.G.Smith, Phys.Rev.Lett., 71, 3589-3591 , (1993)


2 C.L.Jackson and G.B.McKenna, J. Chern. Phys., 93, 9002 (1990)
3 C.Choi , B.J.Balcom, S.D.Beyea, T.W.Bremner, P.E.Grattan-Bellew, R.L.Armstrong, J. App.
Phys., 88, 3758 (2000)
4 A.V.Filippov,A.M.Zaripov, G.G.Pimenov, Colloid J., 62, 372 (2000)
5 E.W.Hansen, R.Schrnidt and M.Stocker, 1. Phys. Chern., 100,11396 (1996)
6 K.R.Brownstein and C.E.Tarr, Phys. Rev. A, 19,2446 (1979)
7 F.D'Orazio, S.Bhattacharja, W.P.Halperin and R.Gerhardt, Phys. Rev. Lett.. 63,43 (1989)
8 F.D'Orazio, S.Bhattacharja, W.P.Halperin, K. Eguchi and T.Mizusaki, Phys. Rev. B, 42, 9810
( 1990)
9 B.P.Hills, P.S.Benton and V.M.Quantin, MoI.Phys., 78, 893 (1993)
10 B.P.Hills and J.E.M.Snaar, Mol. Phys, 86,1137 (1995)
II S.G.Allen , P.C.L.Stephenson and J.H.Strange, 1. Chern. Phys., 108,8195 (1998)
12 S.G.Allen, P.C.L.Stephenson and J.H.Strange, J. Chern. Phys., 106, 7802 (1997)
13 S.G.Allen, M.1.D.Mallett and J.H.Strange, 1. Chern. Phys., 114,3258 (2001)
14 M.J .D.Mallett, B.Clennell, J.H.Strange and Q.Fisher, Report to CMPT, (2000)
15 J.H.Strange and J.B.W.Webber, Appl. Magn. Reson., 12, 231 (1997)
Keynotes and oral presentations
INVESTIGATION OF RADICAL PAIRS IN MICELLES USING
SPIN POLARIZATION TECHNIQUES

E.G.BAGRA YNSKA YA, N.V.LEBEDEVA, M.V.FEDIN,


R.Z.SAGDEEV
International Tomography Center SB RAS, Insitutskaya 3a
Novosibirsk, 630090, Russia

1. Introduction
Micelles represent a simplified model of biological membranes; therefore significant
effort is devoted to establish their properties and structure. The micelles can serve as
"limited volume micro reactors" for radical reactions. In micelles partners of radical
pair (RP) during their lifetime are separated by a few angstroems, and, hence, electron
exchange interaction as well as spin selective reaction rate significantly affect magnetic
field effects. A large number of studies published during the last few years report the
observation and investigation of the magnetic field and magnetic isotope effects, very
strong nuclear and electron spin polarization and other phenomena for the radical pairs
confined within micelles [1-15]. Investigation of the micellized radical pairs attracts
significant attention since the intersystem crossing in such RPs is critically affected by
the exchange interaction and electron spin relaxation. This opens up the possibility to
study both the exchange interaction and its influence on the intersystem crossing of
RPs. The spin dynamics and chemical kinetics of micellized RPs have been studied by
a number of techniques, e.g. Magnetic field Reaction Yield (MARY) [3], Time
Resolved Electron Spin Resonance (TR ESR) [5,7], Chemically Induced Dynamic
Nuclear Polarization (CIDNP) [8], Reaction Yield Detected Magnetic Resonance
(RYDMR)[3,9], Product Yield Electron Spin Resonance (PYESR) [10,11], Stimulated
Nuclear Polarization (SNP) [12-18], CIDNP with the fast Switching of External
Magnetic Field (SEMF CIDNP) [26,27],etc.
This paper summarizes the results of the investigations of radical pairs (RPs) in
micelles by indirect methods of magnetic resonance CIDNP, SNP and SEMF CIDNP.
SNP technique. The distinctive feature of SNP method is the detection of the
NMR signals of reaction products in the presence of a microwave field. It has been
shown earlier [12-15] that the application of the SNP technique to the study of spin
dynamics and chemical kinetics of the RPs confined within the micelles allowed to
obtain the parameters of exchange interaction for a number of RPs, to measure their
lifetimes and to make conclusions on the mechanism of electron spin relaxation. In the
SNP technique, the resonant microwave field affects the rate of singlet - triplet
conversion within RPs and thereby alters the measured polarization of nuclear spins in
diamagnetic reaction products. The SNP spectra essentially represent the EPR spectra
of RPs; therefore they are more sensitive to the effects of the exchange interaction and
spin-selective decay on the intersystem crossing within the micellized RPs as compared
to the CIDEP technique, which mostly reflects the contribution of free radicals. It has
173
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science. 173-184.
© 2002 Kluwer Academic Publishers.
174

been demonstrated earlier [12,13] that the SNP spectra of micellized RPs under some
conditions are sensitive to the reencounter frequency of the radicals, which depends on
the micelle size and intramicellar viscosity.
SEMF CIDNP. A method of Chemically Induced Dynamic Nuclear Polarization in a
Switched External Magnetic Field (SEMF CIDNP) involves abrupt field switching
during the lifetime of RPs. In this method, after initiating a reaction with a laser pulse,
CIDNP is formed during the variable time delay to in initial magnetic field Bo, and then,
after switching, CIDNP formation goes on in the fmal field B\. As the CIDNP intensity
is different in the initial and final magnetic fields, and the intensity of polarization,
formed as a result of recombination of RPs, is a function of the time delay of field
switching to, the dependence of CIDNP on the time delay to reflects the kinetics of RPs.
Time resolution of the method is determined by both the switch rise-time which can be
about 5 ns, the duration of the laser pulse and jitter between laser pulse and the rise of
the switching field.

2. Experimental.
The experimental SNP and SEMF CIDNP setup has been described earlier (15]. The
reaction mixture was irradiated by laser pulses (Lambda Physik COMPex, A=308 nm
20 Hz, 40 mJ) in photochemical cell located in the field of a magnet (0-78 mT). A flow
system was used to transfer the irradiated mixture to the probe of a Bruker MSL-300
NMR spectrometer. In time resolved experiments the dependence of the polarization on
variable time delay between the laser pulse and RF or magnetic field pulse has been
investigated.
The duration of the RF pulse (f=1530 MHz) in SNP experiments was 5 /is and the
RF pulse edge was 20 ns. To perform the experiments at temperatures above and below
the room temperature, the sample was residing within a stream of warm air or cold
nitrogen gas, respectively. The temperature within the sample cell was monitored by
means of a calibrated thermocouple.
For SEMF CIDNP experiments the external magnetic switching field Bs parallel to
Bo was provided by Helrnholz coils. The maximum Bs field amplitude was ±7.0 mT
and the switch rise-time was about 5 ns. The back edge duration, which has not to be so
short, was about 200 ns, with a total pulse duration being 2 - 5 /is. An experimental
jitter time of about 20 ns existed between the laser pulse and the rise of the switching
field and determined the resolution time of our setup. Note that this time is negligible
on the timescale of kinetics involved.

3. Theoretical model
For the calculation of magnetic field dependencies of stationary CIDNP and SNP
spectra, model calculations based on the numerical solution of the Liouville equation
[14] have been used. To describe the dynamics of RPs confined in the micelles, the
microreactor model was employed [19]. It assumes that one of the radicals is
permanently located at the center of a spherical micelle of radius L, whereas its partner
diffuses freely within the micelle volume with diffusion coefficient D and escapes into
the water bulk. It is assumed that an escaped radical does not re-enter the same micelle.
Along with the terms describing the Zeeman interaction of the electron spins and
Hyperfine Interaction (HFI), the hamiltonian flo includes the term describing the
175

exchange interaction which exponentially depends on the interradical distance


J(r)=Jo·exp(-(r-Ro )/).), where Ro is the radius of the reaction zone. The
characteristic length scale of the spatial overlap of electronic orbitals upon formation of
chemical bonds was used as the parameter of the spatial decay of exchange interaction,
o
).=0.5 A. Two mechanisms of electron spin relaxation were taken into account: the
dipole-dipole interaction of the electron spins and the fluctuations of the local magnetic
fields (anisotropy of g- and HFI-tensors, spin-rotational coupling) characterized by the
magnitude, G, and the rotational correlation time, Te. For acyl radicals, the main
relaxation mechanism is the spin-rotational interaction [20], which contributes along
with the relaxation induced by the anisotropic part of HFI. The anisotropic part of the
Zeeman interaction can obviously be neglected, since the magnetic field Bo is weak.
The description of the relaxation caused by the fluctuation of the local magnetic fields
follows the approach described earlier for the calculation of CIDNP magnetic field
dependencies and SNP spectra ofbiradicals [21). The two channels of the RP decay are
considered: (i) an exponential decay, e.g. due to decarbonylation, (ii) and the escape of
a radical from the micelle with the rate constant k ese. The description of the dipole-
dipole interaction was based on the approach developed by Steiner and Wu [22]. The
differential scheme proposed by Pedersen and Freed [23] was employed to solve the
equation. The RP recombination was characterized by dimensionless parameter ks Ts,
where Ts =RotS/ D is an average RP residence time in the reaction zone of thickness t5,
and ks is the RP recombination rate in this zone. The values of Jo, ks Ts, D, L, kese, G, Te
were treated as variables in the calculations performed.
For the calculation of Time-Resolved SNP kinetics of micellized RPs two different
approaches are usually employed. A quantitative description of Time-Resolved SNP
experiments on studying the micellized radical pairs in terms of a kinetic scheme has
been proposed by Avdievich et al. [24]. More correct approach to the theoretical
description of the SNP kinetics is the numerical solution of the Liouville equation for
the density matrix. It has been shown in ref. [25] that under conditions of rapid
intersystem crossing and large rate constant of decay in the reaction zone the SNP
kinetics exhibit monoexponential decay with a decay parameter kobs =krel+kese being the
sum of the rate constants of escape kese and relaxation k rel and almost insensitive to other
parameters.
The theoretical description of the CIDNP of radical pairs in restricted volume, in
experiments with switched external magnetic field was developed by Parnachev et al
[26] and Fedin et al. [27). The general analytical expression for calculation ofRPs with
limited mobility taking into account the electron paramagnetic relaxation has been
obtained. In micelles the RP lifetime is long (0.1-1 IlS [28]), thus S-To , S-L, S-T+
transitions, occurring during free diffusion of radicals inside a micelle, are usually
strongly averaged. The contribution of S-T _-transitions, occurring in the zone of energy
level crossing [8] as well as electron paramagnetic relaxation, caused by electron
dipole-dipole interaction and modulation of hfi anisotropy that causes T±~To, S -
transitions had been taken into account. We have shown that for the limit case where
the rate of S-L-conversion, kJc is much slower than the relaxation rate, krel (kJc«k rel ),
the obtained expression for CIDNP kinetics is: P(to) = A· exp( -(kesc + k rel )· to) . For
the opposite limit case kJc»krel: P(to) =B· exp( -(kesc + k lc )· to) + const.
176

The expressions obtained have a very clear physical meaning. As the dominant
contribution to CIDNP is due to the S-T_ mechanism, kinetics reflects the time
dependence of L-Ievel population. For a low rate ofT_-S conversion the decay of the
T _-level is determined by the sum of escape and relaxation rate constants. Note, that in
this case the CIDNP time dependence obtained is the same as for SNP kinetics. SNP
kinetics reflects the time dependence of Land T+ levels population. For a low L-S
transition rate one expects the population of T_ and T+ levels to decrease by a similar
amount. For slow electron relaxation rate the population of the L level decays with the
escape rate, while the decrease of the population of T _ level is determined by k lc ' Thus
the comparison of experimental kinetics measured by SEMF CIDNP with that
measured by flash photolysis or SNP can provide information on the rate of S-T_-
conversion, exchange interaction, electron paramagnetic relaxation etc.

4. Results and discussion


4.1 . ELECTRON SPIN EXCHANGE IN MICELLISED RPs.
The 'H, I3C and 31p SNP techniques were used to study the radical pairs formed
during the photolysis of a-methyldeoxybenzoin, dibenzyl ketone, benzoin,
deoxybenzoin, benzophenon in the presence of tert-butylphenol, (2,4,6-
trimethylbenzoyl)diphenylphosphine oxide, dimethylacetonedicarboxylate, 2,2' ,4,4 '-
tetramethyl-dimethyl acetonedicarboxylate in alkylsulfate micelles Na +S03"
O(CHz)nMe of various sizes (n=6-11). As an example Fig.l and Fig. 2 shows the results
T
-------
Oy,-q-:Q
o 00

Oy,o -0 + H
[~
-0
-y,
0
II

ppm
200 195 190 185 180 175

Figure 1. The scheme oJthe photolysis oJbenzoin in micelles. The J3C CIDNP spectrum
obtained upon the photolysis OJllC benzoin.
177
of the investigation of micellized RP fonned upon the photolysis of benzoin
(Ph 13 qO) 13CH(OH)Ph).
The significant differences in the ratios of the HFI constants and the lifetimes of
these radical pairs allowed us to examine the influence of the exchange interaction on
the pattern of the SNP spectra. It was discovered that the line splitting and line width in
the spectra depend on the size of micelles. In small micelles (n=6-9) for radicals with
HFI constants a-2-12 mT the decrease in the spectral splitting is due to both effective
exchange interaction and spin selective decay of the radical pairs. In the larger micelles
(n=10, 11), the splitting in the spectrum is equal to the HFI constant, and the exchange
interaction influences only the widths of individual lines. For RPs with large HFI
constants a-30-40 mT, e.g. in the limiting case where the characteristic time of
intersystem crossing markedly exceeds the characteristic time period between re-
encounters in the micelles, splitting of lines in the SNP spectra does not depends on the
micelle size.

Magnetic field (ruT)

Figure 2. SNP spectra measured during the photolysis of J3e benzoin detected via
J3e NMR Signal of product III (a-in Sodium Dodecylsulphate (SDS) micelles, b-in
Sodium Octylsulphate (SOS) micelles) and for product I (c-in SDS micelles, d-in
SOS micelles). Solid lines are simulated spectra with following parameters Jo=-60
mT, J.,=0.5 A, k.c.=1 .6, L=15.4A (SDS), D=0.8 ·106 cm 1/s (SDS), L=JJ .6A (SOS),
D=1.49 ·J06 cm1/s (SOS).
The main parameters detennining the molecular dynamics of radicals are the
coefficient of mutual diffusion D, micelle radius L, the radius of effective radical
recombination R, and the rate constant for the radical escape to the bulk kesc . Splitting in
the SNP spectra depends on the ratio of the following values: the HFI constants, the
characteristic time between the collisions of radicals inside the micelles T2~L3/3RD, the
exchange interaction and the rate of the spin selective decay. Upon a decrease in the
micelle size, the time the RP spends in the reactive state increases in comparison to the
time spent in the non-reactive state. Thus the time-average exchange interaction and the
time-average recombination rate increase.
178

From the comparison of experimental and calculated SNP spectra in low and high
magnetic fields the parameters of exchange interaction of micellized RPs have been
obtained [12,13,24,25,27]. The parameters of exchange interaction obtained under the
investigations of micellized RPs formed during the photolysis of a number of
photochemical reactions in alkyl-sulfate micelles of different sizes are summarized in
Table 1.

N2 RP JomT n SNP koos.1 0" s' SEMFCIDNP k,et·IO"s· (k",,+k.<o)ol 0'· 5. 1


koos.1 0-6 5. 1
la Ph I3 CO••CH(CH 3)Ph 60.0 12 7.I±O.1 7.5±O.5 5.5 1.6
Ib [12],[13],[27] II 7.6±O.1 S.8 1.8
Ie 9 8.4±O.1 6 2.4
Id 8 8.6±O.1 6.1 2.S

2a PhCH/ 3CO ••CH 2Ph 60.0 12 17±2 12.0±1.2


2b [12],[24],[18] 8 IS±!
3a Ph I3CO••CH 2Ph [121 60.0 8 7.5±1.4 6.1 1.4
4a Ph 13CO ••CH(OH)Ph 60.0 12 8.2±O.2 5.5 2.7
4b 8 11 ± 3 6.1 4.9
Sa PhCO•• PO(Ph)2 91.0 12 6.6±O.6
Sb (17] 8 16±4
6 40.0 12

H~ O~ ••

6a RPI 0.84±0.04 0.67S 0.19


(Bo=2mT)
6b RPII 0.4S±O.01 1.13±O.04 0.26
(Bo=S3.SmT) (Bo=2mT) (Bo=S3.SmT) 0.19
[25],[27],[16] 0.82
(Bo=2mT)

7 CH30C(O)CH2· 12 4.2±O.1 4.59±O.03


-CH2CH(CHd·
8 CH3OC(O)C(CH3)2· 12 8.S±O.1 8.3±O.2
CH3OC(O)C(CH3)2CO.

4.2. LIFETIMES OF MICELLIZED RADICAL PAIRS


Time-resolution of SNP experiments is determined by the RF pulse edge and is
equal to 20 ns. The lifetimes of micellized radical pairs lie in the range from ten to
hundreds of nanoseconds and thus their kinetics can be studied using SNP technique.
The observed kinetic curves for time-resolved SNP for the number of RPs are described
by an exponential function (with the exception of the initial parts) in accordance with
theoretical predictions. In the case when the lifetime of excited triplet molecule is
comparable with RP lifetime the SNP kinetics exhibit plateau at short time delays. The
initial parts of the curves are associated with the lifetime of the triplet precursor and the
duration of the laser pulse. Analysis of the resulting time dependencies made it possible
to estimate the relaxation and escape rates, and lifetimes of triplet molecule. The
observed rate constants of RP decay as well as the relaxation and escape rate constants
179

13 1.0 o RP 5b
for a number of investigated RPs
• RP Sa
§ 0.8 are listed in Table 1. Typical
..
.ci
.. 0.6
• RP 7 SNP kinetics obtained for RPs
formed in the photolysis of
~il 0.4 (2,4,6-trimethylbenzoyl)
] diphenylphosphine oxide
~ 0.2
rn (PhCOPO(Phh) and
dimethylacetonedicarboxylate
800 1000 1200 1400
(CH 30 C(O)CH2C(O)-
time delay, os
13= =1.0 o RP 8
CH 2C(O)OCH3) are shown in
Fig. 3a. The observed rate
t!
.... 08 .. RP 6b constants of RP decay formed
o
";.. 0.6 o RP 6a upon the photolysis of (2,4,6-
.5 trimethylbenzoyl)
~ 0.4 diphenylphosphine oxide are
Q
0", 02. determined mainly by the rates
:!: of spin-rotational relaxation of
1;l 00 -\,-Je.....-O'---O'~;:::;:i=..;;::::;::=;::=~ PhCO. and .PO(Ph)2 radicals.
2000 4000 6000 8000 10000
time delay, os The rate of the relaxation of RP
formed upon the photolysis of
dimethylacetonedicarboxylate is
Figure 3. a) SNP kinetics upon the photolysis of rather slow therefore we propose
dimethylacetonedicarboxylate in SDS micelles and that the observed rate constant of
(2,4,6-trimethylbenzoyl) diphenylphosphine oxide in RP decay is determined by the
SDS and SOS micelles. rate of radical escape from the
b) SEMF CIDNP kinetics measured during micelles.
photolysis of 2,2 ',4,4 '-tetramethyl-dimethyl The possibilities for
acetonedicarboxylate and BPITBP in SDS micelles: application of SEMF CIDNP
experiment (symbols) and simulation (lines) with the method for obtaining the
follOWing set of parameters: RP 6a: molecular and spin dynamics
k'c=4·]02 s·', kre,=6.7 ·]05 s·', kesc =2 ·]05 s·', parameters of micellized RPs has
ks=4·]0 s·'. RP 6b: k'c=]·]04 S·', kre,=9.2·]OS s·', been
7 tested using three
5
kesc=2·10 s·', ks=4·] 0 S·' . 7 photochemical reactions [27].
Three photochemical reactions
have been chosen which
correspond to different ratios
between rates of escape from
micelle, electron relaxation and S-L-conversion, and satisfy the requirement of the
approximation of theory, which takes into account only one magnetic nucleus. These
reactions are the photolysis of l3C (90%) carbonyl labelled dibenzyl ketone (DBK) and
t3C (99%) carbonyl labelled a-methyldeoxybenzoin (MDB) photolysis in sodium
dodecyl sulphate (SDS) micelles and the photolysis of benzophenone (dlO) (BP) in a
presence of 2,4,6-tri-tert-butylphenol (TBP) with a natural abundance of l3C atoms.
The obtained parameters of relaxation and escape rate are listed in Table 1. Values of
relaxation and decarbonylation rate constants obtained agree well with the literature
data obtained by other techniques, which is a confirmation of validity of the SEMF
180

CIDNP method. As an example, Fig. 3b shows the SEMF CIDNP kinetics of RPs
formed upon the photolysis of BP in a presence of TBF and 2,2',4,4'-tetramethyl-
dimethyl acetonedicarboxylate in SDS micelles. One more advantage of SNP and
SEMF CIDNP techniques is the possibility to detect spectra of several different RPs in
the same experiment. The comparison of the kinetics detected by the NMR lines of
different RPs during the photolysis ofBP in the presence ofTBP (fig.3b) allowed us to
obtain the electron relaxation and escape rate constants of both RPs.

4.3. DIFFUSIVITY OF MICELLIZED RADICALS


SNP and SEMF CIDNP give an unique information about the behavior of
micellized RPs [18]. It has been demonstrated earlier [12,13] that the SNP spectra of
micellized RPs under some
conditions are sensitive to the
reencounter frequency of the
radicals, which depends on
[J the micelle size and
intramicellar viscosity. The
size, shape and the
aggregation number of
micelles depend on the
structure of the detergent
molecules and it's
45 50 concentration, on the
concentration of a salt,
temperature, etc. Since the
temperature change alters the
micelle size and intra-
micellar viscosity and there-
fore affects the reencounter
Figure 4. SNP spectra detected via carbonyl J3 C NMR frequency, the shape of the
signal of MDB upon the photolysis ofMDB in SDS (a) SNP spectra is expected to
and SOS (b) micelles at various temperatures. Solid exhibit a pronounced
lines represent the simulated spectra obtained for the temperature dependence in
following set of parameters: Jo=60 mT, 4=0.5 A, the cases when the time of
ksTs=1.6. For T=450C L=9.2A, D=3.3 ·10 cm"/s;for
6 radical reencounters is
T=1 ic L=10.6A, D=1.35 .10 cm"/s; for T=10oC
6 comparable with the time of
triplet-singlet conversion. It
L=1 J.JSA, D=1 .1·106 cm"/s.
was found that the shape of
the I3C SNP spectra, detected in the photolysis of dibenzylketone (DBK) and u-
methyldebxybenzoine (MDB) exhibits a noticeable temperature dependence (Fig.4).
For the relatively large sodium dodecylsulphate (SDS) micelles the temperature
increase leads to the increase in the linewidth of SNP spectra, while in the smaller
sodium octylsulphate (SOS) micelles the splitting in the SNP spectra decreases.
SNP spectra obtained upon DBK photolysis in SOS micelles with added NaCI are
shown in Fig.5. In SOS micelles NaCI additions induced the increase of the splitting
from 8.3±0.1 mT at NaCI=O to 11.4±0.1 mT at NaCI=2.68 M in full agreement with the
181

expected behaviour of the SNP spectrum upon the increase of micelle radius. In SDS
micelles NaCl additions have no effect on SNP spectra obtained via NMR line of
carbonyl carbon (a=12.4 mT). The main reason for that is the low sensitivity of SNP
spectra to the changes in the rate ofreencounters (Z) when lla«lIZ, which is the case
for a=12.4 mT in SDS micelles in the presence ofNaCl.
From the comparison of the experimental data with the model calculations based on
the numerical solution of the Liouville equation, the influence of the temperature and
salt additions on the collision rate of radicals in micelle interior was evaluated. Using
literature data on aggregation number at different temperature and salt concentration,
the temperature and
salt additions
o C(NaO)=O dependence on the
"" C(NaO)=O.6 M transla-tional
o C(NaO)=3M diffusivity of radicals
inside the micelle has
been obtained. The
obtained activation
45 50 energy of the transla-
tional diffusion of
small radicals is
E.=6.7±0.25 kcal/mol
and E.=5.6±0.2
kcallmol for SDS and
SOS micelles,
Figure 5. SNP spectra detected via carbonyl J3 C NMR
respectively. These
signal of DBK upon the photolysis of DBK in SOS micelles
values are very close
at various concentrations of NaC!. Solid lines represent to those obtained in
the simulated spectra obtainedfor the following set of [29]: E. =5.8 kcallM
parameters J o=60 mT, A,=0.5 A, ksr.=1.6. For C(NaCl)=O where dibenzyl ester
L=JO.JA, D=I. 7.106 cm2/s;for C(NaCl) =0. 6M L=1 J.5A, has been used as a
D=J.5·J06 cm2/s,for C(NaCl)=3M L=J2.8A, D=J.J5·J06 fluorescence probe.
cm 2/s. Note that the size of
dibenzyl ester is close
to the size of radicals used in this work. The activation energies obtained in our work
are in a good agreement with the activation energies of rotational correlation times ic
values of the nitroxide radicals [30], (E.=6.5 kcallmol for SDS and E.=5.9 kcal/mol for
SOS). As has been mentioned above, the shape of the SNP spectra is determined by the
rate of radical reencounters and thus the translational motion of radicals determines the
obtained viscosity only. The good agreement between our experimental data and the
viscosity obtained using rotational correlation times of radicals of sizes close to those
used in our work shows the applicability of the Debye-Stokes-Einstein formula for a
micelle. Thus despite the substantial heterogeneity of micelle interior, the translational
o
and rotational correlation times for the radicals with radius of 1.5-3 A are similar.
Besides, the influence of temperature and salt additions on the lifetimes of the
micellized RPs formed in the photolysis of MDB and DBK has been studied
experimentally. It is shown that the temperature dependence of RPs lifetime formed in
182

the photolysis of MDB is determined mainly by the changes in the escape rate. This
conclusion is confIrmed by the fact that the activation energy Ea=6.4±0.3 kcallmole of
the observed decay rate is very close to the activation energy of the escape rate of ketyl
radical in SDS micelles Ea=6.5±1.8 kcallmole obtained by Evans et al. [31 ].The
observed decay rate for SDS micelles is equal to kobs =1.1·10 7 S·I, and is in a good
agreement with literature data for decarbonylation rate in non-polar solvents [32].
The salt additions have no effect on the decay rate of SNP kinetics of RPs formed in
DBK photolysis in SDS. In SOS micelles the decay rate increased upon salt additions.
The RPs lifetime in the case of DBK photolysis is mostly determined by
decarbonylation rate. As is known [32], the decarbonylation rate signifIcantly depends
on polarity: the increase of polarity leads to the decrease of decarbonylation rate. Thus
the main salt effect consists of the changes in the decarbonylation rate due to the
decrease of polarity of the micelle interior. The independence of the decay rate for SDS
micelles on NaCI concentration (0.1-0.5 M) leads to the conclusion that the changes of
polarity averaged over the volume of micelle interior are negligible. For smaller SOS
micelles the changes in the interior polarity are substantial which leads to the increase of
decay rate by about a factor of 1.7 when NaCI concentration changes from 0.025 M to
1.1 M. Note that in the absence of decarbonylation reaction the increase in the observed
rate is expected due to the growth of micelle size upon salt additions.

5. Conclusions
The examples given in this review demonstrate the applicability of SNP and SEMF
CIDNP techniques to the study of chemical kinetics and spin dynamics of micellized
radical pairs. In particular, these techniques allow us to evaluate electron exchange
interaction, radical reencounters rate, electron spin relaxation rate, escape rate and
microdiffusivity of radicals within the micelles. SNP and SEMF CIDNP techniques
prove useful in those cases where the use of flash-photolysis is diffIcult due to the
overlap of optical absorption bands. An additional advantage of both techniques is the
spectral resolution of NMR, which allows one to follow the transformation of radicals
into diamagnetic products. The highest effIciency in the applications of these methods
can be obtained by combining them with the techniques of laser flash photolysis and
Time-Resolved EPR.

6. Acknowledgement
The authors thank INTAS 99-01766, the Russian Foundation for Basic Research
grants N 99-03-33488, 99-03-32459 for fInancial aid.

7. References

1. Gould, I.R., Zimmt, M.B., Turro, N.J., Baretz, RH., Lehr, G.F. (1985) Dynamics of
RP reactions in micelles, J.Amer.Chem. Soc. 107,4607-4612.
2. TUITo, N.J., Tarasov, V.F., Buchachenko, A.L. (1995) How Spin Stereochemistry
Severely Complicates the Formation of a Carbon-Carbon Bond between Two
Reactive Radicals in a Supercage, Accounts Chem.Res. 28,69-74.
183

3. McLauchlan, K.A., Nattrass, S.R. (1988) Experimental studies of the spin-


correlated radical pairs in micellar and microemulsion media; MARY, RYDMR Bo
and RYDMR BI spectra, Mol.Phys. 65, 1483-1503.
4. Turro, N.J. (1981) Micelles, magnets and molecular mechanisms. Application to
cage effects and isotope separation, Pure & Appl. Chem. 53, 259-286.
5. Closs, G.L., Forbes, M.D.E., Norris, J.R. (1987) Spin-polarized electron
paramagnetic resonance spectra of radical pairs in micelles. Observation of
electron spin-spin interaction, J.Phys.Chem. 91, 3592-3599.
6. Van Willigen, H., Levstein, P.R., Martino, D., Ouardaoui, A., Tassa, C. (1997) FT-
EPR study of photoionization in micellar solution, Applied Magnetic Resonance
12, 395-398.
7. Wu, lQ., Baumann, D., Steiner, V.E. (1995) Interference of heavy-atom with
magnetic spin effects in spin-correlated micellar radical pairs, Mol.Phys. 85, 981-
985.
8. Zimmt, M.B., Doubleday, C.Jr., Turro, N.J. (1984) Energetics and dynamics of
radical pairs in micelles. Measurement of the average singlet-triplet energy gap by
means of the magnetic field dependence of \3C CIDNP, J.Am.Chem.Soc. 106 ,
3363-3365.
9. Batchelor, S.N., McLauchlan, K.A., Shkrob, I. A. (1993) Radical pair phenomena
in exiplex and exited radical systems, Z.Phys. Chem. 180, 9-31.
10. Okazaki, M., Sakata, S.; Konaka, R., Shiga, T. (1987) Product yield detected ESR
on magnetic field dependent photoreaction of quinones in sodium dodecylsulphate
micellar solution, J.Chem.Phys. 86,6792-6800.
11. Polyakov, N.E., Okazaki, M., Toriyama, K., Leshina, T.V., Fujiwara, Y.,
Tanimoto, Y. (1994) Product-Yield-Detected ESR Study on the Dynamic Behavior
of Radical Pairs Generated in Photoreduction of Acetylenic Ketones in SDS
Micellar Solution, J.Phys.Chem. 98, 10563-10567.
12. Bagryanskaya, E.G., Sagdeev, R.Z. (1993) Novel aspects ofDNP and SNP, Zfur
Phys. Chem.180, 111-114.
13. Tarasov, V.F., Bagryanskaya, E.G., Shkrob, LA., Avdievich, N.I., Ghatlia, N.D.,
Lukzen, N.N., Turro, N.J., Sagdeev, R.Z. (1995) Examination of the exchange
interaction through micellar size. 3. Stimulated nuclear polarization and time
resolved electron spin resonance spectra from the photolysis of methyl
deoxybenzoin in alkyl sulfate micelles of different sizes, J.Am. Chem.Soc. 117,
110-118.
14. Bagryanskaya, E.G., Tarasov, V.F., Avdievich, N.L, Shkrob, LA. (1992) Electron
spin exchange in micellized radical pairs. III. \3C low-field radio frequency
stimulated nuclear polarization spectroscopy (LF SNP), Chem.Phys. 162,213-223.
15. Bagryanskaya, E.G., Sagdeev, R.Z. (1993) Kinetic and mechanistic aspects of
stimulated nuclear polarization, Prog.Reaction Kinet. 18,63-124.
16. Pamachev, A.P., Bagryanskaya, E.G., Sagdeev, R.Z. (1997) A study of
benzophenone photolysis in SDS micelles in the presence of 2,4,6-tri-tert-
butylphenol: distinctive features of SNP in radical pairs with a natural abundance
of \3C nuclei, J.Phys.Chem.A 101,3854-3862.
17. Ananchenko, G:S., Bagryanskaya,E.G., Tarasov, V.F., Sagdeev, R.Z., Paul, H.
(1996) A 3I p _SNP study of photolysis of (2,4,6-
184

trimethylbenzoyl)diphenylphosphine oxide in micelles of different size,


Chem. Phys.Lett. 255, 267-273.
18. Lebedeva, N.V., Bagryanskaya, E.G., Gorelik, V.R., Koptyug, LV., Sagdeev, R.Z.
(2001) Temperature and Salt Additions Effect on the Micellised Radical Pairs
recombination studied by Stimulated Nuclear Polarization, J.Phys.Chem.A 105,
4640-4647.
19. Tarasov, V.F.; Buchachenko, AL.; Maltzev, V.L (1981) Magnetic isotop effect
and the separation ofisotop in "microreactor", Russ. J.Phys.Chem. 55, 1921-1928.
20. Paul, H. (1975) Electron spin relaxation of HCO· in liquids, Chem. Phys. Lett.
160, 472-475.
21. Koptyug, LV., Lukzen, N.N., Bagryanskaya, E.G., Doktorov, A.B., Sagdeev, R.Z.
(1992) The Influence of the Singlet Radical-Pair Decay on RYDMR and SNP
Spectra, and the Mean RP lifetime, Chem.Phys. 162,165-169.
22. Steiner, V.E., Wu, J.R. (1992) Electron spin relaxation of photochemically
generated radical pairs diffusing in micellar supercages, Chem.Phys. 162,53-68.
23 . Pedersen, J.B., Freed, J.H. (1974) Some theoretical aspects of chemically induced
dynamic nuclear polarization, J.Chem.Phys. 58,2746-2762.
24. Avdievich, N.L, Bagryanaksya, E.G., Tarasov, V.F., Sagdeev, R.Z. (1993)
Investigation of micellized radical pairs in the photolysis of ketones by time-
resolved stimulated nuclear polarization, Z.Phys. Chem.Bd. 182, 107-117.
25. Parnachev, AP., Bagryanskaya, E.G., Tarasov, V.F., Lukzen, N.N., Sagdeev, R.Z.
(1995) Investigation of chemical and spin dynamic in micellized radical pairs by
time-resolved stimulated nucler polarization. Theory and experiment,
Chem.Phys.Lett. 244, 245-251.
26. Parnachev, AP., Purtov, P.A., Bagryanskaya, E.G., Sagdeev, R.Z. (1997)
Theoretical and experimental studies of CIDNP kinetics in recombination of
radical pairs by the method of switched external magnetic field, J.Chem .Phys. 107,
9942-9954.
27. Fedin, M.V., Bagryanskaya, E.G., Purtov, P.A. (1999) Theoretical and
experimental studies of chemically induced dynamic nuclear polarization kinetics
in recombination of radical pairs by the method of switched external magnetic
field. II. I3C CIDNP ofmicellized radical pairs, J. Chem.Phys. 111, 5491-5502.
28 . Turro, N.J. and Cherry, W.R. (1978) Magnetic isotope and magnetic field effects
on chemical reactioIls. Sunlight and soap for the efficient separation of C I2 and CI3
isotopesJ.Am.Chem.Soc. 100,7431-7434.
29. Emert, J., Behrens, c., Goldenberg, M. (1979) Intramolecular eximer-forming
probes of aqueous micelles, J.Am.Chem.Soc. 101,771-772.
30. Wasserman, A, Tarasov, V.F. (2001) Chern. Phys. in press.
31. Evans, C.H., Scaiano, lC. and Ingold, K.V. (1992) Influence of micellar size on
the decay of triplet-derived radical pairs in micelles, J.Am.Chem.Soc. 114,140-143.
32. Tsentalovich, Y.P., Fisher, V . (1994) Solvent Effect on the decarbonylation of
acyl radical studied by laser flash photolysis, J.Chem.Soc.PerkinTrans.2 , 729-734.
MONITORING ULTRASLOW MOTIONS IN ORGANISED LIQUIDS

NMR Experiments and computer simulations

F.A. GRINBERG
University of Ulm
Albert-Einstein-Allee, J J
89069, Ulm, Germany

Introduction

The interest to liquid crystals confined in porous materials has greatly increased in
recent years due to their important applications in new electro-optic technologies
[1]. Confinements by pores and surface interactions produce strong effects on the
properties of liquid crystalline systems both below and above the bulk clearing
point [2, 3]. The aim of this work is the study of a surface induced orientational
anisotropy of constrained liquid crystals above the nematic-isotropic transition
temperature on the basis of the stimulated echo experiments and the computer
simulations.
Surface effects are strongest in the near vicinity of the wall. In this work our
attention will be focused on the materials with pore sizes of a few nanometer much
less studied than the materials with relatively big pores [I].
In liquids, the modulations of dipolar interaction by stochastic molecular
motions determine the most important mechanisms of relaxation. We present the
computer simulations of random walks constrained by uneven surface aiming at a
numerical evaluation of the correlation function of spherical harmonics
proportional to a reduced dipolar auto-correlation function.

1. Materials and Instruments

The material studied was 4'-n-pentyl-4-cyanobiphenyl (5CB) confined in porous


silica glasses Bioran (mean pore radius R== 5 nm), Vycor (Vyc, R == 2 nm), and
controlled porous glasses (CPG-1.5, R = 1.5 nm; CPG-4, R = 4 nm). The bulk
isotropization temperature (TNI) of 5CB is 309.6 K. The attenuation curves of the
transverse proton magnetization were recorded at 90 and 400 MHz using high-
power Bruker instruments. The transverse proton relaxation rates, T2'\, were
measured with the standard Hahn echo pulse sequence (see, for instance, [4]). The
"spin exchange" (cross-relaxation) rates [5, 6], 'tex'), were measured at 90 MHz
using the stimulated echo pulse sequence.
185
1. Fraissard and o. Lapina (ells.), Magnetic Resonance in Colloid and Interface Science. 185-196.
© 2002 Kluwer Academic Publishers.
186

2. Computer Simulations

Diffusion of the molecule within the sphere of a given radius, R, was simulated
using a "random walk" procedure. Within each elementary time cycle, Llt, the
random walker performs a step of the elementary length LlI «R. Spatial positions
of the walker are determined in cartesian co-ordinate system with the origin in the
center of the sphere. The directions of the movement along each of the axes are
determined independently using random-number generators. All sites spanned by
the sphere can be visited. Any sites outside the sphere, that is, with the radial
l
distances from the center r == ) x 2 + + Z2 > R + M , are prohibited.
Within the sphere, the random walker can be found in either of two possible
states: a) it is "inside the sphere" (or "in bulk" ) if the co-ordinates obey
r<R-iSr, or b) it is "in the surface layer" if R-iSr~r~R+M. The
parameter iSr stands here for the thickness of the layer in which the molecules are
supposed to be oriented by the surface. The random walker is additionally
characterized by the probability, Ws-h of performing the next step which always
equals a unity "inside the sphere" and might have smaller values if
R =r ~ R + M . In this work, we do not attribute any specific meaning to the
quantity Ws_1 other than it causes the walker to "hesitate" between the steps if it's
value is less than 1. The connection of W5-1 to the parameters of the adsorption-
desorption kinetics, for instance, can be found in [7).
As long as the particle is "inside the sphere", the value of the spherical
harmonics (for definition see, for instance, [4]), Y2k(9 (t), <pet»~, is set to zero
assuming this to be the "net" result of any fast reorientations typically responsible
for motional averaging in liquids. In this limit, the "bulk-like" relaxation rate, that
is, the relaxation rate due to any relaxation mechanisms in the interior part of the
confining surface, is assumed to be zero.
On the other hand, if the particle is "in the surface layer", the current value of
Y2k (9 (t), <pet»~ is determined by the preferential orientation of the molecule relative
to the surface at its momentaneous position. We assume that the inter-nuclear
vector is oriented along the radius-vector connecting the current position with the
center of the surface. In other words, the inter-nuclear vector is supposed to be
oriented perpendicular to the surface (taken the nearest to its momentaneous
position surfa,ce site). The magnetic field is aligned along the z-axis as usual. The
spherical harmonics are calculated with polar angles relative to z direction. All
time and length scales are evaluated using Lli and the diffusivity, D, given In
advance.
The reduced correlation function evaluated after n steps is defined by

G
k
(t) = (I;kCO)I;kCt))
(I;~ ) ' k =0,1,2
(1)
187
where

(2)

and

This numerical algorithm addresses a reduced dipolar auto-correlation function


for diffusion between the states with zero and finite values of the dipolar coupling
constant. Note that it is not identical to a two-site exchange model, where just the
two sites with different features (two different local orientations in this case)
would be considered. Rather we are dealing here with the process where the
confining structure is probed in a series of visits at numerous sites. As molecules
diffuse between the disoriented interior pore space and the surface sites with
different orientations relative to the external magnetic field, the residual dipolar
interactions undergo (slow) fluctuations owing to repeated returns of molecules to
the surface. The corresponding relaxation mechanism is well known as
"reorientations mediated by translational displacements" (RMTD) [4]. It is
responsible for a strong enhancement of the relaxation in the low frequency (kHz)
range in polar liquids confined in pores [8, 9]. It was also found to be an extremely
important relaxation mechanism in confined liquid crystals [2, 3].
Figures 1 and 2 demonstrate typical simulated correlation functions, Eq. (1).
Only the curves for k == 0 are shown in the plot since the deviations of the results
for different values of k (0, 1 and 2) were limited to a few percents. The
parameters used were R == 20 nm, ~ 1== 0.2 nm and D==1 x 10- 10 m2/s (Fig. 1) and
Ws-I == 0.01 (Fig. 2). The correlation functions in Figs 1 and 2 clearly reveal two
characteristic time scales of the correlation losses with the relative intensities of
the fast and the slow attenuation components not depending on D but strongly
depending on Ws_I . Figures 1, 2 show that the slowest component decays
exponentially. This component is of interest in this context since it is responsible
for the most low frequency part of the NMR relaxation. The time constant of the
exponential tail of the curve (further referred as the correlation time) was
determined using the least square fits.
Figure 3 shows the correlation times as a function of the diffusivity. These
quantities appear to be inversely proportional to each other as expected for a
diffusion restricted to a closed cavity. At the same time, a "hesitation" of the
molecule on the surface between the subsequent steps (Ws-I < I) leads to a strong
slowing down of the attenuation of the correlation (increase of the correlation
time).
188

3. Stimulated echo and "spin exchange"

The "exchange" (or cross-relaxation) rates were determined using one-dimensional


stimulated and primary echo experiment. The basic pulse sequence applied is 90°-
't1-900-'t2-90o.
Generally, factors responsible for the attenuation and the modulation of the
amplitudes of the primary, A(2'tl), and the stimulated echo, A(2'tl+'t2), are given by
the contributions due to the transverse and the longitudinal relaxation, the dipolar
correlation effect, translational diffusion and the contribution due to local
resonance offsets (see for detail our earlier works, for instance [10 - 12] and
references therein). The attenuation due these mechanisms, with the exception of
the very last one, are essentially irreversible. The attenuation due to local
resonance on the contrary can completely be reversed given the quasi-static
conditions, i. e. when. spins experience the same local resonance offsets in each of
the relevant evolution intervals.
As it was already shown in [12], the stimulated echo of 5CB confined in small
cavities is subject to a modulation by "spin exchange" (cross-relaxation [5, 6])
between in-equivalent (due to chemical shifts) proton spins. The effect of this
mechanism is that the stimulated echo amplitude as a function of'tl becomes
modulated with the difference frequency of the chemical shifts. The same
mechanism produces the cross-peaks in a well-known two-dimensional NOESY
NMR experiment [6] on the basis of the nuclear Overhauser effect.
In the frame of a "two-site" ~xchange model, the modulation by "spin
exchange" of the quotient of the stimulated and the primary echo amplitudes is
given by [12], see also Appendix:
A(2'1 +'2) = 1- 2P P. (l-cos(2dW, ))(I_e-r2/rex) (4)
A(2'1) a h 1

where 2~ (t) is the total frequency (chemical shift) difference b~tween two spin
states, Pa and Pb are the relative spin fractions in these states (P a + Pb =1), and
'tex - 1 is the exchange rate. The quotient of echo amplitudes in Eq. (4) is constructed
in order to eliminate the attenuation of echo amplitudes by the transverse
relaxation.
We have measured the exchange rates of 5CB confined in porous glasses with
four different pore sizes. The modulated attenuation curves measured at 313 K are
shown in Fig. 4 for CPG-4 as a representative sample. In Fig. 4, the stimulated
echo amplitudes are divided by the amplitudes of the primary echo which were free
of any modulations of this sort. The different sensitivity of the primary echo, with
the maximal amplitude at 2't\, and the stimulated one, with the maximal amplitude
at 2't,+'t2, to "spin exchange" in the investigated systems is due to the differences
in evolution time of the both echoes. The modulating mechanism appears to be
rather slow on the time scale of the transverse relaxation which limits the range of
189

measured tl-intervals for the primary echo down to a few milliseconds. This
process however becomes efficient on the time scale of the order of 10- 1 s
(observe increasing amplitude of the modulations with increasingt2 in Fig. 4) and
can be probed by measuring the stimulated echo for large enough intervals t2. The
latter is limited by longitudinal relaxation which is of the same order of the
magnitude as the characteristic time scale of the exchange process but much longer
than the transverse relaxation.
The fits of Eq. (4) to the experimental data are shown in Fig. 5 by the solid
lines. The only fitted parameters were t ex - I and 2~0). The fractions Pa and Pb were
obtained directly from the measured spectra as the relative intensities of the
spectral lines. The fitted values of t ex - I as a function of temperature are shown in
Fig. 5. The fitted values of 2~0) were in a good agreement with those measured
from the splitting of the spectra. In Vycor sample, we also performed the two-
dimensional NOESY experiments [6] in order to compare the relaxation rates
obtained by both methods. The data points from NOESY experiments are shown
by the stars. The results of both methods appear to be in a very good agreement.
The stimulated echo experiments however have the advantage - given a single
exchange process as it was the case here - of being less time consuming and easier
to perform.
Figure 5 shows the (inverse) temperature dependences of the exchange rates. In
samples with R < S nm, t ex - I smoothly increases as temperature decreases. The
smooth behavior remains intact also at cooling below TNI (indicated by the vertical
dashed line). In Bioran glass with R = S nm, the. exchange rate of SCB on the
contrary changes jump-like in the vicinity of TNT.
Figure 6 shows the square roots of the exchange rates as a function of the mean
pore radiuses of the investigated samples at 310 K. The data point for CPG-4 is
obtained from the extrapolation of the measured temperature dependence to higher
temperature values (as shown by dashed line in Fig. 5). All values in Fig. 6 are
normalized at the smallest pore radius. Three first data points at small R can well
be fitted by an exponential function (as shown in semi-logarithmic scale), whereas
the data point of the Bioran glass deviates essentially from that behavior. Similar
dependence on the pore radius was earlier [13] observed for the square roots of the
transverse relaxation rates in the same samples. The normalized square roots of the
transverse relaxation rates measured 313 K are shown for a comparison in the same
plot. One can see that both data sets are in a good agreement.

4. Discussion and Conclusions

The amplitudes of the stimulated echo of SCB confined in small cavities (R<5 nm)
are modulated by "spin exchange" process efficient on the time scale of't2
intervals. This phenomenon is observed well above the bulk clearing point and
persists also as the temperature goes down at least 10 K below TNI. This relaxation
190

mechanism was not observed for a bulk counterpart neither above nor below TN!
[10, 12]. The relaxation mechanisms governing the attenuation of the quotient of
the stimulated and primary echo amplitudes in seB above TN! are the same as in
isotropic liquids, that is, the quotient of the stimulated and primary echoes is
constant if the attenuation due to diffusion may be neglected. Below TNh the
dominated mechanism was shown [10] to be due to the dipolar-correlation effect.
The latter also dominates in seB confined to bigger cavities (R > 30 nm).
Earlier [13], we observed the enhancement of the transverse relaxation rates
well above TN! in the same samples when comparing to bulk SeB. We assumed
that the enhancement mechanism is due to a surface induced orientational
anisotropy of liquid crystalline molecules as suggested in works by M. Vilfan et al.
(see for instance Ref. [2] ).
The orientational anisotropy stabilized by a solid interface prevents a complete
averaging of dipolar interactions by molecular motions. Diffusion between the
regions with different surface orientations and different levels of orientational
anisotropy then produces a strong relaxation mechanism in the low frequency
range (KHz) relevant for the transverse relaxation. Figure 6 shows that "spin
exchange" studied in this paper is governed by the same dominant mechanism as
the transverse relaxation (observe a nearly identical behavior of the
relaxation/exchange rates as a function ofR).
The decrease of relaxation rates with increasing the pore size is to be attributed
to the averaging effect due to translational molecular diffusion [4, 14] between the
regions with different degrees of local anisotropies. These must decrease with
increasing the distance, r, from the wall as [2]

S(r) oc exp ( -;) (5)

where S is a local order parameter, ~ is a characteristic decay constant.


The relaxation rates are generally proportional to the square of the (residual)
dipolar couplings. In confined liquid crystals above TNh the square roots of the
relaxation rates then appear [2] to be proportional to the orientational local order
parameter, S, induced by a surface. The exponential decay of ~T:xl with
increasing R (first three data points) shown in Fig. 6 thus reflects the change of
the average S with increasing R. From the fits of the exponential function to the
square roots of the exchange rates in Fig. 6 (for R < S nm) we obtained the value
of 3.4 nm. It is very close to the value fitted to the square roots of the transverse
relaxation rates, 2.67 nm. These values give the estimation for the correlation
length of the surface induced orientational anisotropy. The conclusion is that the
exchange spectroscopy/relaxometry can be effectively used along with the
transverse NMR relaxation for the quantitatve study of a surface induced ordering
in confined liquid crystals above the isotropization temperature. This method can
further provide a tool for the determination of the mean pore sizes of porous
materials with small cavities (below S nm).
191

5. Appendix

In order to estimate the exchange rates, the experimental data were treated
according to a simple two-site ("a" and "b") exchange model. Each spin is assumed
to have a certain probability per time unit, ta(b)-l, to leave its state "a"(or "b"). The
states are characterized by the precession frequencies -~O) and +~O) in the rotating
frame (the total chemical shift difference being 2~O) and the relative spin
fractions, Pa and Pb (Pa+Pb=l). Since the typical situation in our experiments was
tl«t2, a given spin is assumed not to leave its initial state during the time t=tl .
The contribution of the phase shifts ~i due to one particle local field offsets gained
in the first (from 0 to tl) and the third (from tl+t2 to 2tl+t2) time intervals to the
amplitude ofthe stimulated echo can be generally expressed as
A(2TI +T2 )=(COS(lP3 -lPI)) + (COS(lP3 +lPI))' (6)
where the brackets denote the ensemble average. With the above assumptions, the
procedure of averaging over the stochastic process of "spin exchange" and over the
field gradients (assumed to be constant within the sample) leads to:
A(2TI +T2 ):::: PaWaa +~~b + (P"Wab + ~~a)cOs(2~mTI)' (7)
where Wij are the conditional probabilities for a given spin to occur at time t=tl+t2
in state j, if at t=0 it was found in state i . These probabilities obey the following
differential equations:

(8)

aWhb ___ 1 W. ~w. . aWhll = __1 W. +_1 W. (9)


~t - hb+ bll' ~t hll bb·
v ~ ~ v ~ ~

The solution of Eqs. (8, 9) with the initial conditions Waa=Wbb=l and Wab=Wba=O

:J
gives:

W~ =p.+p.ex+ (10)

~b = I1. + ~ exp( __t J (11)


Tex

where T:xl == T~I + T;I. With account of Eqs. (10, II), Eq. (7) reduces to:
192

1~======~======~====~
Ws-I
0 .005

0 .1

0. 0 5.0x10-7 1.0x1 0-6


t, S

Figure 1. Correlation functions, Eq. (1) obtained in the computer simulations for a
sphere with R = 20 nm. The values of ~I and 0 were 0.2 nm and 10-10 m2/s,
respectively. The values of Ws-1 as a curve parameter are indicated in the plot.

0.1

0.0 1.0x10-6 1.5x10 -6


t, S

Figure 2. Correlation functions, Eq. (1), obtained in the computer simulations for
a sphere with R = 20 nm for different values of D indicated in the plot. The values
of ~I, or and W s•1 were 0.2 nin, 2 nm and 1.0, respectively.
193

't c ' S
W5· 1
1x10-4
• 1.0
1x10-5 • 0.01

10~

10.7

10~

10.11 1x10·10 1x10·9

Figure 3. Values of 'tc fitted to the slowly attenuating parts of the simulated
correlation functions as a function of the diffusivity.

A(21: 1 +1( 2 ) 1.6 T2 , s t2, s


1.4 0.010 0.080
0.020 0.160
A(2 1:1 ) 1.2 0.040 0.320

1.0
~.
\
0.8

0.6
0.4

0.2
O. 0 L-_"--_-'--_....o...-_....I....---,--'-_......L...---,-~_ _l

0.0 20. x10 ·3 4.0x10 ·3 6.0x10 ·3


't 1 , S

Figure 4. Attenuation curves of the amplitudes of the stimulated echo divided by


the amplitudes of the primary echo as a function oh l . The curve parameter was 't2
as indicated in the plot. The solid lines are fits of the Eq. (4). The fitted parameters
were 'tex• 1 and 2Llco. The temperature dependence of 'tex• 1 is shown in Fig. 5.
194

-1 -1
'rex I S

-, ... -
CPG~
Vyc2 nm
,,' '
I

,,
... ,
Bioran 5 nm ..
0.0024 0.0027 0.0033 0.0036

Figure 5. Temperature dependences of the exchange rates, tex-t, in 5CB confined


in porous glasses with four different pore sizes.

1
E
c: E
c:
,...
It? It?
,...
II II
~ ~
d~ ~
o~

. 0.1 *"

0;-
N'

l- ~

:::::- t:
on on

.
ci ci~
(T -1 )O.5 f (T -1) o.5(R=1.5 nm)
0;-

!::..
0;-

~
.... G
*• (t
ex
2
.1)O.5 ft
ex
2
.1)o.5(R=1.5 nm)

0.01

R, nm
Figure 6. Square roots oftex-1 and ofT2- 1 as a function of R measured at 310 K and
313 K, respectively. All values are normalized at the value of the smallest R = 1.5
nm. The solid line is the fit of the exponential function to the square roots of the
transverse relaxation rates for R < 5 nm. The characteristic length constant was
2.76 nm. The dashed line is the fit of the exponential function to the square roots
of the exchange rates. The characteristic length constant was 3.14 nm.
195

Acknowledgements. I would like to thank the Deutsche Forschungsgemeinschaft


and the Minesterium flir Wissenschaft, Forschung und Kunst Baden-Wurttemberg
for financial support. Further I would like to express my special thanks to
Prof. R. Kimmich and Prof. M. Vilfan for valuable discussions. I cordially thank
Prof. M. Mackowiak for his help in course ofNOESY- experiments.

6. References

1. Crawford, G. P., Zumer, S. (1996) (Eds.) Liquid Crystals in Complex


Geometries, Taylor&Francis, London.
2. Vilfan, M., Vrbancic-Kopac, N., Ziherl, P., Crawford, G. P. (1999) Deuteron
NMR relaxometry applied to confined liquid crystals, Appl. Magn. Reson. 17,
329-344, and references therein.
3. Zumer,S., Ziherl, P. and Vilfan, M. (1997) Dynamics of microconfined
nematic liquid crystals and related NMR studies, Mol. Cryst. Liq. Cryst. 292,
39-59.
4. Kimmich, R. (1997) NMR: Tomography, Difjusometry, Relaxometry, Springer-
Verlag, Heidelberg.
5. Neuhaus, D. and Williamson, M. (1989) The Nuclear Overhauser Effect in
Structural and Conformational Analysis, VCH Publishers, New York.
6. Ernst, R. R., Bodenhausen, G. and Wokaun, A. (1987) Principles of nuclear
magnetic resonance in one and two dimensions, Clarendon Press, Oxford.
7. Valiullin, R., Kimmich, R., Fatkullin, N. (1997) Levi walks of strong
adsorbates on the surface: Computer simulation and spin-lattice relaxation,
Phys. Rev. E 56, 4371-4375.
8. Zavada, T., Kimmich, R. (1998) The anomalous adsorbate dynamics at
surfaces in porous media studied by nuclear magnetic resonance methods. The
orientational structure factor and Levi walks, J Chem. Phys. 109, 6929-6939.
9. Zavada, T., Stidland, N., Kimmich, R. and Nonnenmacher, T. F. (1999)
Propagator representation of anomalous diffusion: The orientational structure
factor formalism in NMR, Phys. Rev. E,60, 1292-1298.
10. Grinberg, F., Kimmich, R. (1996) Pore size dependence of the dipolar-
correlation effect on the stimulated echo in liquid crystals confined in porous
glass, J Chem. Phys. 105,3301-3306.
11. Grinberg, F., Garbarczyk, M. and Kuhn, W. (1999) Influence of the cross-link
density and the filler content on segment dynamics in dry and swollen natural
rubber studied by the NMR dipolar-correlation effect, 1. Chern. Phys., Ill,
11222-11231.
12. Grinberg, F., Kimmich, R. and Stapf, S. (1996) Investigation of molecular
order and dynamics in liquid crystals confined in porous media using the
dipolar-correlation effect on the stimulated echo, Magn. Reson. Imaging, 14,
883-885.
196
13. Grinberg, F., Kimmich R. (2001) Surface effects and dipolar correlations of
confined and constrained liquids investigated by NMR relaxation experiments
and computer simulations, Magn. Reson. Imaging, in press.
14. Karger, J., Heitjans, P. and Haberlandt, R. (Eds.) (1998) Diffusion in
Condensed Matter, Vieweg&Sohn, Wiesbaden.
CHARACTERIZA TION OF MASS TRANSPORT AND RELATED PHENOMENA
IN POROUS CATALYSTS AND SORBENTS BY NMR IMAGING AND
DISPLACEMENT NMR SPECTROSCOPY

LV. KOPTYUG, L.YU. ILYINA, A.V. MATVEEV, R.Z. SAGDEEV,


V.N. PARMON!
International Tomography Center
3A Institutskaya St, Novosibirsk 630090, Russia, and
JBoreskov Institute of Catalysis
5 Acad. Lavrentiev Pr., Novosibirsk 630090, Russia

1. Abstract

Applications of NMR imaging in heterogeneous catalysis are rapidly gaining popularity.


Due to its nondestructive and noninvasive nature, the technique can be successfully
employed to characterize many of the essential stages of catalyst preparation and use,
including multiple coupled processes. The dynamics of Pt or Pd redistribution and the type
of their ultimate macroscopic distributions in supported Ptlalumina (Pdlalumina) catalysts
can be mapped indirectly by !H NMR imaging. Flow imaging of liquids and gases provides
flow velocity maps and can be used to guide the design of shaped structures such as
monolithic catalysts. Displacement NMR spectroscopy yields information on the average
flow velocities and hydrodynamic dispersion for gases, liquids and solids flowing through
porous media. Imaging of drying of porous pellets and beds comprised of such pellets
initially saturated with various liquids, as well as imaging of water vapor adsorption by
porous pellets and beds, can provide useful and reliable infonnation on mass transport
mechanisms within porous media and interphase mass transfer between the catalyst and the
flowing gas. Spatially resolved imaging of gas adsorption on supported catalysts can be
employed to study the selectivity of adsorption by active components. Finally, NMR
imaging can be applied to monitor the progress of heterogeneous catalytic reactions in real
time. The examples of such applications are presented and discussed.

2. Introduction

NMR has become a versatile tool with a very broad range of applications in numerous areas
of research. Both NMR spectroscopy and imaging are widely and successfully employed in
materials science and chemical engineering. NMR is becoming a promising nondestructive
and noninvasive tool for heterogeneous catalysis which can be successfully employed to
characterize many of the essential stages of catalyst preparation and use. It is now possible
to use NMR not only to address certain isolated aspects of the complex problems
encountered in chemical engineering and catalysis, but also for a comprehensive study of a
broad range of interrelated issues. In fact, in many cases it should be possible to treat a
197
198

problem at hand and the way to solve it as a pyramid, addressing simpler sub-problems first
and then building the higher levels based on the results obtained at the previous stages. In
this report we intend to demonstrate that NMR is mature enough to be employed in such an
approach to the investigation of an operating reactor, by going through several levels of
increasing sophistication, from studying simple uncoupled processes and properties through
the investigation of complex processes toward addressing a coupled transport-adsorption-
reaction system.

3. Experimental

All IH NMR microimaging experiments were performed at 299 MHz on a Bruker DRX
spectrometer equipped with a vertical bore superconducting magnet and the micro imaging
accessory. Distilled water, "purum" grade cyclohexane, and acetylene and butane gases
were used. GC MS has shown the presence of other hydrocarbon gases (a few %) in butane.
Texture characteristics of alumina (y-AI 20 3) and silica gel pellets used in the
experiments were obtained by conventional techniques. Monoliths were made of y-Ah03
and had 14 nm average pore diameter, transport channels of (4mm)2 square cross-section
and wall thickness of 1 mm. The roughly cylindrical pieces ca. 21 mm in diameter were cut
out of bigger monoliths, with the cylinder axis parallel to the transport channels. Some of
the drying studies were done with a monolith which was cut into 10 equal sections
perpendicular to the direction of the channels and then glued back together with epoxy to
restore the original sample shape.
For the mapping of active components distribution in the supported catalysts, the
following experimental procedures were employed. All pellets were initially saturated with
water. The egg-shell distribution of Pt was obtained by diffusional impregnation in a 0.02N
aqueous solution of H2PtCI 6 • For an egg-white distribution, the pellets were treated for 10
min with 0.3 N oxalic acid aqueous solution, and then placed in the 0.02 N HzPtCl 6 + 0.3 N
H1C20 4 aqueous solution. Similar procedure was used for Pd. For the studies of the mixed
Cu+Pt supported catalysts, the pellets were placed for two hours in a 2.5x 10- 2 M aqueous
solution of CU(N03)2, or a 0.01 N solution of H2 PtCI 6, or their mixture. After the adsorption
stage, all samples were dried for several hours at 130° C, and then saturated with
cyclohexane. The 2D spin-echo sequence with 3 mm slice selection and inversion-recovery
preconditioning was employed. The detection of each 2D image took 15-30 min. A set of
16 images with various degrees of T I-weighting was obtained, and the spatial maps of the
cyclohexane spin density and the spin-lattice relaxation time were reconstructed with the
in-plane resolution of (78 !lml
For the imaging of the dynamics of the impregnation process in real time, alumina pellet
was treated with 0.3 N H2C20 4 aqueous solution for 10 min, then transferred to the NMR
tube containing 5 ml of 0.02 N HzPtCl6 + 0.3 N H 1C10 4 aqueous solution, and the imaging
experiment was initiated immediately. In this case the nuclear spin relaxation times of
water hydrogens are sensitive to the presence ofhexachloroplatinate dianion. A single value
for the recovery delay which gave the best contrast in the TI-weighted images (256 ms) was
used. Several 2D water spin-echo images were acquired sequentially after the initiation of
the impregnation process; the acquisition time of each 2D image was 9 min 40 s.
For single pellet drying experiments, alumina pellets initially fully saturated with water
were used. For single pellet adsorption experiments, alumina or silica gel pellets were
impregnated with CaCh (or CuCh). Depending on the desired distribution of the salt, the
199

pellets were initially fully saturated with water or partially saturated with cycIohexane, then
placed in a saturated aqueous solution of CaCI 2 for several hours and later dried at l30oC.
All pellets were cylindrical, 3.5-5 mm in diameter. During the experiments, the two ends of
the cylindrical sample were covered to ensure radial mass transport. A stream of room
temperature air or nitrogen gas of a desired relative humidity (ca. 0% for drying, 32-100%
for adsorption) was passing along the pellet surface in the axial direction. One-dimensional
water concentration profiles along the diametrical slab of the cylindrical pellet with spatial
resolution of 73 Ilm were detected employing a 10 spin-echo pulse sequence with slice
selection in two orthogonal directions. Each averaged profile was accumulated within 35-
40 s.
The silica gel pellets used in the bed adsorption experiments were roughly spherical, 1-5
mm in diameter. In the bed adsorption and drying experiments, the porous pellets were
packed in a glass cylinder of 21 mm internal diameter, and the flow of dry or wet air was
passing through the bed. The 2D images in the plane perpendicular to the cylinder axis in
the bed drying experiments were obtained by conventional spin-echo method. The 1D
projections of signal intensity on the axis of the cylindrical bed with no slice selection were
detected in the bed adsorption experiments by either spin-echo or the single point imaging
(SPI) methods.
For 2D visualization of the redistribution of CuC/z, a slice-selective spin-echo sequence
with the repetition time of 0.15-1.5 s was used. The pellet was dried after each sorption
experiment, saturated with cycIohexane and imaged, and then dried again before the next
water sorption experiment.
For water vapor adsorption studies in the absence of air, alumina or silica gel pellets
containing CaCh were evacuated down to 2.5 Pa residual air pressure. After positioning the
cell in the NMR probe, the sample was connected to a flask containing degassed water or a
saturated aqueous solution of CaC/z, Mg(N03h or NaCI. Water vapor sorption through one
or both flat edges of the pellet was studied. AID SPI method was used to detect I D water
content profiles along the pellet axis with spatial resolution of 206 Ilm or 365 Ilm and the
total accumulation time per profile of 130 s.
Monoliths were saturated with water and dried with the stream of dry air passing
through the transport channels. The ID SPI projections of the water content on the direction
of the channel axes were detected every 130 s with spatial resolution of 417 Ilm.
In the gas flow experiments, the gas was expanding from the tank into the supply tubing
attached to the bottom part of the cell, and was leaving into the fume hood through the open
end of the exhaust tubing attached to the top of the cell. A thermostat pump was used in the
water flow experiments.
The 2D spin-echo pulse sequence with a slice selective 90°-pulse and a hard 180°-pulse
was employed in all liquid and gas flow imaging experiments, with the slice thickness of 2
mm for water and 4 mm or 15 mm for gas studies. The echo time was 5.1 ms (gas) or 2.7
ms (water). The true in-plane spatial resolution of (400 Ilmi for gases and (200 Ilm)2 for
water was improved to (100 Ilm)2 by zero-filling. Two extra gradients were applied to
phase encode the flow velocity. The velocity maps were constructed from two 2D images
detected with and without flow. It took 20 to 40 minutes to acquire each 2D image. PFG
NMR experiments were performed using the stimulated echo sequence with an extra pair of
gradients for encoding of the displacement along a certain direction with or without slice
selection.
Studies of the gravity driven granular flow utilized 100-200 Ilm porous alumina grains
partially saturated with vaseline oil (8.5 weight %) flowing through a cylindrical bed with
200

23 mm diameter comprised of 5 mm alumina beads. The transverse 2D maps of the axial


velocity component were detected using the stimulated echo based imaging pulse sequence
with (0,4 mm)2 in-plane resolution, 5.3 mm slice thickness and 6 min 40 s acquisition time.
PGSTE sequence was employed to detect the average propagator.
The 2D images of the gas filling the reactor and adsorbed within the Ptlalumina catalyst
grain were detected with the conventional spin-echo technique. In-plane resolution was (0,4
mm)2 and acquisition time was 17 min. No slice selection was used.

4. Results and discussion

NMR spectroscopy and imaging techniques are widely employed for the characterization of
porous materials, and can provide information about their homogeneity, pore size
distribution and connectivity of the pore space. Several techniques exist for measuring pore
size distributions, such as NMR cryoporometry [1] and surface induced spin relaxation
measurements [2,3]. Besides, our results have shown that the information about pore size
distribution can in principle be obtained form the drying studies of porous objects initially
saturated with water [4,5]. In this case, pore size distribution has a pronounced influence on
the capillary transport of liquid within the pore space under the developing concentration
gradients, and thus on the temporal transformations of liquid distribution profiles in the
course of drying. Moreover, the drying experiment can reveal the non-uniformity of the
pellet in terms of the variation of pore sizes, since such variation across the pellet leads to
the distortion of an otherwise smootn transient distribution profile.
NMR imaging c!an be successfully employed for the characterization of supported
catalysts. If a support is impregnated with a paramagnetic active component, the relaxation
times of a liquid permeating· porous support will be reduced in the impregnated
macroscopic parts of the support. We have employed this approach to confirm the egg-shell
type of CuCIz distribution in an alumina pellet [6,7], Furthermore, diamagnetic active
components are also known to alter the relaxation of liquids permeating porous materials.
We have demonstrated that the nuclear spin relaxation times of liquids permeating the pores
increase in the regions of porous alumina pellets impregnated with Pt or Pd [8-10]. In
general, the relaxation times of liquids permeating alumina and other types of pellets are
considerably reduced as compared to bulk liquids due to the interactions with the pore walls
and due to the presence of paramagnetic impurities (e.g., iron) within the walls.
Presumably, adsorption of PtCI/ or PdCl/ on the pore surface increases the average
distance between the molecules of the liquid and paramagnetic impurities thereby reducing
the paramagnetic contribution to nuclear spin relaxation [11]. This effect can be employed
to study macroscopic distribution of active components such as Pt and Pd in supported
catalysts. Furthermore, the dynamics of their redistribution can be visualized in real time in
the course of supported catalyst preparation [8,10]. Since diamagnetic and paramagnetic
active components change the relaxation times of liquids in the opposite directions, it is
possible to image the dynamics of the simultaneous impregnation of a support with two
active components, such as Cu 2+ and PtCI/, and to visualize the final distribution of both
components after the impregnation [9].
NMR can be used to explore various types of mass transport. We have studied the
drying of alumina pellets initially saturated with liquids such as acetone, benzene,
cyclohexane and water [4]. For pellets saturated with water, a more or less uniform
distribution of the liquid within the sample is maintained throughout most of the drying
201

process, despite the fact that evaporation proceeds mainly at the sample periphery. This is
the result of an efficient capillary flow of water which leads to its rapid redistribution
during drying, preventing the formation of large water content gradients. A detailed
analysis has shown that the fine details of the profile transformations during the sample
drying depended on the details of the pore size distribution in the sample under study. For
the alumina samples with a pronounced "bimodality" in the pore size distribution, the
profile shape undergoes a rectangular ~ round-top ~ rectangular ~ round-top sequence of
transformations in the course of the drying process. This peculiar behavior arises because
the concentration and pressure gradients are interrelated through the slope of the cumulative
pore size distribution curve corresponding to the current saturation, and because of the
tendency of the local concentration gradient, i.e., the profile shape, to readjust itself
permanently in order to maintain a certain capillary flow of liquid toward the sample
surface [5]. The transport of liquid within the pellet during its drying was modeled
successfully in terms of the diffusion equation [5,12]. In the case of water, the extracted
effective diffusivity (D) values at large saturation (C) levels were 1-2 orders of magnitude
larger than the water self-diffusion coefficient. The peculiar non-monotonic transformations
of the profiles were reflected in the non-monotonic behavior of the D(C) dependence,
which exhibited local maxima at those pellet saturations which corresponded to plateaus in
the cumulative pore size distribution [4,5].
The inverse process, namely the adsorption of water vapor by a single alumina or silica
gel pellet impregnated with CaCh, was also investigated. In that case, the rate of water
transport was limited by the penetration of liquid into the dry regions of the pellet, and the
propagation of the sorbed water was characterized by a sharp front. It was established that
liquid transport at non-zero initial saturations was more efficient than that through a dry
pellet. The results of the drying studies of the alumina pellets described above have shown
that capillary diffusion in unmodified alumina pellets was very efficient and was
characterized by the diffusivity values which at large saturations exceeded the water self-
diffusion coefficient. It was concluded and verified experimentally that an egg-shell
distribution of the salt within the adsorbent pellet could increase the efficiency of the
transport of adsorbed water within the pellet. As mentioned above, the preparation
procedure of the egg-shell distribution of the salt was checked by substituting CuCI 2 for
CaCl z and detecting the relaxation time-weighted image of the pellet saturated with a
liquid. The adsorption process was modeled based on the diffusion equation [13].
We have extended the drying experiments to the next level of sophistication and studied
the drying of beds comprised of wet porous grains. The interesting issues here include the
dependence of the drying rate on the pellet position in the bed, e.g. the distance from the
center of a cylindrical bed. Care must be taken, however, since if the properties of the
individual pellets differ, the radial dependence will be masked by the variation in the drying
rate caused by the pellets differences. The two types of beds were studied, one comprised of
porous pellets and another comprised mostly of solid pellets with a small number of porous
pellets added. Similar results obtained for the two types of beds indicated the absence of
significant readsorption during bed drying. A reliable modeling of bed drying can be done
if it is based on the results obtained in the single pellet drying experiments.
NMR imaging can provide useful information on mass transport of liquids, and possibly
gases, both within the porous walls of the shaped catalysts (internal mass transport) and
across the gas/solid interface at the channel walls of the catalyst (external). As an example,
we have studied the process of drying of the alumina monolith, initially saturated with
water, by a stream of dry air passing through its channels [6]. The I D water content profiles
202

along the monolith were detected in the course of drying. Due to the existence of an
efficient capillary flow within the walls, the water content decreases uniformly along the
monolith until the contiguous system of liquid elements within the pores is disrupted at the
late stages of drying. The efficient capillary redistribution was already observed in the
alumina pellet drying experiments, and for the monolith was confirmed as follows. The
original monolith was sliced in 2 mm thick sections perpendicular to the channel axes. The
sections were then glued back together with epoxy to restore the original monolith shape,
but with the impermeable epoxy layers preventing water from moving between the sections.
Whereas the original monolith dried uniformly, the behavior of the structurally modified
sample in the course of drying was entirely different. With the modified monolith it was
possible to measure the rates of external mass transfer at various distances from the inflow
edge of the monolith and to demonstrate that the rate of external mass transfer decreases on
going from the inflow edge toward the outflow edge, possibly due to the variation of the
gas flow pattern along the sample. Work is currently in progress to extract mass transfer
parameters from the experimental data.
Spatial distributions of flow velocity which allegedly cause the non-uniform drying
behavior of the sliced monolith sample can be visualized directly by NMR flow imaging. In
general, this approach is useful for liquids, gases and solid granular materials in large
channels and in porous media with relatively large pores.
In the next set of experiments we have employed NMR flow imaging to visualize flow
patterns of water and protonated gases with equilibrium nuclear spin polarization flowing
through the channels of monolithic catalysts. Figs 1 and 2 show 2D flow velocity maps
detected transverse to the direction of the monolith channels at various distances from the
inflow edge of the monolith, and the lD cross-sections of these maps. Fig. 1 visualizes the
flow velocity component directed along the channels, while Fig. 2 shows the velocity
component directed horizontally in the image plane. The flow maps of the third velocity
component are analogous to the maps of Fig. 2 rotated by 90° in the image plane.
The experiments were performed with a cylindrical fragment of the monolith possessing
channels of square cross-section. It was placed in a cylindrical cell 20 mm above the
bottom edge of the cell, with the cell axis coincident with the axis of the monolith and
parallel to the direction of its transport channels. Water inflows into the cell at the bottom
through the narrow supply tubing approximately along the axis of the sample (central peak
in Fig. 1a). As the stream approaches the lower edge of the monolith, it undergoes distinct
transformations and acquires velocity components in the image plane, i.e., perpendicular to
the direction of the main flow (a negative and a small positive peaks in the center of Fig.
2a). Eventually, this causes the main stream to separate into four parts entering the four
central channels of the monolith (positive peaks in Fig. 1c). The velocity of displacement of
the four streams away from the center of the monolith is largest at the entrance section of
the monolith, which for the left and right hand side streams manifests itself as the two peaks
of opposite sign in Fig. 2b (the upper and lower streams do not have a horizontal velocity
component but rather shift in the vertical direction within the image plane). The flow
pattern complicates further as the streams proceed along the monolith channels. In Fig. lc
the four streams are crescent-shaped, with the centers of the crescents still shifting away
from the center of the monolith, while their edges have already started to shift in the
opposite direction (Fig. ld). This is confirmed by the sign alteration of the horizontal
velocity component within each channel in Fig. 2c. As the streams proceed further along
the channels they gradually expand to fill out the entire cross-section of the channels, which
Figure I. Two-dimensional now maps showing the spatial distribution of water now ve loci ti es in a cylindrical
cell containing the monol ith, with the one-dimensional horizontal cross-sec tions superimposed. The velocity
component perpendicular to the image plane is shown. The 20 ma ps were detected at various distances from the
innow edge of the monolith. These distances arc indicated in the figure , with the negati ve values corresponding to
the cross-sections outside of the monolith.

naturally leads to the reduction of the peak flow velocities (Figs I c-t). Eventually, the
developed flow pattern characteristic of the square cross-section channels is established.
We have estimated the Reynolds number for this experiment as Re=72 assuming that
most of the flow passes through the four central channels. This corresponds to the entrance
length of ca. 17 mm (0.06Rexd), in a reasonable agreement with the experimental
observation that the transition to the developed flow pattern within the monolith is almost
complete at 18.5 mm from the inflow edge. It is evident from Fig. I that channels with
smaller velocities of flow are characterized by smaller entrance lengths.
Furthermore, the flow maps clearly reveal the presence of another four streams
(negative external peaks in Figs la,b) flowing in the direction opposite to that of the main
flow. These streams come out of the external comers of the four central channels and
gradually shift away from the sample axis (Fig. I b, Fig. 1a), in agreement with the
observation of the horizontal velocity components for these streams (two external peaks of
opposite sign in Fig. 2a). Besides, another four reverse streams are observed inside the
monolith near the inner comers of the four central channels (two small negative inner peaks
in Fig. lc).
Similar experiments were performed with acetylene and butane gases flowing in the
channels of the monoliths at the atmospheric pressure. They resulted in a very first
observation of a flow map for gases with thermal equilibrium polarization of nuclear spins
[14], demonstrating the feasibility of such studies at moderately high magnetic fields (7T)
with true in-plane resolution of 400 11m and reasonable detection times. The 2D gas flow
204

1.5
a)
-.
-1.5 mm
4
1.0
CII as
.] 00
6.Q5
'uo -10.
~-15
> . ~~~~~~~~~
-10 -5 0 5 10 -10 -5 0 5 10
distance, mm distance , mm

d) 10.5 mm

as
CII

E
()

~OO
()
o
~.Q5
>
'{}5 00 05 -10 -5 0 5 10
ve locity, em's distance , mm
Figure 2. Same as in Fig. I, but for the veloc ity component directed horizontally in the image plane. Note that 10
cross-sections in Figs c and d are displaced from the middle of the 20 map. In all instances the zero velocity leve l
of the 10 cross-section defines the position of the cross-section relative to the 20 map.

images exhibited flow patterns that were not fully developed, in agreement with the range
of Reynolds numbers (190-570) and the length of the sample used in gas flow experiments.
The flow maps have revealed the highly non-uniform spatial distribution of shear rates
within the monolith channels of square cross-section, the kind of information essential for
evaluation and improvement of the efficiency of mass transfer in shaped catalysts.
NMR flow imaging becomes progressively more difficult as channel or pore sizes are
diminished. This is especially true for gases due to lower SIN and faster diffusion as
compared to liquids. In the situation when velocity is no longer constant across the voxel of
an image, averaging of the velocity over each voxel inherent in the measuring technique
leads to a substantial distortion of the observed flow pattern. In this case, dynamic NMR
imaging (combination of NMR imaging and PFG displacement spectroscopy) can be
employed to recover the distribution of velocities or displacements within each voxel. As
an extreme case, spatial resolution can be abandoned completely, and the entire object can
be considered as a single voxel. PFG displacement spectroscopy has an advantage that it
can detect both flow and diffusion, and is widely used to characterize transport of liquids in
porous materials.
In certain cases, such as flow of liquids or gases in relatively wide channels and in beds
comprised of large grains, the results obtained by NMR flow imaging and PFG
displacement spectroscopy can be compared. We have performed such comparison for
water and gas flow in a straight circular pipe with 7.6 mm inner diameter [15]. The 2D flow
maps and their 1D cross-sections along the tube diameter have revealed the parabolic
205
velocity profiles for water flowing at Re=285 and butane at Re=580, while for water at
Re= 1825 the flow velocity distribution was distorted and the 1D cross-section clearly
deviated from a parabola. In the latter case, the entrance length exceeded the length of the
straight section of the tube, leading to the observation of the flow pattern which was not
fully developed.
The flow velocity distributions were then constructed for each of the 2D flow maps as
histograms of the number of pixels corresponding to each velocity value. Besides, the PFG
experiments were conducted under the same experimental conditions to measure such
distributions directly: There is a general agreement between the histograms obtained from
velocity maps and the average propagators measured directly. Some differences are caused
by the inflow/outflow of water or gas during the measurement. Besides, for gases the
propagator appears broadened at the edges owing to the substantial diffusive displacements
within the observation time interval [15] .
Propagator studies of water flow in beds of solid beads are performed routinely and
yield average velocities and hydrodynamic dispersion coefficients. We have extended
successfully the PFG NMR experiments to study the flow of butane with equilibrium spin
polarization in a cylindrical bed of solid glass beads (Fig. 3a). Similar to the studies
performed with liquids, such experiments can provide information on the average axial
velocity and on the coefficients of both axial and transverse hydrodynamic dispersion.
First, a set of propagators were detected with the flow of gas turned off in order to have
a useful reference. In that case, the displacements of the gas molecules were solely due to
molecular diffusion. For the entire range of the observation time intervals from 1 ms to 600
ms, the propagators exhibited Gaussian shape with a progressively increasing width. The
corresponding diffusivity values appeared to decrease with time, as expected, but the scatter
of data prevented us from making an unambiguous conclusion. Two similar sets of
propagators were then obtained with the gas flow turned on, and with the magnetic field
gradient appl ied either parallel or transverse to the bed axis for the measurement of the
axial or transverse displacements, respectively. The transverse propagators also had the
Gaussian shape for all observation times. At short times (1 ms), the value of the transverse
dispersion coefficient was close to that of the diffusivity measured when the flow was
turned off (0.034 cm2/s), in agreement with the fact that at short times the displacement is
dominated by diffusion. However, as the observation time increased, the · transverse
dispersion coefficient first increased, reached a maximum of 0.064 cm 2/s at around 80 ms
and then decreased below 0.055 cm2/s. In the case of the axial displacement measurements,
the propagators for short «5 ms) and long (>150 ms) observation times were Gaussian in
shape, while for the intermediate times they exhibited pronounced tails toward large
displacements. The extracted axial dispersion coefficient first rapidly increased with
observation time and then leveled off reaching a value of more than 0.22 cm2/s, which was
at least a factor of 6.5 larger than the diffusivity value. The experiments were performed at
the average flow velocity of 2.91 crn/s. The information on average flow velocity and
hydrodynamic dispersion should be useful in a variety of applications which rely on the
flow of liquids and gases in porous materials.
Flow of fine solid materials through porous beds is important in many technological
processes. We have applied both NMR flow imaging and PFG NMR to study the gravity
driven granular flow of 100-200 11m porous alumina grains partially saturated with vaseline
oil through a cylindrical bed comprised of 5 mm alumina beads. The transverse 2D maps of
the axial velocity component have revealed that flow was not localized exclusively near the
walls, the information not readily available with other techniques for optically non-
206

a) b)
1 1
E E
..!2 ..!2
..-
..-

Of!!!!lo"""'" o
~--~--~--~--~--~
-0.5 0.0 0.5 1.0 1.5 2.0 -1.0 0.0 1.0 2.0 3.0
Displacement, cm Displacement, cm
Figure 3. Probability density distributions of the axial displacements for butane (a) and water (b) flowing through
a bed of3.2 mm glass beads for an observation time of300.1 ms. Solid line in (a) shows a Gaussian fit.

transparent beds. However, SIN and the spatial resolution were not high enough to avoid
velocity averaging within the voxels. Therefore, PFG NMR was employed to obtain
displacement distributions without spatial resolution. The experiments were carried out in
the regime when the bed was almost completely flooded with the granular material due to
the large flow rate. The propagator was found to exhibit a large sharp peak at velocities
<0.5 cmls and a broad declining wing which stretched beyond 15 cm/s. In a number of
experiments, an additional small peak was observed at 4.5 cm/s. The studies of other
granular flow regimes are currently in progress.
The peak near zero displacement is often observed in the liquid flow experiments and
indicates that some fraction of the liquid is partially immobilized within the bed. Such a
situation is usually encountered when the bed is comprised of porous grains, with some
liquid residing within the pores. In such case, the flow can be coupled to the transport of
liquid within the grain pores through the exchange between the entrapped and the flowing
fractions of the liquid. However, even for non-porous grains some liquid can be
immobilized near the grains contacts and in the dead-ended pores. The exchange can be
driven by a number of mechanisms, such as diffusion or localized eddies. Our experiments
have revealed the presence of the stagnant fraction of water for a cylindrical bed of 23 mm
in diameter comprised of solid 3.2 mm glass beads (Fig. 3b) or cylinders. The peak at zero
displacement was insensitive to the flow rate, but decreased as the observation time
increased. This result seem to imply that the exchange process is driven by diffusion, since
for a convectively driven exchange one could expect a variation of the exchange rate as
flow velocity is altered. With the characteristic time of the disappearance of the stagnant
fraction being on the order of Is, the size of the stagnant pools can then be estimated as ca.
70 J.lm. The fact that the flow velocity distribution histograms constructed from the flow
maps obtained under the same experimental conditions were lacking the peak near zero
displacement indicated that the stagnant pools were indeed much smaller than 0.2xO.2x2
mm 3, the voxel size of the images. However, the conclusion about the diffusive nature of
exchange is tentative, since the range of velocity variation was limited, and the possibility
that higher flow rates could increase the rate of exchange can't be excluded completely.
We have encountered a coupling of flow with the intragrain transport when studying
flow of gases in porous beds comprised of porous silica gel or alumina grains. This
situation, however, is further complicated by the strong adsorption of the protonated gases
on the pore surface of the silica gel and especially the alumina grains. Thus the two coupled
transport processes are further coupled to the adsorption process. For the bed of silica gel
pellets impregnated with CaCl 2 containing I atm of butane, the PFG NMR experiment has
yielded a propagator which represented a sum of the two Gaussian curves. The narrow one
207
(D=4.8x 10-4 cm2/s) had larger intensity and was tentatively ascribed to the adsorbed butane,
while the broader component (D=5 .7x 10-3 cm2/s) presumably corresponded to the gaseous
fraction. This assignment was in agreement with the following observation. When the flow
of gas was turned on, the larger component remained unchanged, while the smaller one
became asymmetric and shifted in the direction of the flow. The increase in the observation
time and gas flow rate led to the observation of the dominating tailing contribution of the
desorbed gas due to the coupling of the adsorption/desorption and transport processes. In
the case of the alumina grains the adsorption of butane was much more efficient, and the
static propagator did not reveal any contribution of the gaseous fraction.
The information on the dispersion coefficient for the gas flowing through the bed of
silica gel grains impregnated with CaCl2 would be useful in our studies of the water vapor
adsorption by such beds when moist air is passed through the bed. We have performed a
number of such experiments [6], but the mathematical modeling of the process is hampered
by the coupling of a number of processes and the lack of information on some essential
parameters governing these processes. Such information, however, is accessible. The
experiments described in the previous paragraph were intended to yield the dispersion
coefficient essential for the description of the gas flow through the bed. The adsorption
process and the transport within the individual grains constituting the bed were already
studied in the single pellet adsorption experiments described above. The same approach can
be employed to model the bed drying experiments based on the gas flow studies and the
results of the single pellet drying studies.
The discussion of the drying and adsorption processes presented above was clearly
oversimplified in that it did not account for the temperature variations associated with these
processes. While in some situations this could be justified, this is no longer the case for our
studies of water vapor adsorption in the absence of air by the silica gel pellets impregnated
with CaCh [9,10]. The direct measurements indicated that the process was accompanied by
a temperature rise from RT to about 50 DC. Thus the situation is characterized by the further

en
:!: 0.8
b)
c:
::l

...
0.6
.0
ro 0.4
>-
~ 0.2
c:
~ 0.0
c:
45 QO Q5
Distance, cm
Figure 4. A 2D image of a Ptlalumina catalyst pellet inside a glass tube filled with butane at I atm pressure (a)
and its ID cross-section (b). The position of the cross-section is shown with the solid line in (a).

coupling of phase transition and mass transport processes to heat transport within the
sample. Further experiments and modeling of vacuum adsorption are currently in progress.
The top of this pyramid is occupied by a heterogeneous catalytic reaction, which
embraces all sorts of mass transport processes, phase transitions, adsorption/desorption and
heat transport coupled to each other and to the chemical transformations. The real time
NMR imaging of a heterogeneous catalytic reaction is still a challenging problem to be
addressed. As a preliminary step in this direction we have performed NMR imaging of the
208

actual catalyst pellet (Pt! AhO)) residing within the non-operating reactor by filling the
entire system with butane. Strong adsorption of butane on alumina discussed above allowed
us to visualize the catalyst pellet (Fig. 4). The presence of a thermocouple inserted in the
catalyst pellet produced an artifact which was larger than the actual hole in which the
thermocouple was residing. Nevertheless, the results demonstrated that the images of
reasonable quality could be obtained. Further work in this direction is currently in progress.

Acknowledgments. Liquid and gas flow experiments are performed in collaboration with
New Mexico Resonance, Albuquerque, NM, USA. We thank our colleagues from Boreskov
Institute of Catalysis for providing the samples and for useful discussions. The work was
partially supported by the grants from the Russian Foundation for Basic Research (# 99-03-
32314a and supplementary grants 01-03-06091 and 01-03-06092), Russian Academy of
Sciences (grant # 150, competition of the research projects of young scientists #611999), the
Siberian Branch of RAS (integration project #46). A.V. Matveev acknowledges a
scholarship awarded by the Zamaraev International Charitable Scientific Foundation. L.Yu.
I1yina acknowledges the young scientists' competition grant from SB RAS.

5. References
1. Hansen, E.W., Tangstad, E., Myrvold, E. , and Myrstad, T. (1997) Pore structure characterization of
mesoporous/microporous materials by IH NMR using water as a probe molecule, 1. Phys. Chem. BIOI,
10709-\0714.
2. Allen, S.G., Stephenson, P.C.L., and Strange, J.H. (1997) Morphology of porous media studied by nuclear
magnetic resonance, 1. Chem. Phys. 106,7802-7809.
3. Latour, L.L., Kleinberg, R.L., Mitra, P.P., and Sotak, C.H. (1995) Pore-size distributions and tortuosity in
heterogeneous porous media, 1. Magn. Reson. A 112, 83-91.
4. Koptyug, l.V., Fenelonov, V.B., Khitrina, L.Yu., Sagdeev, R.Z., and Parmon, V.N. (1998) In situ NMR imaging
studies of the drying kinetics of porous catalyst support pellets, 1. Phys. Chem. B 102, 3090-3098.
5. Koptyug, LV., Kabanikhin, S.\., Iskakov, K.T., Fenelonov, V.B., Khitrina, L.Yu., Sagdeev, R.Z., and Parmon,
V.N. (2000) A quantitative NMR imaging study of mass transport in porous solids during drying, Chem.
EngngSci. 55, 1559-1571.
6. Koptyug, l.V., Sagdeev, R.Z., Khitrina, L.Yu., and Parmon, V.N. (2000) A nuclear magnetic resonance
microscopy study of mass transport in porous materials, App/. Magn. Reson. 18, 13-28.
7. Koptyug, \'V., Khitrina, L.Yu., Aristov, Yu.L, Tokarev, M.M., Iskakov, K.T., Parmon, V.N., and Sagdeev, R.l.
(2000) An IH NMR microimaging study of water vapor sorption by individual porous pellets, 1. Phys. Chem.
B 104, 1695-1700.
8. Khitrina, L.Yu., Koptyug, \'V., Pakhomov, N.A., Sagdeev, R.Z., and Parmon, V.N. (2000) An IH NMR
microimaging visualization ofhexachloroplatinate dianion redistribution within a porous gamma-AhO) pellet
in the course of supported catalyst preparation, 1. Phys. Chem. B 104, 1966-1970.
9. Koptyug, l.V., Khitrina, L.Yu., Parmon, V.N., and Sagdeev, R.Z. (2001) NMR imaging of mass transport and
related phenomena in porous catalysts and sorbents, Magn, Reson. Imaging 19, 531-534.
10. lIyina, L.Yu. (2001) Investigation of mass transport of liqUids in porous catalyst and sorbent grains by /H
NMR imaging in situ, Ph.D. Thesis, Novosibirsk (in Russian).
II. Boddenberg, B. and Beerwerth, B. (1989) Proton and deuteron magnetic resonance relaxation of benzene
adsorbed on alumina and on a platinum/alumina catalyst, 1. Phys. Chem. 93, 1440-1447.
12. Kabanikhin, S.\., Koptyug, \'V., Iskakov, K.T. , and Sagdeev, R.l. (1998) Inverse problem for a quasi-linear
equation of diffusion, J. Inv. l//-Posed Problems 6, 335-351 .
13. Kabanikhin, S.l., Koptyug, l.V., Iskakov, K.T., and Sagdeev, R.Z. (2000) Inverse problem for the diffusional
transport of water upon single pellet moisture sorption, Int. 1. Nonlinear Sci. Num. Simul. 1,29-41.
14. Koptyug, l.V., Altobelli, S.A., Fukushima, E., Matveev, A.V., and Sagdeev, R.Z. (2000) Thermally polarized
IH NMR microimaging studies of liquid and gas flow in monolithic catalysts, J. Magn. Reson. 147,36-42.
15. Koptyug, LV., lIyina, L.Yu., Matveev, A.V., Sagdeev, R.Z., Parmon, V.N., and Altobelli, S.A. (2001) Liquid
and gas flow and related phenomena in monolithic catalysts studied by IH NMR microimaging, Catal. Today,
accepted.
NMR SPECTROSCOPY CONTRIBUTION TO THE STUDY OF
BIOMATERIAL MINERALISATION

A.P. LEGRANDt, B. BRESSON!, R.GUIDOIN 1, R. FAMERy2,


J-M. BOULER3
iLaboratoire de Physique Quantique. FRE CNRS 2312. ESPCl. 10, rue
Vauquelin 75005 PARIS. France
2Laboratoire Ceramiques et Materiaux Mineraux UMR 7574
ESPCl. 10, rue Vauquelin 75005 PARIS. France
3Centre de Recherche INSERM 99-03 sur les Materiaux d'Interet
Biologique Faculte de Chirurgie Dentaire - BP84215. 44042 Nantes
Cedex 1. France

ABSTRACT. High resolution solid state NMR spectroscopy is a powerful tool for
understanding bone structures, bone substitutes and implants. In particular it can be
used to estimate osteoformation via bioceramic bone colonisation. Pathological
calcification occurring in bioprosthetic heart valves and breast prostheses can be
characterised.

1. INTRODUCTION

The study of calcification processes involved in the mechanisms of biodegradation,


osteoconduction and osteoinduction requires numerous methods of investigation.
Complementary techniques such as X-ray, Infrared and Raman spectroscopies can be
used. More recently Solid State Nuclear Magnetic Resonance has proved its value for
the characterisation of biomaterials, bone mineral phases and pathological
calcifications.
The characterisation of bioresorbable ceramics used in orthopaedic and
maxillofacial surgery through osteoformation (1), the selection of the best
characteristics for such materials (2-3), and the evaluation of the influence of drug
delivery into the bone structure (4-5), require a variety of approaches. The inorganic
part of bone mineral can be characterised using solid-state NMR. This method was
used a long time ago, but the evolution of the NMR technology makes it possible to
improve such analyses.
209
J Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 209-218.
© 2002 Kluwer Academic Publishers.
210

After implantation of bioprosthetic heart valves or silicone breast implants a


calcification process develops, the origin of which has been discussed. This unwanted
pathological phenomenon has to be understood and eradicated.

2. EXPERIMENTAL

X-ray. Powder X-ray diffraction experiments between 10° and 80° in 20 were
conducted using a Philips automatic diffractometer with Cu-Ka radiation. The step
scanning conditions were 0.02° for 20 and 3 mn for the acquisition time. Transmission
mode powder X-ray diffraction experiments were run on an INEL curve detector
calibrated with NAC.
NMR. Measurements were made at room temperature on a ASX-300 Broker
spectrometer using Magic Angle Spinning (MAS) at a rotation speed as high as 14
kHz depending on the nucleus studied. Powder samples averaging 100 mg are
necessary to fill the rotor of the NMR probe. Due to a long spin lattice relaxation time
particular attention has to be taken regarding the recycle time for the acquisition of
each spectrum.

3. MATERIALS

3.1 MACROPOROUS BIPHASIC CERAMIC PHOSPHATES (MBCP)


These consist of mixtures of hydroxyapatite and ~-tricalcium phosphate. They are
widely used as bone substitutes in orthopaedics and oral surgeries (6-10). The interest
in such materials has necessitated the definition of specific standards based on the
application of the X-ray diffractIon method (11-13). The recent development of
minimal invasive surgery requires that such biomaterials be adapted to the medical
devices involved and also to new surgical indications (14). These reasons led to the
development and study of an injectable bone substitute consisting of a suspension of
MBCP granules associated with a viscous polymeric carrier (15). Various in vivo
implantations have been perfonned in rabbit bones (16-17).

3.2 BREAST IMPLANTS


In spite of continual advancements in surgical techniques and development of new
devices, late complications such as rupture and/or mineralisation of prostheses remain
a significant factor in breast augmentation surgery (18). The durability of these
implants does not exceed the life expectancy of the patients. The reoperation rate of
patients is very high due to:
• excessive fibrous capsules (19-20)
• percolation of silicone gel through the silicone shell
• change in shape
• rupture of prostheses (21)
• mineralisation (22)
Breast prostheses were collected at different hospitals in Canada and the USA
participating in the explanted devices retrieval program launched by the Quebec
211

Biomaterials Institute (Laval University, Quebec, Canada). Different highly


mineralised models of devices implanted for ten years or more were selected (23).
They include silicone bags (saline-filled breast implants), gel-filled breast implants
with or without polyester (Dacron ®) patches and double lumen breast implants. No
prosthesis was PU-coated or filled with hydrogel or oil. All silicone rubber surfaces in
contact with tissues were smooth, i.e. none of them was texturised.
At explantation, capsules of tissues are dissected. Both prosthesis and capsule are
left in formalin (buffered 10%). Some minerals are then left on their own and
precipitate in a formalin solution. Most of them can be dissected from the tissues.
Contrary to a previous protocol (24-25), the capsules were not boiled in sodium
bicarbonate and sodium hypochlorite to eliminate the tissues. They were air-dried and
subjected to analysis.

3.3 PORCINE AND PERICARDIUM VALYES


Such bioprostheses, also collected at different hospitals in Canada and the USA,
have two origins:
• porcine: aortic leaflets and/or valves of pigs treated with glutaraldehyde are
mounted on stents (rigid or soft) and chemically processed in a buffered solution
of glutaraldehyde at low pressure.
• pericardium: the pericardium of the heart of calf is trimmed, mounted on a stent
and fixed with glutaraldehyde
Replacement of human valves concerns aortic, mitral and tricuspid positions. It is
frequent at the aortic (dia. 20-23 mm) and mitral (dia. 23-31 mm) sites. It is less
frequent at the tricuspid site (dia. 22-30 mm). Reoperations on the valves implanted at
the aortic and mitral positions are very frequent after 7 to 10 years.
The origins of the progressive complications, in particular for the left side of the
heart, are degradation and mineralisation which reduce the opening of the leaflet and
permit blood leaks at closure (26-28). Despite these weaknesses, these valves are
recommended for elderly patients and pregnant women as systemic heparinisation is
not necessary.

4. RESULTS AND DISCUSSION

4.1 PROGRAMMED MINERALISATION


MBCP.
Analysis of the materials as prepared. Two MBCP were used, 45/55 and 60/40
(±4) HAP/~-TCP, the composition of which was determined by X-ray analysis
following (11). In figure 1 are presented 31p MASINMR spectra of pure HAP and ~­
TCP compared with 45/55 and 60/40 HAP/~-TCP MBCP samples. MBCP samples are
neither a pure nor a simple mixture of ~-TCP and HAP. There is a broad line around
2.8 ppm related to the existence of crystalline and amorphous hydroxyapatite. For the
tricalcium phase, the broader line, under the previous one, does not look like that of a
pure ~-TCP. Consequently, spectra must be decomposed to obtain a precise estimate
of the relative proportion HAP U /~-TCP. This shows that the standard X-ray
212

determination for such partially crystalline materials is inadequate for structure and
composition analysis.
Analysis of the materials after explantations. X-ray, IH, 31 p MASINMR and
CPIMAS where used for different purposes.
• 45/55 HAP/~-TCP MBCP. 31p MAS 78 weeks explant (Figure 2) shows an
important evolution of the material. A bone-like peak around 2.8 ppm is observed.
31 p MAS saturation recovery shows that the spectrum is constituted of two type of
materials (Figure 3). The same peak at 2.8 ppm is observed and a residual broad
line attributed to unresorbed MBCP part.

45% HAP/55% 13-TCP MBCP


,',
" ~ pure ~-TCP

I
I
, ''/
,' II
I
I
,
I ,
, I

,,
I

I
I
,
I

, I

",.,--'
31 pMAS ,, ,,
/ ,,
,, ,,
I

,, "
............ - .... _---
--------'

10 5 0 -5
ppm (/HAP=2.8ppm)
FIGURE 1. 31 p MASINMR spectra of pure HAP, pure ~-TCP and 45/55 or 60/40
MBCP.

• 60/40 HAPI~-TCP MBCP. 31 p MAS spectra of the two explants and of the
original MBCP are shown in figure 4. For comparison, spectra are presented with
the same maximum amplitude. Differences are observable, in partic~lar
concerning the influence of the amount of hGH (recombinant human growth
hormone, UMATROPE, Lilly, 16UI) added before implantation. It seems that the
process of resorption is not just a simple transformation of the implant. It is
generally accepted that dissolution-reprecipitation uccurs and it can be seen that
this phenomenon is influenced by the presence of hGH. The greater the amont of
hGH, the broader and smoother the signal. Analysis of the X-ray (INEL) diagrams
show that two phases are present in both samples studied (Figure 5):
hydroxyapatite (JCPDS 9-0432) and whitlockite (JCPDS 9-0169).
213

"PMAS-N 78 week Explant


IritialMBCP 45/55 HAP~-TCP
45155 HAP.j}TCP MBCP

W 9 8 7 8 5 432 1 0 ~ 4 ~ 4 ~

10 ppm (/HAP=2.8
ppm (IHAP=2.8 ppm)
FIGURE 2. 31p MASINMR of initial FIGURE 3. 31p MASINMR saturation
45/55 MBCP and 78 weeks explant. recovery of a 78 weeks 45/55 MBCP
explant showing the composite structure
of the spectrum in figure 2.

MBCP l.p MASINMA


(HAP/II-TCP:W40)

MBCP + 10 ~g hGH

10 2 044 ~
10 o -2 4 -6
ppm (HAP: 2.8 ppm)
ppm (HAP: 2.8 ppm)

FIGURE 4. 31p MASINMR of 60/40 MBCP with 0.1 and 10 Ilg hGH amount added
(left). IH ~ 31 p CPIMAS spectrum (5 ms contact time, 10 kHz speed of rotation)
compared with the 0.1 Ilg hGH amount added 31 p MAS spectrum (right).
- MBCP 1110 hGH
_ Whitlockite
_ Hydroxyapatite

2k

lk

~ ~ ~ ~ ~ M ~ ~

D~fraction angle (2 8)

FIGURE 5. Explanted MBCP lllg hGH X-ray diffraction pattern (INEL).


214

4.2 PATHOLOGICAL MINERALISATION


Breast Implants.
It is not uncommon that breast implants become mineralised to various levels after
decade of implantation. As the fibrous capsules surrounding the implants shrink,
mineralisation occurs at the interface between the prosthesis and the fibrous capsule.
They accumulate as visible mineral pellets or as discrete dispersion in the fibrotic
tissues in the vicinity of the foreign surface. These samples were analysed by X-ray
diffraction (Figure 6) and 31p MAS NMR spectrometry (Figure 7).

2,01< Breast Prosthesis X-ray Reference

/
Calcified Deposit JCPDS 9-0432

500,0

M ~ ~ ~ ~ ~ u ~ ~ •
Diffraction angle (2 9)
FIGURE 6. X-ray diffraction pattern (Philips) of typical breast calcified deposit from
breast prosthesis.

Three different
Breast prostheses

~~f;~~

15 10 o
5 -5 -10

ppm (!HAP: 2.8ppm)


FIGURE 7. 31p MAS-NMR spectra of hydroxyapatite compared to calcified deposits
on breast prostheses. Recycle time 1000s, rotation speed 5.3 kHz, frequency 121.49
MHz, number of scan 4.
215

X-ray diffraction. Analysis of all the diagrams shows that hydroxyapatite (JCPDS
9-0432) alone is always present. Fluorohydroxyapatite (JCPDS 34-0010) was also
detected. Microanalysis reveals Ca and P in all samples. Si, Na and Mg are also
detected.

31p MAS NMR. Figure 7 shows a series of spectra compared to that of a well
crystallised hydroxyapatite. The line width of such samples is greater (3 ppm) than of
the latter (0.5 ppm). This effect is attributed to the partially amorphous structure of
such deposits. Comparison shows that the organisation of the calcified deposits is
different from that of a newly formed bone (figure 4; < I ppm).

19F MAS NMR. Fluorohydroxyapatite was detected as a trace and requires a very long
accumulation time. Recycle time 100 s, 13 kHz rotation speed, frequency 282.4 MHz,
number of scans 200 to 500.

Cardiac Valves.
Mineralisation was frequent in the porcine bioprostheses of the first generation. Most
of the explanted devices submitted for investigation devices were mineralised after 7
years of implantation. The deposits were analysed by X-ray diffraction (Figure 8) and
31 p MAS NMR spectrometry (Figure 9). Microanalysis reveals Ca and P in all
samples. Na and Mg are detected, and more rarely S.

1000
2.0k Mineralisations from
bioprosthesis
...3'
CD
:::J1.5k
cardiac valve
eoo~
:::J

- -
(II (II
;:;
'< eoo~
n n

- -
~1 .0k ~

~ 400~
500.0
200

0.0
0
20 22 24 26 28 30 32 34 36 38 40 42 44
Diffraction angle (2 e)
FIGURE 8. X-ray diffraction patterns (Philips) of typical calcified deposits from
cardiac valve and breast prostheses.
216

Calcified Deposits 31p MAS-NMR


of S,•• ,' PW~ "

,,
I, , I

,, I ,

,, ,I
, I

C alcified Deposits : Hydroxyapatite


of Cardiac Valve

I-----~--:::.~------" "
15 10 5 o -5 -10

ppm (/HAP: 2.8ppm)


FIGURE 9. 31 p MAS NMR spectra of hydroxyapatite compared to cardiac valve and
breast prosthesis calcified mineralised deposits. Recycle time 1000s, rotation speed 5.3
kHz, frequency 12l.49 MHz, number of scan 4.

X-ray diffraction. Analysis of all the diagrams shows that hydroxyapatite (JCPDS 9-
0432) only is always present.

31p MAS NMR Figure 9 is similar to that observed above (Figure 7). Similar
conclusions can be made about the organisation of such deposits.

Origin of the pathological mineralisations


Breast Prosthesis
As the mineral deposits are mainly hydroxyapatite, it is unlikely that the
mineralisation process is device-related:
• no mineralisation in contact with the polyester patchs and the adhesive.
• zinc acting as a catalyst for mineralisation was not found (24).
• mineralisation is therefore more likely to be patient and procedure related.
• mineralisation occurs in contact with natural tissues which have shrunk and then
could be partially necrosed (29) because prosthesis displacement particularly
when the patient is moving (30).
The presence of bacteria can also have some influence and this approach and must
be further investigated as the results are still controversial (31). Ideally, the prostheses
together with .the surrounding tissue capsules must preserve their integrity and their
softness. In addition the potential durability of the prostheses must outlast the life
expectancy of the patients. The situation is still far from this goal.
217

Porcine and Pericardium Valve Bioprostheses.


The degradation of bioprostheses is related to the appantlon of tears and
mineralisation (endophytic, inside the leaflets and exophytic, outside the leaflets)
which hinders the mobility of the leaflets resulting in a poor closure and regurgitation
After 5 years of implantation reoperations are frequently necessary.

Hypotheses
• endophytic: tissue necrosis, lipid uptake; debris are not removed because of the
absence of lymphatic drainage (30).
• exophytic: thrombostic debris colonised by bacteria of the biofilm (32) as the
result of an induced bacteremia (31).
Despite alternatives, as kangaroo heart valves (33), and different fIxation
methodologies (34-36), the prevention of mineralisation in bioprosthesis heart valves
are still a challenge. Acute and short-term calcifIcations have been eliminated but the
long term goal has not yet been reached.

CONCLUSIONS
Mineralisation occurring after surgical implantation of different medical
devices has been known for a long time. Nevertheless, this has positive and negative
effects. The characterisation of the calcifIed deposits or tissues is of interest for
improving the preparation of new materials for implantation. Solid-state NMR, in
association with other methods of analysis, is useful for the follow-up of the evolution
of the material implanted, although this method needs explanted samples for the
measurements.

ACKNOWLEDGEMENTS
Authors thank Miss Genois for X-ray sample preparations and for Mrs. Ollitrault for
TEM quantitative determinations.

REFERENCES
1. Marchandise X., Belgrand P., Legrand A.P., Magnetic Resonance in Medecine. 1992; 28:1-
8.
2. Miquel 1.L., Facchini L., Legrand A.P., Marchandise X., Lecouffe P., Chanavaz M.,
Donnazan M., Rey c., Lemaitre 1., Clinical Materials. 1990;5:115-125.
3. Bohner M., Lemaitre 1., Legrand A.P., d'Espinose de la Caillerie J.B., Belgrand P., J. Mat.
Sci. Mat. in Medicine. 1996;7:457-463.
4. Legrand AP., Sfihi H., Bouler J-M., Bone. 1999;25, 2:103-105S.
5. Bohic S., Rey C., Legrand A.P., Sfihi H., Rohamizadech R., Martel c., Barbier A, Daculsi
G., Bone. 2000;2:341-348.
6. Legrand A.P., Bruno B., Bouler J.M., Bioceramics. S. Giannini, A Moroni Edts.
2000; 13:759-764.
7. Daculsi G., Bagot d'Arc M., Corlieu P., GersdorffM., Macroporous Biphasic Annals of
Otology, Rhinology and Laryngology, 1992; 101:669.
8. Ransford AQ., Morley T., Edgar M.A., Webb P., Passuti N., Chopin D., Morin c., Michel
F., Garin C., Pries D., l Bone Joint Surg. (Br).1998; 80: 13.
9. Cavagna R., Daculsi G., Bouler J-M., l Long-Term effect ofMed. Imp!. 1999; 9, 4:403.
10. Gouin F., Deleerin l, Passuti N., Touchais S., Poirier P., Bainvel lV., Rev. Chir. Orthop.,
1995; 81:65.
218

II . Toth J.M.,. Hirthe W.M, Hubbard W.G., Brantley W.A., Lynch K.L., J. AppJ. Biomater.,
1991; 2:40.
12. Ishikawa K., Ducheyne P., Radin S., 1. Mater Sci Mater Med, 1993; 4:168
13. Association FranIYaise de nonnalisation (AFNOR), Nonne S94-066, 1993, AFNOR, Paris
14. Daculsi G., Weiss P.,. Bouler J-M, Gauthier 0., Millot F., Aguado E., Bone. 1999; 25,2:
59-61.
15. Weiss P. , Gauthier 0 ., Bouler J.M., Grimandi G., Daculsi G., Bone 1999; 25(2):67-70.
16. Gauthier 0., Bouler J.M., Weiss P., Bosco J., Aguado E., Daculsi G., Bone. 1999; 25,2:71 -
74.
17. Gauthier 0., Bouler I.M., Weiss P., Bosco J., Daculsi G., Aguado E., 1 Biomed Mater Res.
1999; 47,1:28-35.
18. Collis N., Sharpe D.T., Plast. Reconstr. Surg. 2000; 105:1979-1985
19. Kuhn A, Singh S., Smith P.D., Ko F., Falcone R, Lyle W.G., Maggi, S.P., Wells K.E.
Robson M.C., Ann. Plast. Surg. 2000; 44:387-391
20. Janowsky E., Kupper L.L., Hilka B.S. N. EngJ. J. Med. 2000; 342:781-790
21. Brown S.L., Middleton M.S., Berg W.A. , Soo M.S., Pennello G., AJ.R 2000; 175:1057-
1064
22. Peters W., Smith D., Lugowski S., Pritzker K., Holmyard D., Plast. Reconstr. Surg. 2001;
107:356-363
23. Guidoin R., Rolland C., King M. W., Roy P.E., Therrien M., Silicone in medical devices.
FDA Conf Proceed. Baltimore Feb 1-2, 1991
24. Rolland c., Ledoux R., Guidoin R. , Marceau D., lnt. 1. Artif. Org. 1989; 12 : 180-188
25. Rolland c., Guidoin R., Ledoux R., Zerguini R., Canadian Mineralogist. 1991; 29 :337-345
26. Grabenwoger M., Fitzal F., Gross c., Hutschala D., Bock P., Brucke P. Wolner E., 1. Heart
Valve Dis. 2000; 9: 104-109
27. Herijgers P., Ozaki S., Verbeken E., Van Lommel A, Jashari R, Nishida T., Leunens V.,
Flarneng, Thorac. Cardiovasc. Surg. 1999; 11,4:171-175
28. Vasudev S.c., Chandy T., Shanna C.P., Mohanty M., Umasankar P.R, 1. Biomat. AppJ.
2000; 14 :273-295
29. Andre-Frei V., Chevallay 8., Orly l., Boudeulle M., Hue A, Herbage D., Calcif. Tissue lnt.
2000; 66 :204-211
30. Proudfoot D., Skepper J.N., Hegeyi L., Bennett M.R., Shanahan C.M., Weissberg P.L.,
Circ. Res. 2000; 24: 1055-1062
31. Cisar J.O., De-Qi Xu, Thompson J., Swaim W., Lan Hu, Kopecko DJ., Proc. Nat. Acad.
Sci. (US) 2000; 97,21: 11511-11515
32. Hyde J.A., Daromeke R.O., Costerton 1.W., J. Heart Valve Dis. 1998; 7,3:316-326
33. Neethling W.M., Papadimitriou 1.M., Swarts E., Hodge, A.J., 1. Cardiovasc. Surg (Torino)
2000; 41:341-3<!·8
34. Suh H., Hwang Y.S., Park J.C., Cho B.K., Artif. Org. 2000; 24:555-563
35. Jasinski M.J., Kadziola Z., Keal R., Sosnowski A W., 1. Cardiovasc. Surg (Torino). 2000;
41: 181-186
36. Mirzaie M., Meyer T., Schwartz P., Dalichou H., 1. Heart Valve Dis. 2000; 9:576-582
MULTINUCLEAR MAGNETIC RESONANCE CHARACTERIZATION OF
SOLID CATALYSTS AND THEIR REACTIONS IN THE ADSORBED STATE

J. B.NAGya, P. LENTZa, A. FONSECAa, F. TESTAb, R. AIELLOb, Z.


KONYAC, I. HANNUSc and I. KIRICSlc
a) Laboratoire de RMN, Facultes Univesitaires Notre-Dame de la Paix,
61 Rue de Bruxelles, B-5000 Namur (Belgium)
b) Dipartimento di 1ngegneria Chimica e dei Materiali, Universita della
Calabria, 1-87030 Rende (Cs) (Italy)
c) Applied Chemistry Department, University ojSzeged, H-6720 Szeged
(Hungary)

1. Introduction

The high resolution magic angle spinning solid state Nuclear Magnetic Resonance has
opened new horizons in the characterization of solid catalysts [1,2]. Moreover, the
strongly or covalently adsorbed molecules as reactants, intermediates or products can be
easily detected and characterized [1,3].
The present authors have already published several review papers on the study of
the reactions in the adsorbed state using mainly l3C NMR [4-7]. On the other hand, the
characterization of zeolitic materials before, during and after the hydrothermal synthesis
allowed one to get a deep insight in the mechanism of formation of these microporous
materials. Emphasis was also put on the characterization of various atoms which
isomorphously substituted silicon in a tetrahedral form in the framework [8-11].
The present paper contains both already published as well as original materials. It
will be shown that more than four boron per unit cell can be introduced in the MFI
zeolitic framework using fluoride-containing synthesis media. The siting of aluminium
atoms will be tempted using 29Si NMR and both 1D 27Al NMR and 2D MQ MAS 27Al
NMR. Multi NMR characterization of both the solid phase and the adsorbed molecules
will provide a rather complete understanding on the reaction of chlorofluorocarbons
with zeolites and alumina. Finally, the reaction of l3C labeled methanol will be
presented on levyne catalysts.

2. Experimental

The zeolites and solid catalysts studied were either commercial products-Y-FAU
(Union Carbide), y-alumina (Degussa)-, or home made materials-HZSM5-MFI, B-
ZSM5, Levyne.
The in situ MAS l3C and 19F NMR measurements were performed on a Bruker MSL
400 spectrometer operating at 100.6 MHz (4.0 f.ls - e = n/2 pulse length) and 376.4
219
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 219-229.
© 2002 Kluwer Academic Publishers.
220
MHz (5.0 IlS - 9 = 1(/2 pulse length), respectively and using a 5.0 s repetition time. The
samples were packed into the NMR tubes and evacuated at 723 K for 2 h. The activated
zeolite or alumina samples were loaded with CClzF2 or CH30H, the tubes were
carefully sealed to achieve proper balance and a high spinning rate (ca. 4 kHz). Spectra
were recorded after heating the tube at various preselected temperatures.
For 29Si, 27AI, liB and 2~a NMR investigations the following parameters were
used. For 29Si (79.4 MHz), a 4.0 Ils ( 9 = 1(16) pulse, with a repetition time of 6.0 s; for
27Al (104.3 MHz), a LOllS (9 = 1(/12) pulse with 0.2 s repetitition time; for liB (128.3
MHz), a LOllS (9 = 1(112) pulse, 0.2 s repetition time and for 23Na (105.8 MHz), a 1.0
IlS (9 = 1(/12) pulse with 0.1 s repetition time. All chemical shifts are referred to
external standards: aqueous NaCI and AICh solutions for 23Na and 27AI,
tetramethylsilane for 13C and 29Si, liquid CCl3F for 19p and ChB·O(CH3)2 for liB.
The 3QMAS experiments of 27AI were perfomed on a Bruker ASX-400 at 9.4 T
using a 4 mm MQMAS probehead spinning at 15 kHz, with a radiofrequency field VRF
estimated at ca. 280 kHz. The pulse sequence was composed of three pulses
corresponding to the Z-filter MQMAS method [12], which yields pure absorption
spectra. The pulse lengths were experimentally adjusted to 2.1 Ils, 0.65 IlS (VRF = 280
kHz) and 9.0 Ils (VRF = 10 kHz). The recycle delay was the same as for the one-pulse
27Al MAS experiments. The delay t1 between the first and second pulses was regularly
incremented by 67 IlS, according to the method of rotor synthronization [13], which
allows one to remove spinning side bands along the isotropic axis and to significantly
reduce the acquisition time.

3. Results and discussions

3.1. CHARACTERIZATION OF B-ZSM-5 BY 29Si AND liB NMR

The synthesis of boron containing ZSM-5 zeolite is very efficient in fluoride media.
Indeed, more than four boron per unit cell can be introduced as tetrahedral boron in the
framework, if the hydrothermal synthesis is made at 170°C from gels of composition
9MF-xH3B03-10SiOz-1.25TPABr-330H20 with M = K or Cs, s = 6 or 10 and TPA =
tetrapropylammonium [14].
Figure 1 shows the liB NMR spectra of the as-made and calcined B-ZSM-5
samples synthesized with 9CsF and 10H3B03• It has to be emphasized that Cs+ at high
concentration of H3B03 behaves quite peculiarly. For example, the B/u.c. in the
framework increases during calcination with x = 4,6 and 10 moles. During calcination,
the NMR line at -4.0 ppm increases, while the one at -2.9 ppm decreases. There is also
some increase in the broad 5-17 ppm line. It seems that if the -2.9 ppm line which was
previously attributed to extra framework tetrahedral boron were transformed
essentially to the -4.0 ppm line, i.e. the boron in the structure in tetrahedral
configuration, and partially to non framework tetrahedral boron at 5-17 ppm. Hence,
we have to modify the attribution of the -2.9 ppm line, which is generally considered as
extra-framework boron. It is possible, that the extra framework tetrahedral boron in the
precursor is really not an extra framework boron, but characterizes a tetrahedral boron
which is partially linked to the structure. Note, that a similar species was previously
221

I I , ,

20 10 0 -10 -20
ppm
Figure I. MAS liB NMR spectra of as-made (a) and calcined (b) Cs-BZSM-5 sythesized with 10 H3B03.

supposed for aluminium, where it is called a defonned tetrahedral framework


aluminium (see below). If this is the case, a rather high amount ofSiOH defect groups
should be present in these samples. This is confmned by the 29Si NMR spectra, where
indeed a high concentration ofSiOX groups at -103 ppm is detected. Interestingly, the
K-borosilicate samples do not show any anomaly. No SiOX defect groups were
detected in the precursor samples at high boron concentration.
As a conclusion, the defonned tetrahedral boron can be represented as:

SiO OH HO OSi
~/
~B{; Si
/ ~ / ~
SiO OSi SiO OSi

where the tetrahedral framework boron is in a distorted configuration.


The attribution of the -2.9 ppm line to framework tetrahedral boron is further
supported by the comparison between the framework negative charges linked to the
tetrahedral boron and the positive charges (It and M") neutralizing them in the
222
calcined samples. It is supposed that H+ is stemming from the thennal decomposition
of TPA+ ions at high temperature, 470 °C, as these ions are known to neutralize the
(BOSi)" negative charges of the framework. The framework tetrahedral boron or
BT/u.c. is computed taking into account both -4.0 ppm and -2.9 ppm lines.
For the high boron content samples (x = 4,6 and 10) the agreement is reasonable
between the amount of positive charges (H + M) / u.c. and the framework negative
charges BT/u.c .. These data reinforce the attribution of the -2.9 ppm line essentially to
framework defonned tetrahedral boron. Note that no agreement was found at all if only
the -4.0 ppm line was taken into account.

3.2. SITING OF ALUMINIUM ATOMS IN THE LEVYNE STRUCTURE

Levyne is one of the oldest natural zeolite. The unit cell composed of 54 tetrahedral
atoms is rhombohedral. The cages are composed of hexagonal primsms and are
heptadecahedral (called levyne c~e). The structure contains two crystallographically
different sites, 36T 1 and 18T2. Levyne has a bidimensional system of channels of
elliptical pores of dimension 3.6 x 4.8 A.

Donna!
CP
as-made calcined

iii iii iii iii iii i iii iii i i j iii i i i i I

-70 -80 -90 -toO -110 -120 (ppm) -70 -80 -90 -100 -110 -120 (ppm)

Figure 2. MAS 29Si NMR spectra of as made and calcined samples at 700°C.

The magic angle spinning 29Si NMR spectra of the as made and calcined levyne
samples with and without cross polarization are shown in Figure 2. The levyne was
obtained using methyl quinuclidine iodide (MeQI) in the reaction batch [15]. It is
clearly seen that the spectra of the as made sample obtained with and without cross
polarization are similar, while they are quite different for the calcined sample. The
great change stems fonn the -98 ppm and -103 ppm lines which are attributed to
=Si== (OM)2 and == SiOM defect groups, respectively (M = H,Na,K or MeQ). The
different NMR lines were decomposed following the hypothesis that the chemical shift
ofSi on T2 sites are smaller [16]:
- 115 ppm: Si atomes ofSi(OSi)4 configuration on site T2, noted SiT2(OAl)
- 108 ppm: Si atomes on site Th SiT1(OAl) and Si atomes on site T2 with one Al
neighbour, Si(OAl) (OSi)3 noted SiT2(1Al)
223
- 103 ppm: Si atoms on site Tl with one Al neighbour, Si(OAI)(OSi)3, noted SiT1(lAI)
and Si atoms on site T2 with two Al neighbours, Si(OAIMOSi)2, noted SiT2(2AI)
- 98 ppm: Si atoms on site T1with two Al neighbours, noted SiT 1(2Al) and Si atoms on
site T2 with 3Al neighbours, Si(OAIMOSi)2, noted SiT2(3Al).
Taking into account the presence ofSiOM and Si(OM)2 defect groups at -103 and
-98 ppm, respectively, the decomposed spectra lead ·to the presence of a rather high
contribution of defect groups, ca. 18 % in the uncalcined sample.
These results cannot be explained adequately at present. Indeed, the levyne
crystallites could be inhomogeneous, the defect groups could occupy specific sites (e.g.
more defect groups on site T2), or the levyne crystallites also contain some additional
phases like LZ-132, LZ-133 or octadecasil. Let us note, that a borolevyne sample only
contains two NMR lines at -107 and -115 ppm.
The 2D 3QMAS 27Al NMR spectra of the as-made and calcined samples are more
relevant for the Al distribution. Figure 3 shows clearly the presence of two tetrahedral
Al atoms at 62 and 56 ppm, in an intensity ratio of 2: 1. These Al atoms occupy thus the
Tl and T2 sites, respectively, and the ratio of 2 shows that the Al atoms occupy
randomly both sites. The quadrnpolar product

(1)

with CQ the quadrupolar coupling constant and l1Q the asymmetric parameter of the
electric field gradient, is equal to 2.9 MHz and 1.9 MHz on site T1and T2, respectively.

6K+
as-made

"

Figure 3. 20 3QMAS 27 AI MR spectra of as made and calcined samples at 700°C.

During calcination, a new NMR line appears at 65 ppm (PQ = 4.8 MHz) and
another one at 4 ppm (PQ = 3.3 MHz). The former is attributed to a 'deformed
tetrahedral' Al atoms, while the latter characterizes an extra-framework octahedral Al
atom. The deformed tetrahedral Al atom is still in the framework, but one AI-O-Si
bond is broken:
224

Si-O O-Si

"'-/ Si

Si-O OH HO
/ "'- O-Si

The higher electric field gradient around this Al atom explains the rather high PQ value.
Note finally, that both the deformed tetrahedral Al atoms and the extra framework
octahedral Al are formed essentially from the Al atoms occupying T 1 sites.

3.3. REACTIONS OF CHLOROFLUOROCARBONS ON ZEOLITES AND


ALUMINA

Multi NMR of both solid and adsorbed phases is an excellent method to study the
decomposition of chlorofluorocarbons on zeolites and oxides. Chlorofluorocarbons are
environmentally harmful molecules and their decomposition or transformation to other
valuable products is a challenge for the chemical industries.
CChF2 (0 = 126.2 ppm, JCF = 323 Hz) is decomposed on NaY zeolite at 150 °c. An
intermediate COCh is formed which is identified in the adsorbed state at 8 = 142.7
ppm. The fmal product is CO2 (8 = 125.5 ppm). The decomposition is accompanied by
a dealumination of the framework that is shown by both 29 Si and 27Al NMR spectra.
After the reaction, the initial tetrahedral Al atoms (8 = 50 ppm) partially leaves the
structure and form extra-framework octahedral species at 8 = -3.4 ppm (ca. 15 %). The
29Si NMR spectra show that some 52 % of the initial silicon became amorphous [17].
The 23Na NMR spectrum ofthe initial dehydrated NaY zeolite shows three NMR lines
at -6, -25 and -55 ppm. They are attributed, respectively, to sodium ions in the
hexagonal prisms, in the sodalite units and in the supercages. The absence of the NMR
lines characteristic of the Na+ ions in the supercages and in the sodalite units of the
fmal sample suggests that the adsorption and reaction of CChF2 have taken place on
these centers. At the same time a new NMR line appeared at 6 ppm and at 0 ppm. The
former is attributed to NaCI while the latter to the presence ofNaF.
During the reaction of CChF2 with HZSM-5 zeolite, CClt (8 = 96.5 ppm) is also
detected as reaction intermediate. The intensity of the 19p NMR line of CChF2
adsorbed on HZSM-5 (8 = 6.1 ppm) decreased with increasing time at 200 °C. Finally,
this signal has disappeared and a new line at -163.4 ppm, due to AlF 3 appeared in the
spectrum.
The decomposition of CChF2 on y-alumina has been analyzed more completely in
the adsorbed sate. Figures 4 and 5 show, respectively, the \3C NMR spectra of CChF2,
intermediates and fmal products and the corresponding kinetic curves at 100 °C. Figure
4 shows that the decomposition of CChF2 already occurs at room temperature. Besides
the expected triplet of CChF2, a quadruplet attributed to CClF3 (8 = 125.2 ppm, JCF =
325 Hz) and a singlet due to CO2 (0 = 125.5 ppm) complicate the spectrum. CClt is
also detected at 96.6 ppm.
225

CChF1m
CChF (Q)
ca.
CO2 (S)

150 140 130 120 1 iO 100

Figure 4. MAS \3C NMR spectra ofCChF2 adsorbed on y-AhO) at room temperature (a) and at 100°C (b-d)
for 10 min, 30 min and 10 hours, respectively.

100

90 - - CCI2F2
- 0 - CCIF 3

80 - - - CCI4
----cr-- CO2
70 _ ____ _ _ 0 ----- 0 _______ 0
0
- 0 - 0
60 0 ·0
./'
0'
50

-----------------
40

30

20 '.-",,-
/ -
10 .-----
o _____ •-. _
o - o _~::::=--O
======== •0
o~~--~--~--~~~~--~--~--~~--~--~--~~--~

o 2 3 4 5 6 7 8 9 10 11 12 13 14

Time (hours)

Figure 5. Kinetic curves for the reaction ofCChF2 on y-AhO) at 100°C.


226

At 100°C, the reaction proceeds slowly toward the total conversion ofCChF 2 into
products, however the amount of CO2 remains small (Figure 5). After 14h, the
temperature was raised to 200°C. At this temperature the intensity of the CO2 line
increased at the expense of the CClF3 line, but the relative intensity of CC4 remained
constant. Finally, only at 300°C are the intensities of both CClF3 and CC4 lines
decreased and CO2 was observed as the final product. These observations prove the
intermediate character of CClF3 and CC4 in the decomposition. COCh is scarcely
detected in the course of the reaction.
The 19F NMR spectra confIrm the J3C NMR data. In addition to the -12.1 ppm line
due to CChF2, another line appeared at -33.9 ppm stemming from the presence of
CClF3. The intensity of the latter passes through a maximum as a function of time,
showing its intermediate character. The final product detected is AIF3 with 0 = -172.8
ppm.

3.4. TRANSFORMA nON OF METHANOL ON LEVYNE ZEOLITE

The J3C NMR spectra of methanol adsorbed on levyne (SiiAI = 13.6) as a function of
time at 100°C are shown in Figure 6. The CH30H is J3C labeled for 30 at %. The
methanol loading corresponds to 0.2 CH30WAI atom. The NMR lines at ca. 50 and 60
ppm stem from CH30H(MeOH) and CH3OCH3 (DME), the first product due to
dehydration of methanol. In most of the cases, each line shows two components, a
narrow one and a broader one. The narrow ones ·characterize the mobile molecules
(Physisorbed or in the gaseous state) and the broad ones the more strongly adsorbed (or
chemisorbed) molecules. The NMR line ofDME is narrower (M1- 150Hz) compared
to that of MeOH (M1 - 400Hz). Figure 6 shows clearly that the reaction of
dehydration reaches an eqUilibrium. The kinetic curves were computed based on a
second order equilibrium reaction:

with

where the [CH30H]o means initial concentration and [CH30H]eq is the equilibrium
concentration of methanol.
227

t= 0

30'

4h

16h

80 70 60 50 40 30 ppm
Figure 6. BC NMR spectra of BC labeled (30 at %) CH10H adsorbed on levyne
as a function of time at 100 °C.

The so-determined value ofk2 is equal to 0.36 ± 0.06 mor' s·'. From the variation
of k2 as a function of temperature an activation energy of 75 ± 15 kJ mor' was
obtained. This value is much smaller than the theoretical value or computed for
HZSM-5 (145 kJ mort), but it is close to the experimental values determined on
HZSM-5 (108 kJ mor l ) or mordenite (80 kJ mor l ). Note that the activation energy in
the gaseous phase is equal to ca 775 kJ mort.
As far as the apparent equilibrium ratios are concerned, e.g.,
[CH30CH3][H20]l[CH30H] , they are much smaller than the gas phase equilibrium
constants. This can be explained, if the adsorption of both methanol and dimethylether-
water are strong enough:

2CH30l\g) CH30CH3(g) + H2

). ("". CH,O!q 1~ ). ("". CH'OCH')1~ ). ("'. H,o)1


KR,ads,..
228

With
~HpCH3g IH20gICHpHadsP (3)
rHpHgf[cHpcH3adsIH2oadJ
The Kads value computed from the l3 C NMR data is equal to 2.9.10-2• From the
variation of Kads as a function of temperature the enthalpy of adsorption is
endothermal, equal to 52 ± 7 kJ mor l and the entropy of adsorption is equal to 110 ±
15 JK- 1 mOrl. However, the reaction of the methanol to dimethylether in the gaseous
state is exothermal. This difference can be explained if one is taking into account the
strong adsorption of both the reactants and products. Figure 7 illustrates clearly the
correspondence between the various values. Three steps can be distinguished:
1. Adsorption of two methanol molecules forming either a hydrogen bridged complex,
a surface methoxy group and a physisorbed methanol molecule or a pair of
CH30H2+ and CH30- formed on neighbouring Broensted acid and Lewis base pair.
2. Dehydration of the two methanol molecules forming dimethylether and water in the
adsorbed state.
3. Desorption of dimethylether and water from the surface.
Note that the endothermal nature of the reaction is well illustrated in the adsorbed state.

CH 30CH 3 (9)
+ H20 (9)

t~H~'{ads)
l..-_--' . . .... .. - CH 30CH 3 (ads)
2 CH 30H (ads) + H20 (ads)

Figure 7. Thermodynamic model for the dehydration of methanol on levyne.

At higher temperatures aliphatic products start to appear. Propane, n-butane, and


isobutane are detected after 12 h of heating at 250°C. At even higher temperatures
olefmes (propene, butenes, pentenes) are also detected, but only in a small amount. The
major part of the product is composed of aliphatic compounds suggesting the
deposition of coke in the levyne channels.
229

4. Acknowledgement

This work was supported by the Belgian Program PAl P4/10 on Reduced
Dimensionality Systems and OTKA T025248, OTKA F025245 and FKFP-0400/2000.
P. Lentz thanks F .R.LA. for fmancial support.

5. References

1. Engelhardt, G. and Michel, D. (1987) High-Resolution Solid State NMR of Silicates and Zeolites,
Wiley, Chichester.
2. Bell, A.T. and Pines, A. (1994) NMR Techniques in Catalysis, Marcel Dekker, New York.
3. Derouane, E.G., Gilson, J.P. and B.Nagy, J. (1982) In situ characterization of carbonaceous residues
from zeolite catalyzed reactions using high resolution solid state 13C_NMR spectroscopy, Zeolites, 2,
42-46.
4. B.Nagy, J., Engelhardt, G. and Michel, D. (1985) High resolution NMR on adsorbate-adsorbent
systems, Adv. Colloid Interface Sci, 23,67-128.
5. Derouane, E.G. and B. Nagy, J. (1984) Applications of high-resolution J3C-NMR and magic-angle
spinning NMR to reactions on zeolites and oxides, ACS Symposium Series n0248, American Chemical
Society, Washington, DC, 101-126.
6. B.Nagy, J. and Derouane, E.G. (1988) NMR spectroscopy and zeolites chemiStry, ACS Symposium
Series n° 368, American Chemical Society, Washington, DC, 2-32.
7. B.Nagy, J. (1996) The application of 13C-NMR of adsorbed molecules to heterogeneous catalysis,
Bull. Soc. Chim. Belg., 105, 57-66
8. B.Nagy, J. Bodart, P., Collette, H., Gabelica, Z., Nastro, A. and Aiello, R. (1989) Modification of
crystalline and amorphous phases during the synthesis of M-ZSM-5 zeolites (M=Li,Na,K), J.
Chem.Soc., Faraday Trans. J, 85,2749-2769.
9. Gabelica, Z., B.Nagy, J., Bodart, P. and Debras, G. (1984) High resolution solid state MAS IlB-NMR
evidence of boron incorporation in tetrahedral sites of zeolites, Chem. Letters, 1059-1062.
10. Bodart, P., B.Nagy, J., Debras, G., Gabelica, Z. and Jacobs, P.A. (1986) Aluminium siting in
mordenite and dealumination mechanism, 1. Phys. Chem., 90, 5183-5190.
11. Testa, F., Crea, F., Diodati, G.D., Pasqua, L. , Aiello, R. , Terwagne, G., Lentz, P. and B.Nagy, J.
(1999) Synthesis and characterisation of Fe- and [Fe, A1]-MCM-22 zeolites, Microporous Mesoporous
Mater, 30, 187-197.
12. Amoureux, J.-P., Fernandez, C. and Steuemagel, S. (1996) Z-Filtering in MQ MAS NMR,1. Magn.
Reson., A123,1I6-1I8.
13. Lentz, P., Massiot, D. (1996) Sensitivity and Iineshape improvements of MQ-MAS by rotor-
synchronized data-acquisition, 1. Magn. Reson., A122, 240-244.
14. Testa, F. Chiappetta, R., Crea, F., Aiello, R., Fonseca, A. and B.Nagy, J. (1995) Synthesis of
borosilicalite-I with high boron content from fluoride containing media, Stud. Surf. Sci. Catal., 94,
349-356
15. Tuoto, c., Regina, A., B.Nagy, J. and Nastro, A. (1998) Influence of Si(hlA1 20 3 ratio on the synthesis
and physicochemical characteristics of Levyne-type zeolite, Microporous Mesoporous Mater, 20, 247-
257.
16. Lentz, P., B.Nagy, J., Delevoye, L., Dumazey, Y., Fernandez, C., Amoureux, J.-P., Tuoto, C.V. and
Nastro, A. (1999) 20 multiple quantum 21A1 NMR and 29Si NMR characterization oflevyne, Colloids
and Surfaces A: Physicochem. Eng. Aspects, 158, 13-20.
17. Konya, Z., Hannus, J., Kiricsi, 1., Lentz, P. and B.Nagy, J. (1999) In situ MAS 13C-NMR studies of
surface intermediates formed upon interaction of CChF z with NaY-FAU and HZSM-5-MFI zeolites,
Colloids and Surfaces A: Physicochem. Eng. Aspects, 158,35-42.
ORBITAL ORDER AND ORBITAL FLUCTUATIONS IN COLOSSAL
MAGNETO RESISTIVE MANGANITES. AN INVESTIGATION WITH sSMn
AND 139La NMR.

G. PAPAVASSILIOU; M. BELESI, and M. FARDIS


Institute of Materials Science, NCSR Demokritos, 153 10 Aghia Paraskevi,
Athens, Greece

C. DIMITROPOULOS
Institut de Physique Exprimentale, EPFL-PH-Ecublens, IOI5-Lausanne,
Switzerland

M. PATTABIRAMAN, G. RANGARAJAN
Department of Physics, Indian Institute of Technology, Madras, Chennai
600036, India

Abstract

55Mn and 139La NMR techniques are shown to be an excellent tool for detecting changes
in the local electronic properties of colossal magnetoresistance manganites. Based on
this we discuss the possibility of detecting orbital ordering in ferromagnetic
Lal _xCaxMn03 (0.20:::x<0.5). We also provide evidence about the formation of an
unusual kind of phase separation in the low temperature regime of the ferromagnetic
conductive regime of the magnetic phase diagram for x~O.25 . Specifically, we show
that by lowering temperature, an extra NMR signal is fonned in the ferromagnetic
conductive phase, which originates from regions with extremely small magnetic
anisotropy, possibly produced by strong orbital fluctuations.

1. Introduction

Doped mang~nese ~erovs~ites ~ith the ~enera! formula R 1: xDxMn03, where R ~s+ a raJ:
earth metal lIke La' +, Pr3 , Nd' and D IS a dIvalent alkalIne metal such as Ca , Sr- ,
Ba2+, attain currently a lot of theoretical and experimental interest. Although the
existence of these systems was known from the 1950s [1], the research interest
increased only recently, after the first report about the presence of huge negative
magnetoresistance [2], i.e. the decrease of the electrical resistivity upon the application
of an external magnetic field, and the possibility of technological applications. Since
then detailed studies on the transport, magnetic, and structural properties of
RI_xDxMn03 have been conducted, which revealed a very complex magnetic phase
diagram with anti ferromagnetic (AFM) , ferromagnetic (FM), metallic, or charge
231
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science. 231-243.
© 2002 Kluwer Academic Publishers.
232

ordered regions [3]. In case of the Lal.xCaxMn03 family, experiments have shown that
for hole doping 0.2< x<0.5 these materials exhibit a paramagnetic (PM) to FM phase
transition with maximum negative magnetoresistance right below the Curie temperature
Te. At half filling, x=0.5, the FM metallic phase becomes unstable against the formation
of a low temperature insulating AFM phase, whereas for x>O.5 only the AFM phase is
established by cooling. The general framework to describe this behaviour is the double
exchange model (DE) [4], according to which strong Hund's rule coupling enhances the
hopping of eg holes in successive Mn 3+ and Mn4+ ions by establishing FM electron spin
order. This explains in a reasonable way the fact that the FM phase is metallic (FMM),
whereas the PM and AFM phases insulating. It was thus surprising when recent
experiments have shown that for x<0.2 the observed unsaturated magnetization was
produced by a FM and insulating (FMI) phase, and not by a canted AFM phase as until
that time believed, an observation which is in contrast with the DE model. Experiments
in the low doping regime of Lal_xSrxMn03 have shown that most probably such a FMI
phase is associated with orbital ordering (00), and strong reduction of the cooperative
Jahn-Teller lattice distortions [5], [6], [7], whereas the anisotropy in the orbital sector is
considered to favour or disfavour the eg hole mobility in an orbital orientation depended
way. In the more complex case of Lal_xCaxMn03, experiments have shown that for low
doping the formation of an orbitally disordered (OD) FMI phase is also possible [8], [9].
Particularly, sSMn and 139La NMR measurements on Lal_xCaxMn03 [9], [10] have
shown the presence of a transition line T(x) deep into the FMI phase, which has been
associated to 00 effects. At the same time, electron and neutron scattering experiments
on Lao.8SCao.ISMn03 revealed an unconventional layered 00, which is abruptly
enhanced below 70K [11]. In another neutron scattering study of Lal_xCaxMn03, for
doping 0.125< x< 0.20, a bell-shaped transition line TB(x) at around 80K was observed,
which defines a gradual "re-entrant" structural transition with strongly reduced
orthorombicity [12]. Low- T 00 effects in FMI LaI-xCa,,;Mn03 systems were also
confirmed by resistivity measurements under pressure [13].
In the presence of 00, another important issue that raises up is the stability of the
orbital lattice against thermal and quantum fluctuations. Recent theoretical studies have
shown that orbital fluctuations in the fully spin polarized FMM state might explain the
anomalous optical conductivity behaviour in FM compounds of the Lal_ xSrxMn03
family [14]. It is also possible that the quasi-two-dimensional nature of the orbital
fluctuations enables the quantum OD state to persist down to zero temperature, i.e., an
orbitally liquid state is formed, as recently progosed [15].
In the present work, we show that Mn and 139La NMR are able to detect
different Mn electron spin and orbital configurations, as well as the dynamics of such
configurations. Specifically, we provide evidence about the formation of FMI states in
Lal_xCaxMn03 in the doping range 0.ISxSO.2, which at low temperatures exhibit
sufficiently higher local magnetic anisotropy than the FMM phase for x~O.2. This may
be attributed to the 00 nature of the FMI phase in comparison to the orbitally
disordered (OD) FMM phase. Experiments also indicate that by increasing temperature,
the NMR signal from the low temperature 00 FMI phase "wipes out." This effect
accompanies the structural phase transition detected by neutrons [12] . Most remarkably,
233

139La NMR rf enhancement experiments show that for x~0.25, FM clusters with
vanishing small magnetic anisotropy are formed within the "homogeneous" high-T
FMM phase. This feature, which persists even in the AFM spin ordered phase above
x>0.5, is explained as showing the onset of strong orbital fluctuations at low
temperatures.

2. Sample preparation and characterization

Lal_xCaxMn03 samples were prepared by thoroughly mixing high-purity stoichiometric


amounts of LaZ03, MnOz and CaC03. The mixing powders were pelletized and
annealed in air at 1300 to 1400 °C, depending on the composition, for 4 days with
intermediate grinding and reformation into pellets. Finally, they were slowly cooled to
room temperature by turning off the furnace. All samples were then characterized
structurally at room temperature with a D500 Siemens x-ray diffractometer, and
magnetically with a SQUID magnetometer. The obtained crystallographic and magnetic
data were found to be in accordance with literature. Figure I demonstrates
magnetization measurements as a function of temperature for Lal_ xCaxMn03, x = 0.10 ,
0.15 , 0.25 , and 0.33 at a field of H =1000 G. All M vs. T curves are those of typical
FM systems. The corresponding transition temperatures, defined as the inflection

-
x
<II
E
~ 0.5
~

0.0 -h-,....,....,...,...,...,.""T""T"",...........,...,...,.""T""T"";;::;.~~~f;Bl
o 100 200 300
Temperature (K)

Figure 1. Magnetization M vs. T curves for La) .•Ca.Mn03, x::O.1 (6),0.15 (0),
0.25 (®), and 0.33 (0) in an external magnetic field 1000 Gauss.

points of the M vs. T curves, are Tc-169 K, - 180 K, - 240 K, and - 265 K, for x=O.I,
0.15, 0.25, and 0.33 respectively. In addition, we have performed magnetization
234

measurements at very low field, Bext=1O G, in order to minimize the influence of the
external magnetic field on possible 00 effects.

0.12,------------,
x=O.10
0.10

0.08
"C"
~ 0.06
~
~ 0.04

0.02

0.00

2.5
x=O.15

"C" 2.0
~...
~
;, 1.5
E
CD
~ 1.0

0.5

0.0

3.0 x=O.175

2.5
~

~ 2.0
;,
E 1.5
CD
~ 1.0

0.5

0.0 -j-.-'-"'--'~"-'--'-"""':==>"""-r",......j
o 40 80 120 160 200 240 280 320
Temperature(K)

Figure 2. Magnetization M vs. T curves for the Ferromagnetic and insulating systems
La,o,Ca,MnO). x=0.1O. 0.15 . and 0.175 in external magnetic field of 10 G.

Figure 2 shows the M vs. T curves at 10 G for the FM and insulating systems x=O.1O,
0.15, and 0.175. Remarkably, at -80 K. all three curves exhibit a slope change in both
the zero field cooling (ZFC) and field cooling (FC) branches, not shown in the 1000
Gauss. For T -80 K the ZFC branch shows a rapid increase of the magnetization, which
- considering the sudden decrease of 00 above 80 K [91, [11], [13] - may be explained
as showing better alignment of the electron spins into the external magnetic field due to
a decrease of the magnetic anisotropy field HA . On the other hand, in the FC route, the
electron spins are already aligned into the external magnetic field and the slight increase
of the magnetization below 80 K, which is expected to follow the formation 00, as
235

pointed out by Endoh and co-workers [5]. These results are shown below to correlate
nicely with the NMR results.

3. NMR results and discussion·

In contrast to bulk experimental techniques which measure macroscopic properties,


55 MnNMR in zero external magnetic field probes the local magnetic environment of the
resonating Mn nuclei via the hyperfine field, Bhf = (lIfiy)A(S) , where A is the
hyperfine coupling constant and <S> the average Mn electronic spin. According to this
formula 55 Mn NMR is possible to resolve the different Mn charge states, i.e., localized
Mn4+,3+ and intermediate FM metallic valence states. Previous works in manganese
perovskites [9], [16] have shown that at low temperatures NMR signals from Mn4+ and
Mn 3+ charge states have peaks at frequencies -330 MHz and -420 MHz respectively,
whereas signals from FM metallic regions are located at intermediate frequencies, due
to the fast electron transfer between Mn 3+ and Mn4+ ions. By increasing temperature the
signal frequency is expected to decrease, as the order parameter --> 0 on approaching Tc
from below. On the other hand, 139La NMR (S=O), which probes the average spin state
of the surrounding Mn octant through transferred hyperfine interactions, is very
sensitive to changes of the local Mn spin configuration. Since Bhr(La) arises indirectly
from overlapping between the Mn I 3d> and the oxygen 12p> wave functions, in
conjunction with (j bonding of the oxygen with the ISp3> hybrid states of La3+, the 139La
NMR signal should reflect [17] (i) the parallel alignment of the t2g electron spins, and
(ii) possible deformations of the Mn-O-Mn bonding, which alter the hyperfine coupling
constant A. Symmetry arguments also indicate that Bhr(La) should vary from zero value
for collinearly AFM ordered Mn octants, to a maximum value for collinearly FM
ordered Mn octants. In the case of a canted AFM state Bhr(La) is partially cancelled and
therefore the La NMR frequency is expected to be sufficiently lower than in the FM
state.
In the present work we also take advantage of the fact that in PM materials, very strong
NMR signals are produced at very low rf irradiation fields HI , due to coupling of the rf
field with the magnetic moments of the electrons. In this way, the effective irradiation
field at the nuclear sites is sufficiently stronger than the applied HI [18].
Correspondingly, the emitted NMR signal is amplified by the oscillating electron
moments, thus giving rise to the strong NMR signals that characterize FM materials. It
is thus straightforward that any change in the magnetic anisotropy, which will
accompany possible changes in the orbital configuration of the e g electrons, will modify
the response of the electronic moments to the applied external rf field, i.e., will modify
the magnetic anisotropy field H A , and consequently the rf field H I.max that is required to
attain the maximum NMR signal [18]. We stress that in case of NMR signals from
coexisting regions with different local magnetic anisotropy, multiple maxima in the
signal intensity I vs. HI curves are expected.
236

55Mn NMR spectra of Lal.xCaxMn03 in zero external magnetic field were


obtained by using an untuned probe head and a 1 ~sec-1:-1 ~ec spin-echo pulse
sequence with t =3 ~sec, and recording the signal at equidistant irradiation frequencies,
as in previous works [9], [16]. The used rf power level was very small as explained
above, due to the strong enhancement of the applied rf at the position of the resonating
nuclei. 139La NMR spectra in zero external magnetic field were acquired by applying a
two pulse spin-echo technique on a tuned probe, with pulse widths 0.6 IlseC-t-0.6 Ilsec
and t =10 J!sec. In addition, 139La NMR rf enhancement experiments were performed by
recording the signal intensity I as a function of HI at the peak of the spectra. The
obtained curves were found to follow an asymmetric bell-shaped law with maximum at
nyH I=21t13, which allows the calculation of the rf enhancement factor n [18].
Figures 2, 3, and 4 demonstrate 55Mn NMR spectra as a function of temperature for
doping concentrations x=O.l25, 0.175, and 0.33 respectively. According to the magnetic
phase diagram, the first two systems are ferromagnetic insulators, whereas the x=0.33

x=0.125
Mn 3 +

6K

20K

400 550
Frequency (M Hz)

Figure 3. sSMn NMR spectra of Lao.mCao.mMn03 in zero external magnetic field


at various temperatures.

system is ferromagnetic and metallic. For x=0.125 (Figure 3), spectra exhibit a strong
peak at v- 330 MHz, which may be assigned to FM Mn4+ sites, and a broad frequency
237

distribution with a peak at v- 420 MHz, which may be assigned to the Jahn-Teller
distorted Mn 3+ sites. The large difference in the line widths between the Mn4+ and Mn 3+
signals is explainable if we take into consideration that the hyperfine field at the Mn 3+
sites has a strong anisotropic contribution from the spin-dipolar field of the d, x -y
,
orbital states [9]. Consequently, the Mn 3+ NMR frequency depends on the local
environment, thus spreading over a very broad frequency range. By increasing
temperature the overall signal intensity decreases very fast, and the signal disappears at
T B- 80 K, where the re-entrant structural transition occurs [12]. According to recent
NMR relaxation studies [19], the origin of this effect is the extreme enhancement of the
spin-lattice Iff. and spin-spin Iff 2 relaxation rates, due to slowing down of the
fluctuations of the Bhf on approaching T B from below. This makes most of the NMR
signal invisible in the time scale of an NMR experiment. We notice that the wipe out
effect of NMRlNQR signals is produced by coupling of slow fluctuating degrees of
freedom with the nuclear spin. In such case, the relaxation rates Iff •.! become so fast
that most of the signal relaxes before it can be observed [20]. Previously, wipeout
effects have been observed in classical spin glass systems [21] and stripe-ordered
cuprates [20], [22] in the latter case due to the glassy nature of the charge striped phase.

30K
40K
§.OK enm
60K
70K
._-"""'"..............-..-
80 ~"--~--.

100K
120K
140K
280 320 360 400 440
Frequency (MHz)

Figure 4. 55Mn NMR spectra of Lao.825Cao.I75Mn03 in zero external magnetic field


at various temperatures.
238

In the case of x=0.175, the system exhibits a mixed FMI-FMM phase, as


deduced from the 55 Mn NMR sRectra of Figure 4. At low temperatures, spectra show
the coexistence of localized Mn +.3+ peaks and delocalised Mn states at -370-380 MHz
, whereas similarly with the x=0.125 system, by increasing temperature the Mn4 +. 3+
signal intensities reduce rapidly, and finally disappear for T>80 K. The FMM phase
component, does not show any kind of relaxation enhancement through TB, and it's
NMR signal intensity decreases gradually on approaching Tc from below.

190K
210K
250 300 350 400 450
Frequency (MHz)

Figure 5. 5~n NMR spectra of Lao.61CacJ.33Mn03 in zero external magnetic field at


various temperatures.

Finally, in the case of x =0.33, only the FMM signal is detectable, as shown in Figure 5.
This signal varies with temperature in the same way as the FM metallic component for
x=0.175.
In order to examine whether the FMI and FMM phases differ in other electronic
properties, we have applied 139La NMR rf enhancement measurements on
Lal_xCaxMn03, and in the doping range O.lS:xs:O.SS. The sensitivity of the rf
enhancement to the FMI-FMM phase transition is clearly shown in Figures 6a and 6b,
which show 139La NMR spectra and signal intensity I vs. HI curves of Lal _xCaxMn03 as
a function of x at 5 K. The NMR spectra in Figure 6a exhibit a peak at -20 MHz [17],
which corresponds to fully FM polarized Mn octants and remain almost unchanged by
crossing the critical doping at xc-O.20. On the contrary, by increasing x and for x 2:xc,
Hl,max shifts rapidly to lower values. Such a rapid increase of the rf enhancement by
239

12 16 20 24 28
Frequency (MHz)

1.00
~
~ 0.75
>.
~
c:
Q) 0.50
"E
(ij
§, 0.25
U5
0.00..Ly--...---r----.-...---.--,..--I
o 2 4 6
H, (Gauss)
Figure 6. (a) 139La NMR spectra of Lal.xCaxMn03 in zero external magnetic field at
SK. (b) l3~a NMR rf enhancement measurements of Lal.xCaxMn03 at SK.

crossing the phase boundary is explicable if we consider that the 00 FMI phase has a
sufficiently higher magnetic anisotropy than the OD FMM phase.
However, a clear demonstration of the relation between rf enhancement and the local
magnetic anisotropy is given by Lao.825Cao.175Mn03 , which exhibits a mixed FMI and
FMM phase. Figure 7a presents 139La NMR signal intensity I vs. HI curves at various
temperatures. At low temperatures, the maximum of the I vs. HI curves is located at
H I,max=2.4 G, which corresponds to the FMI phase. By increasing temperature the 139La
NMR signal intensity from FMI regions decreases rapidly and disappears at T -80 K, in
accordance with the 55Mn NMR results. Similarly with the 55Mn NMR results, the
origin of the 139La signal wipe out is the dramatic increase of the T 1,2 relaxation rates on
approaching T B from below [19] .
The disappearance of the FMI signal is accompanied by a shift of HI,max at lower values,
whereas at 80K it takes the value 0.3 G, which is characteristic for the FMM phase
component. The system thus comprises of two strongly intermixed phase components
240

with different rf enhancements that reflect the different local electronic properties, and
specifically the different 00 of the two subphases. For reasons of comparison, we have

(a)

T (K)
Figure 7. (a) 139La rf enhancement measurements for Lao.mCao.I75Mn03 as a function of temperature.
(b) The rf field with the maximum signal intensity, HI,max as a function of temperature for x=0.175 (e),
0.20 (0), and 0.25 (0). For reasons of comparison, the M vs. T curve for x=0.175 is also shown in the
same Figure.

also plotted in Figure 7b the HI,max vs. T curves for x=0.20 and 0.25. In both systems
and at low temperatures, HI,max <0.2 G, whereas at high temperatures HI,max increases
slightly in a way similar to x=0.175. The relatively higher HI,max values of the 0.175
system in comparison to those of the x=0.20, and 0.25 systems for T:::::80 K, might
indicate a slight increase of the high-T magnetic anisotropy for x=O.l75 , due to
enhanced orbital correlations.
An unusual effect was observed by performing detailed I vs. HI measurements on
Lal_xCaxMn03 for x=0.20, 0.25, 0.33, and 0.5 as a function of temperature (Figure 8).
At high temperatures, the I vs. HI curves exhibit the FMM maximum, which in the
doping range 0.20<x<0.33 takes values of HI ,max=OA- 1.0 G. Remarkably, for x >0.25
and by lowering temperature, a second maximum appears at HI ,max:::::0.05 G. Such an
extremely low H I.max value is indicative of an NMR signal produced in FM regions,
where the electron spins are almost freely responding to the external rf field. It is also
worthwhile to note that this feature persists even in the AFM phase, as implied by its
appearance in the low temperature regime of x=0.5, and x=0.55 (Figure 6b). In the case
241

of x=O.5, two maxima at Hl.max=O.05 and 2.3 G are observed, whereas neutron
diffraction [23] and resistivity [24] data show the presence of a minority FM phase
down to the lowest measured temperature. However, for the x=0.55 sample, only the
H I,max=O.05 G maximum was detected (Figure 6b), whereas no FM peaks are observed
in the corresponding neutron diffraction pattern performed on the sample used in the
present

l- x=O.25

YJ.~~--to
100 52
'-
200 I-
5
HI (GauSS)

x=O.33
.
l- x=O.50

~~~o ~
,.p~~~100~
nO::'-1-:2::--"3-4-1
5 200 I-
HI (Gauss)
Figure 8. 139 La NMR rf enhancement measurements of Lal.xCaxMn03 at various
temperatures.

work [23]. This indicates that for this doping concentration, the detected 139La NMR
signal is produced in FM clusters with a very short correlation length, which are finely
dispersed within the AFM matrix in agreement with previous electron microscopy
results. It is thus tempting to argue that e g electrons within such FM nanoclusters do not
possess any energetically preferential orbital orientation, thus enabling the obtainment
of extremely high rf enhancement. This picture is in compliance with recent
calculations, which show that the FMM state may vary among different orbital
configurations. Remarkably, the calculated energy difference of these configurations is
very small [25] suggesting the presence of strong orbital fluctuations . In such a case, the
formation of islands of an orbital liquid with extremely low HA cannot be ruled out. The
possibility of orbitally liquid states may explain the appearance of the puzzling I vs. HI
maximum at the extremely low HI ,max=O.05 G. In this scenario, by increasing
temperature, the onset of Jahn-Teller distortions lifts the degeneracy of the various
242

states, and the orbital liquid would freeze in the high temperature orbitally disordered
FMM configuration.
In conclusion, by using 55Mn and 139La NMR techniques, it is possible to
differentiate among two different types of FM configurations, a FM metallic and a FM
insulating that differ significantly in their local magnetic anisotropy. The agreement of
the NMR results with theoretical models including orbital degrees of freedom [25]
implies that most probably 00 is responsible for the difference in the local magnetic
anisotropy. For x>O.25 and at low temperatures, experiments show the formation of
regions with extremely low magnetic anisotropy, presumably due to strong orbital
fluctuations in accordance with theory [25].

References

I. Jonker, G. H., Van Santen, J. H. (1950) Ferromagnetic Compounds of Manganese with Perovskite
Structure, Physica 16.337-349.
2. Von Helmolt . R.. Wecker. J., Holzapfel. B.. Schultz. L. and Samwer. K. (1993) Giant negative
magnetoresistance in perovskitelike La2l3Bal13MnOx ferromagnetic films. Physical Review Letters 71.
2331-2334.
3. Schiffer. P. Ramirez. A. P.• Bao. W.. and Cheong. S -W (1995) Low Temperature Magnetoresistance
and the Magnetic Phase Diagram of Lal.xCaxMn03. Physical Review Letters 75. 3336-3339.
4. Zener. C. (1951) Interaction between the d-Shells in the Transition Metals. II. Ferromagnetic
Compounds of Manganese with Perovskite Structure. Physical Review 2.403-405.
5. Endoh, Y.• Hirota. K.• Ishihara. S .• Okamoto. S .• Murakami. Y.• Nishizawa. A.. Fukuda. T.. Kimura.
H.• Nojiri. H.• Kaneko. K.. Maekawa. S. (1999) Transition between Two Ferromagnetic States Driven
by Orbital Ordering in Lao.88Sro.12Mn03. Physical Review Letters 82.4328-4331.
6. Yamada. Y.. Hino. 0 .. Nodho. S .• Kanao. R. (1996) Polaron Ordering in Low-Doping
La,.xSrxMnOJ. Physical Review Letters 77. 904-907.
7. Uhlenbruck. S .• Teipen. R. Klingeler. R.. Biichner. B.. Friedt. 0 .. Hiicker. M.. Kierspel. H.. Niemoller.
T .• Pinsard. L.. Revcolevschi. A.. Gross. R (1999) Interplay between Charge Order. Magnetism. and
Structure in Lao.mSro mMn03. Physical Review Letters 82, 185-188.
8. Dai. P.• Fernadez-Baca. J. A.. Wakabayashi. N.. Plummer. E. W.• Tomioka. Y.• Tokura. Y. (2000)
Short-Range Polaron Correlations in the Ferromagnetic
Lal_xCaxMn03, Physical Review Letters 85. 2553-2556.
9. Papavassiliou. G .• Fardis. M.• Belesi. M.• Maris. T.G., Kallias. G.• Pissas. M.• Niarchos. D..
Dimitropoulos. C.• Dolinsek. J. (2000) sSMn NMR Investigation of Electronic Phase Separation in Lal_
xCaxMn03 for 0.2~x~O . 5 , Physical Review Letters 84.761-764.
10. Belesi. M.• Papavassiliou. G ..Fardis. M.. Pissas. M.• Wegrowe. J. E.. Dimitropoulos. C. (2001) NMR
as a local probe of magnetic anisotropy: The possibility of orbital ordering and orbital liquid states in
colossal magnetoresistance manganites. Physical Review B 63. 180406R.
II. Lobanov. M. V.• Balagurov • A. M.• Pomjakushin • V. Ju .• Fischer, P.• Gutmann. M.• Abakumov . A.
M.. D·yachenko. O. G.• Antipov . E. V.. Lebedev. O. I.. Van Tendeloo. G. (2000) Structural and
magnetic properties of the colossal magnetoresistance perovskite Lao.8sCao.lsMn03. Physical Review B
61,8941 -8949.
12. Biotteau. G.• Hennion. M.• Moussa. F.• Rodriguez-Carvajal. J .. Pinsard. L.. Revcolevschi. A..
Mukovskii. Y. M.. Shulyatev. D. (2001) Approach to the metal-insulator transition in Lal-xCaxMn03
(O.2~x~0.S): magnetic inhomogeneity and spin-wave anomaly. cond·mattlOlO14 14.
243

13. Okuda, T., Tomioka, Y., Asamitsu, A., Tokura, Y. (2000) Low-temperature properties of Lal_
xCaxMn03 single crystals: Comparison with Lal_xSrxMn03' Physical Review B 61,8009-8015.
14. Okimoto, Y., Katsufuji, T., Ishikawa, T., Urushibara, A., Arima, T., Tokura, Y. (1995) Anomalous
Variation of Optical Spectra with Spin Polarization in Double-Exchange Ferromagnet: Lal.xSrxMn03,
Physical Review Letters 75, 109-112.
15. Tokura, Y., Nagaosa, N. (2000) Orbital Physics in Transition-Metal Oxides, Science 288, 462-467.
16. Allodi, G.. De Renzi. R.. Licci, F., Pieper, M. W. (1998) First Order Nucleation of Charge Ordered
Domains in Lao.sCao.SMn03 Detected by 139La and sSMn NMR, Physical Review Letters 81, 4736-4739.
17. Papavassiliou. G., Fardis, M., Belesi, M., Pissas, M., Panagiotopoulos, 1.. Kallias, G., Niarchos, D.,
Dimitropoulos, c.. Dolinsek, J. (1999) Polarons and phase separation in lanthanum-based manganese
perovskites: A 139La and sSMn NMR study, Physical Review B 59. 6390-6394.
18. Weisman, 1. D.. Swrtzendruber. L. J., Bennet, L. H.. (1973) Techniques of Metal Research. Volume
VI, John Wiley & Sons, New York.
19. Papavassiliou, G., Belesi. M., Fardis. M., Dimitropoulos, C. (2001) 139La NMR investigation of quasi-
static orbital ordering in Lal_xCaxMn03 0.15x50.2, submitted to Physical Review Letters.
20. Cllrro, N. J .. Hammel, P. c.. Suh. B. J., HUcker. M.. BUchner. B.. Arnmerahl, U., Revcolevschi. A.
(2000) Inhomogeneous Low Frequency Spin Dynamics in LaI.6SEuo.2SrO.ISCu04. Physical Review
Letters 85. 642-645.
21. Chen, M. c., Slichter, C. P. (1983) Zero-field NMR study on a spin-glass: Iron-doped 2H-niobium
diselenide, Physical Review B 27. 278-292.
22. Hunt. A. W.• Singer, P. M., Thurber, K. R.. Imai. T. (1999) 63CU NQR Measurement of Stripe Order
Parameter in La2_xSrxCu04 Physical Review Letters 82. 4300-4003.
23. Pissas. M., Kallias. G. (2000). unpublished results.
24. Roy. M.. Mitchell, J. F., Ramirez. A. P., Schiffer P. (1998) Doping-induced transition from double
exchange to charge order in Lal_xCaxMn03 near x =0.50. Physical Review B 58. 5185-5188.
25. Maezono. R.. Ishihara. S., Nagaosa. N. (1998) Orbital polarization in manganese oxides, Physical
Review B 57, R13993-13996.
THE FEATURES OF PFG NMR TECHNIQUE AND SOME METHODICAL
ASPECTS OF ITS APPLICATION

V. D. Skirda
Kazan State University
Kremlevskaja 18, Kazan 420008, Russia,
< Vladimir.Skirda@ksu.ru>

INTRODUCTION
Translational mobility of molecules under thermodynamic equilibrium conditions
(self-diffusion) is a very important characteristic of the physicochemical state of molecular
systems. In view of the rapid development of experimental methods for studying molecular
dynamics, such as neutron scattering, Rayleigh light scattering, and nuclear magnetic
resonance (NMR), studies of diffusion in complex systems acquire special interest. Such
systems include solutions and melts of polymers, systems with chemical interactions,
heterogeneous (e.g., porous) media, and membranes. Recently, the pulsed field gradient
NMR (PFG NMR), which was proposed by Tanner and Stejskal in 1965 [1], is widely
applied for the investigation of transport properties of liquids in various media. Since 1974,
the field of our scientific interests connects directly to the development of PFG NMR and
its applications for the investigation translational mobility in the complicated molecular
systems: polymeric solutions and melts, polymeric networks, biological systems, porous
media. Size limitations of this paper prevent us from giving a detailed description of all
available data on applying this method.

GENERAL PRINCIPLES
The use of PFG NMR teclmiques to study the translational dynamics of molecules is
based on recording the loss of phase coherence of spins as a result of their translational
motion in magnetic field gradients. Information about diffusion processes is obtained by
analysing diffusion damping (decay) A(q,t), or the dependence of the amplitude of the
spin echo signal on magnetic field gradient parameters and time t. The q = (2JlY' rog
value, where r is the gyromagnetic ratio of resonating nuclei, is directly related to
amplitude g and duration 0 of magnetic field gradient" action; this value is an analogue of
the wave vector, for instance, in neutron scattering. It follows that diffusion damping
A(q,t) can be represented by the van Hove correlation function
A(q,t) = Hp(r)p,(r;r',t)exp(i2Jl't](r' -r»drdr' (1)
where per) is the initial density of spins; and P,(r;r',t) is the conventional probability
density or the propagator for observing a spin at a point with radius vector r' at time t if,
at the initial time, the spin had radius vector r. Thanks to the direct relation between
P,(r;r',t) and A(q,t) , the method has a wide scope of applications. For free diffusion in a
one-component system and times different from microscopic, P,(r;",!) has the form of the
Gauss function
245
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 245-254.
© 2002 Kluwer Academic Publishers.
246

- -, ,I):;:
P (r,r
s (41fDi)
1
I 3/2 exp -IF'-
- FI2
4Di
-) (la)

with the root-mean-square deviation ([r'(/) - r(0)]2) :;: 6D,t; parameter Ds is the self-
diffusion coefficient.
It follows that, for a single-phase (from the point of view of NMR) system with a
single D, value and under the condition Og» !go , where go is the constant magnetic
field gradient, diffusion damping for the most widely used stimulated spin echo sequence
(Fig. 1) is written as
A(2" 'I ,g2) :;: A(2" 'I ,O)exp(- q2 Dstd) (2)
For exponential relaxation,

A(2, , 0) = A(O) exp(- 2,


'I' 2 T2
_.2)To
1
(2a)

where A(O) is the initial amplitude of the free induction signal after the first 900 pulse; T2
is the spin-spin relaxation time; ~ is the spin-lattice relaxation time, ' and 'I
are the time
intervals between the first and the second and between the second and the third 90 0
radiofrequency pulses, respectively; 11 is the time interval between gradient pulses; and
td =(11- J I 3) is the diffusion time. For such simple systems, it makes no difference what
pulse sequence parameter is considered variable 11, g or J .
The spatial and time resolution PFG
NMR is directly proportional to the __~~~r--r------------~ 600
amplitude of magnetic field gradient 500
pulse. Generation of large magnetic field Maximum
"'" Gradient
gradients involves serious technical (Tim}
difficulties up to the point (see for 300
instance [2]). Nevertheless, whereas in
the 1980s, a maximum value in pulsed 200

techniques did not exceed 10 TIm, at '00


present, magnetic field gradients of
about 100 Tim and higher have been
obtained in several laboratories (see
F ig.l), which makes it possible to record
translations by less than 1000 A. The
gradient with value 50-200 TlIm we use Figure I . Maximfllll gradiellls achieved ill Ih e iabormOI)' of
long ago. For special probe we are Kazall Swte Ulliversity (V Skirda) ill compariSOIl with results
a selection 0/ other laboratories. (Prof J Karger.
succeeded to achieve the real gradient fromUniversittll Leipzig. Germany; Prof R. Kimmich. UllIversitat
value up to 500T/m. This considerably U/m. Germany: Prof P. Callaghall. Massey Ulliversity. New
broadens the scope of the method and its Zealalld: alld Prof H. SillesclI. Ulliversillil Mainz. GermanyJ.
applications.

MULTIPHASE SYSTEMS
Simple objects for which equation (2) is exact are rather an exception than a rule. But
there is a huge number of multiphase (from the point of view of NMR) molecular systems
(here and throughout, the term phase does not necessarily refer to a thermodynamic phase).
In such systems characterised by a set of relaxation times and self-diffusion coefficients if
the diffusion propagator of each phase obeys equation (la), diffusion damping can be
247
written as the sum of exponential terms

A(g2) = ±P;exp(-l DsJd)' (3)

where apparent population P; is given by

Pi, = Pi exp( -2r - )/~


- - r, (2r r,)
L.. Pi exp - - - - (3a)
~i'T2i T2i ~i
N is the number of phases; and Pi is the true population (fraction) of resonating nuclei
characterised by ~i' Tu , and Dsi • Clearly, experimental conditions should be such that
time parameters rand r, remain constant. This can be achieved if diffusion damping is
induced by varying magnetic field gradient or, less often, gradient pulse width 8 . For this
reason, the term diffusion damping will hereafter be understood to imply the dependence of
the amplitude of the spin echo signal on the magnetic field gradient, (A(g2»). It is assumed
that the other parameters are kept constant.
Often, because of instrumental limitations and complexity of objects, an averaged
characteristic (mean self-diffusion coefficient 75,) is analysed. This characteristic is
unambiguously determined from the initial slope of diffusion damping; for (3), it equals
75, = LP;D'i (4)
We again stress that, generally, fractions in (4) are apparent values depending on the
time parameters of the experiment according to (3a).
On the whole, the possibility of obtaining data on the behaviour of all molecular
components of a multiphase system is the remarkable property of the method making it
exceedingly informative. Nevertheless, equation (3a) shows that the results should always
be interpreted with caution. Indeed, for a normal sequence of stimulated spin echo pulses,
the form of diffusion damping is a function of the time of diffusion, because a change td in
(3) implies a change in interval r, in (3a), which results in an apparent dependence of
populations P; on td • One of the methods for taking into account this undesirable effect
involves renormalization of diffusion measurement results with the use of additional
nuclear relaxation measurements. More reliable information on the time dependence of
translational dynamics of molecules can be obtained by using a special five-pulse sequence;
this technique makes it possible to vary diffusion time at constant time parameters of NMR
relaxation. The problems of taking into account nuclear relaxation effects are considered in
detail in [3).
For multiphase systems, the question of the nature of separate phases (components)
and the interaction between them is always topical. Often, equilibrium between phases is
controlled by exchange. A study of exchange characteristics is a fundamental problem of
physical chemistry.

MULTIPHASE SYSTEMS WITH EXCHANGE


For the first time, the problem of measuring self-diffusion coefficients in the presence
of exchange between phases was considered in [4] for the example of a two-phase system,
on the assumption of exponential distribution functions for lifetimes in phases. A more
general problem was considered in [5,6). The results for slow and fast exchange conditions
coincide. Under slow exchange conditions, diffusion damping is described by (3), and for
fast exchange, A(g2) = exp(- q2 D,td)'
248
Intermediate exchange is most interesting. For instance, even for a two-phase system
(phases a and b), diffusion damping is described by a continuous spectrum of self-
diffusion coefficients. The most important result is that the spectrum is unambiguously
bounded by Dro and D,b values and populations Pa(t,,) and Pb(/,,) of the components are
(I ) '<I
decreasing functions of diffusion time I" :
Pa.h " = 1- flfla.b(Ta.b)dTa .b ,
Pa.b(/" =0) 0

where Pa.h(/" = 0) is the true population of phase a or b, and the integral is the
probability for a species to leave the corresponding phase at least once for time I" .
To summarise, an analysis of Pa.h(t,,) dependence provides the possibility in principle
of determining distribution function lfI(r) for lifetimes in phases; this function makes it
possible to determine easily the mean lifetime in a phase. An important criterion of
exchange is independence of the mean self-diffusion coefficient Ds = PP'i fromL
diffusion time. This method, different from the classic approach [4], was used to gain
insight into the reasons why the form of diffusion damping in melts and solutions of
polymers and binary mixtures of polymer molecules with different molecular weights
depended on diffusion time [5]. The results, which provided a fairly solid basis for the
claim that polymer systems contained dynamic macromolecular clusters, were used to
estimate cluster lifetimes. The concept of the existence of dynamic clusters as maximal
kinetic units in self-diffusion of macromolecules is of importance for understanding the
special features of the translational dynamics of polymer chains and should be considered a
contribution to theory.
Very interesting results were reported for aqueous solutions of saccharides [6], in
which proton exchange was for the first time detected by the PFG NMR method. Diffusion
experimental data were used to determine the mean lifetimes of protons in saccharide
hydroxyls as a 1fI( T) function.
In several works (e.g., see [7]), anomalous behaviour of self-diffusion coefficients of
liquids partially filling a porous space was observed; the suggestion was made that this
behaviour was due to fast (on the NMR scale) exchange between liquids and their saturated
vapours. This exchange was detected [8] by PFG NMR and the distribution function for
lifetimes in the liquid phase was obtained. The form of this function was shown to be very
sensitive to the strength of interaction between the liquid and the surface. Under the
conditions of incomplete filling of a porous medium by a liquid, this interaction determined
the character of the distribution of the liquid over the porous space.

SELF-DIFFUSION OF MACROMOLECULES IN SOLUTIONS AND MELTS


The Rouse model describes the dynamics of linear macromolecules with molecular weights
below the critical value (Mc). At higher molecular weights, the state of molecules
characterised by the so-called entanglement effect is treated in terms of dynamic scaling
and the phenomenological reptation model [9]. Its predictions concerning the molecular
weight dependence of the self-diffusion coefficient (D oc M- 2 ) are in satisfactory
agreement with the majority of experimental data. An analysis of arguments for and against
the reptation model is beyond the scope of this work. We will only mention the most
important results obtained by PFG NMR. First, consider the concentration dependencies of
the self-diffusion coefficients of flexible-chain macromolecules. It was shown for more
than 20 different polymer-solvent systems [10] that the translational dynamics of
macromolecules obeys some universal law f(j{;), which is invariant under temperature,
249

polymer molecular weight, and structure


variations and solvent changes from ideal ()-
solvent) to "good." It was found that,
preliminarily, the initial concentration
dependence of the mean self-diffusion
Cl
0.01 coefficient D,(rP) should be normalised with
the use of the concentration dependence of
the local dynamic characteristics L(rP) of the
polymer chain. Such a normalisation can be
16-4
performed with nuclear magnetic relaxation
data obtained in independent measurements
0.01 0.1 10 for each polymer-solvent system.
In the high-temperature approximation,
L(rP) = r.. (rP)/r.. (0), and the self-diffusion
Figure 2. Universal concentration dependence
coefficient normalised to the local
- - = f( -;::::
D; (rP) rP) obtained in flO} based on characteristics of the polymer chain can
D,(O) rP therefore be written as
more than 130 initial D,(rP) dependence'sfor D;(tp) = D,(tp)L- 1 (rP) =D,(tp)r..(O)/r..(tp) ,
macromolecules in 20 different polymer-solvent where r.. (rP) and r.. (0) are the spin-lattice
systems relaxation times for polymer protons at the
current concentration rP and in the rP ~ 0
0'(",)/0(0)
limit, respectively. Remarkable results of
such a normalisation are plotted in Fig. 2.
Note that, for each system, only one critical
concentration depending on the molecular
weight of the polymer as 'i(M) oc M-(2-jJ)/3
had to be introduced. An estimate of f3 can
• 3
. 0- .. 4 also be made from the dependence of
-5
10 lim D,(M) oc M-P in dilute solutions.
10' ; -> 0

The function implies the


Figure 3. Concentration dependence's
existence of very general laws governing the
D; (tp) for
the fifth (1), sixth (2), and translational dynamics of flexible-chain
seventh (3) generations of dendrimer in macromolecules in solutions of polymers. All
solutions with deuterium chloroform, were special features of a polymer-solvent system
obtained by II normalize »-procedure oj _ __
experim ental concentration dependence with are determined by the D, (0) = lim D, (rP) , rP ,
using L-functions for each dendrimer ¢ .... o
generation, accordingly and »generalized" and f3 parameters values and on the form of
concentration dependence for globular
proteins in aqueous suspensions (4) in the L(rP) function.
comparison with data for a linear polymers Additionally the generalized
concentration dependence of D;(tp) for macromolecules ofpolyallylcarbosilane dendrimer
solutions with deuterium chloroform [II] and for globular proteins in aqueous
suspension [12] are shown in Fig. 3, too. It is shown that within the limits of investigated
macromolecule concentration range the 'generalized' concentration dependence for
polyallylcarbosilane dendrimers coincides with 'generalized' con centration dependence for
250

globular protein in aqueous suspensions. The difference of the curve shapes for
"generalized" concentration dependence of D;(rp) for dendrimers and universal
dependence of D; (rp) for linear flexible polymers gives reasons to characterize dendritic
macromolecule as a system, which does not significantly change itself conformation and
has mechanisms of intermolecular interactions similar to globular proteins.
Concerning a discussion about the reptation model, some results of the
investigation of self-diffusion process of polymer chains in a medium with geometric
restrictions (for example poly(ethylene glycol) macromolecules with different molecular
masses, confined in pores of glass
"Vycor") are presented (Fig. 4).
(The diameter of pores was about
IE·1O 40 AO, which corresponds to
"tube" diameter in the reptation
I~.II
model). The curve I form is
typical for polymer melts.
co " Experimental data obtained for
... · Il
poly( ethylene glycol)
Me
macromolecules in porous glass
11::-1 .,1
I (curve 2) formally can be
described by the umque
1000 10000
dependence of the type
D, oc M- '4 , that is in contrast to
Figure 4. The molecular-mass dependencies of averaged self-
diffusion coefficients of poly(ethylene expected D, oc M- 2 from the
glycol)
reptation model. This fact can be
macromolecules in bulk (I) and confined in porous glass
., Vycor" (2). The asymptotic dependencies with corresponding
interpreted as a consequence of
power laws are shown by solid lines. The temperature of
excluding in porous medium some
measurements was +60°C. The diffusion time was 0.2 sec.
of macromolecular dynamics
factor, which is neglected in reptation model. By our opinion, this factor is related with
effect of the intermacromolecular linkage formations (quasi-crosslinks), which provide the
realization of more isotropic mechanism of the polymer chain dynamics in the comparing
with mechanism of reptation. Moreover, as it was shown in our previous studies, these
intermolecular linkages (entanglements) can result to formation of the stochastic
macromolecular groups (clusters) [3,5].

TRANS LA TIONAL DYNAMICS OF MACROMOLECULES FORMING THREE-


DIMENSIONAL NETWORKS
A three-dimensional network is an analogue of a macromolecule with an infinite
molecular weight. For this reason, translational mobility of macromolecule elements is of
interest rather than the whole macromolecule. Clearly, this translational mobility essentially
differs from the dynamics of a freely diffusing particle and can be characterized by some
effective self-diffusion coefficient depending on diffusion time td ,
D *, (t D) = < r2 (t d ) >/6td • For polymeric networks, of greatest interest are studies at long
td times, when the mean-square displacement is constant, < r\td ) >= const , and directly
related to network parameters such as the molecular weight (Mee) of chains between cross-
links, the conformation of these chains, and the functional groups that form cross-links. The
D*,(tD) dependence obeys the law of fully constrained diffusion, D*,(tD)oct d - ', and
D*, (tD) is temperature-independent. Examples of the first results obtained for such
251

systems are shown in Fig. 5, where the


1E-13
D*., (tD) dependencies observed in
N!{!
E studies on radiationally cross-linked
;,
a t -1
polybutadienes [13] are plotted; the
1E-14
...-- d samples had different densities of
cross-links and were swelled in
benzene to equilibrium. The

0
1E· 15 proportionality law D*s(ID)octd-1 is
.. -2 v
o -3
obeyed in a wide range of td
• -4 0 variations, and the expected
dependence on the density of cross-
0.01 0.1
links, < r2(td) >oc M ee , IS indeed
td·s
observed. All these results show that
Figure 5. Dependencies D *., (td ) for radiation-cross- the method of NMR with pulsed
linked polybutadienes with Mee =.9.J(t (1). (2) 8.8 ·J(1. (3) magnetic field gradient is an effective
tool for studying gel formation
6.8.J(1 and (4) 3.8HI under the conditions of equilibrium
swelling in benzene. T = 303-333 K.
processes leading to three-dimensional
macromolecular networks.

TRANSLATIONAL DYNAMICS OF MOLECULES AND GEL FORMATION


We studied the translational mobility of molecules in gels for the example [14] of
three systems (gelatine-water, agarose-water, and cellulose triacetate-benzyl alcohol). All
of them-formed thermoreversible gels upon cooling. The diffusion decay can be described
by the equation
2 2, 'I 2 2, 'I "'f 2
A(g ) "" Pa exp(-----q tdDsa) + Ph exp(----) P.. (D,b)exp(-q IdDsb)dDsb +
T2a T.,a T2b T.,b 0
+ P exp(-~-!L-q2t D* . (I ))+ P exp(-~-.!i-q2t D*. (I ))
e T2e ~
Ie
d K d • T2. ~
Ie
d. d
Phase a comprises solvent molecules. Unlike relaxation time T2a , self-diffusion
coefficient Dsa usually does not change substantially as a result of gel formation. This
observation conforms to the concept of free and bound states of solvent molecules and fast
exchange between them. The fraction of the bound solvent is small. The other phases
comprise polymer molecules. The parameters with subscript b characterise free (not
incorporated into the gel network) macromolecules, those with subscript c refer to
macromolecule fragments between network nodes, and the parameters labelled e describe
macromolecule fragments situated in network nodes. Note in conclusion that the
possibilities of experimentally recording diffusion damping for phase e are often limited
because the relaxation times T2e are too short. The coefficients Dm and D,b were
independent of diffusion time td , whereas the dependencies of D *se (Id ) and D *,. (td )
corresponded with the model of completely constrained diffusion - D* oc 1;1.
The region of the onset of gel formation manifests itself by nonzero Pe and P. values
and a decrease in Pb . A decrease in Ph is as a rule characterised by an anomalous increase
in the mean D,b value and a decrease in the width of the spectrum, even during further
252

cooling. These observations evidence


the formation of a broad particle-size
1,0
spectrum of clusters of
macromolecules under the conditions
close to gel formation; they also
show that the spatial structure of gels
is predominantly formed from
maximum-size clusters. The values
of p"" = Ph /(Ph + Pc + Pe) and
Pge' = (Pc + Pe) /(Pb + Pc + Pe) are
the fractions of the sol and gel
phases, respectively.
I f I The Pge/(t) kinetic dependencies
4 6 8
describe gel formation under
t1(}3, rek
isothermal conditions and obey the
Kolmogorov-Avraami equation
Figure 6. Fraction p gel as a function of gel formation
duration for aqueous solutions of gelatine with Pge' = P;, [1- exp( -kilt")],
M,,=9·10' at gelatine concentrations (1, 2) 5 and (3-5)
10 wt % and temperatures (1,3) 303, (2, 4) 300, and (5) where P;e, = lim Pgel (t), and k and
1-400
265
n are the rate and acceleration
constants. The P;, value heavily
lOll
I, ~'I-I~
l' :-a~- ~ depends on polymer concentration

~~~
UI
Or and molecular weight, temperature,
E
etc., but virtually never reaches one.
OW" Additional conclusions on network
rigidity can be drawn from a
1012 A-A-~~~-~
comparison of T2h and T 2c ' Of the
~'\
~:-.
(~~
2 "
systems mentioned above, the least
rigid network is formed in gelatine
:.~
-.-.+-l.,¢ gels because of intramolecular non-
1013 Ql uniformity of gelatine
3,0 3,1 3,2 3,3 3,4 macromolecules imposing
T-l'l(ll(~l) limitations on the size of network
nodes. Estimates show that
Figure 7. Temperature dependencies of (1. 1') mean
D *,c and D *\e values have usually
self-diffusion coefficient Dsb and (2, 2') sol phase
the same order of magnitude. In
fraction (P.w' = 1- Pge' ) in a 10 wt % aqueous chemically cross-linked gels, phase's
solution of gelatine with M,,=9·104• Curves 1 and 2 were c and e are as a rule
obtained by increasing the temperature ofgel formation, indistinguishable. The Pgel(t) kinetic
and curves l' and 2', by decreasing this temperature
dependencies are shown in Fig. 6 for
the example of gelatine gels, and typical temperature dependencies of D,h and Pge' , having
the form of hysteresis curves, are represented in Fig. 7.
Notes that NMR with pulsed magnetic field gradient is one of the few methods for
studying the structure of gels without gel destruction. For instance, translational mobility of
linear macromolecules of poly( ethylene glycol) (PEG) within a weakly cross-linked
poly(methacrylic acid) (PMAA) hydrogel were investigated [15] by PFG NMR method in
253

order to reveal the effect produced by PMAAIPEG complex formation. It was found that
inside the collapsed gel a part of PEG macromolecules has the diffusion characteristics
coupled with those for the network chains. This evidences the formation of interpolymer
complex, as a result of which linear molecules acquire dynamic properties of the network
chains. Another part of PEG macromolecules inside the collapsed gel is characterized by
free diffusion which implies that not all the PEG macromolecules, absorbed by the gel, are
included in the complex with PMAA. As opposite to the collapsed gel, in the swollen gel
(at concentrations of linear polymer higher than 5 wt. %) the self-diffusion coefficient of all
PEG molecules is independent on the diffusion time, which indicates to the absence of
interpolymer complex (or at least to the fact that its lifetime is short enough).

SELF-DIFFUSION IN POROUS SYSTEMS


The behavior of the mean self-diffusion coefficient of a liquid in a porous structure is
indicative of the existence of three diffusion time regions:
i) Short-time region, when root-me an-square displacements of molecules of a liquid
are smaller than linear pore dimensions, and there is no substantial effect of pore walls on
the observed self-diffusion coefficient; that is, D* "" Do, where Do is the self-diffusion
coefficient of bulk liquid.
ii) Intermediate diffusion time region, when root-mean-square displacements of
molecules of a liquid are comparable with linear pore dimensions, and effective coefficient
D * depends on diffusion time.
iii) Long diffusion time region. The behavior of D * is then determined by the
structure of the porous space.
For closed pores and long diffusion times, completely constrained diffusion with
D* = De!! (I) oc rl can be observed. Experimental data can then be used to determine linear
constraint dimensions by the Einstein equation Rp 2 "" 6D eff (t)1 .
For connected pores and long diffusion times, root-mean-square displacements of
molecules are larger than pore dimensions, and motion of the liquid becomes averaged over
the space of the system. For porous systems with a random structure, diffusion damping is,
as a rule, exponential and characterized by a diffusion-time-independent effective self-
diffusion coefficient D* = D", .
It follows that the behavior of the mean self-diffusion coefficient of molecules in a
system of constraints is determined by three terms, viz., the self-diffusion coefficient of the
pure liquid, the effect of constraints, and the permeability effect. According to [8,16], a
correct separation of these effects we can be achieved with scaling equations for
Deff(/)=(D*(t)-D",)*( Do ).
Do-D",
The D~ff (I) dependency can be used to determine information about constraint
dimensions (porous medium characteristics) by analysing the overall time dependence of
the self-diffusion coefficient including the three principal time regions.
The self-diffusion of the alkanes molecules with molecular weights from 86 to 212
and water molecules, confined in the controlled pore glass "Vycor" with average pore
diameter 40 A0 are measured within 3 decimal orders of diffusion time td interval from 3
ms to 1000 ms by PFG NMR method at 300 C. We reliably observe the dependence
Ds(td)-tllt is found for all the studied liquids. These results allow us to suggest that the
fractal structure of porous glass "Vycor" with average pore diameter 40 A0 is not displayed
in the spatial scales from 0.1 j.J m to 24 j.J m. The second important result is the observation
254
of power law dependence of Ds on the liquid molecule size. It is shown, that taking into
account of this result is necessary for the estimation of tortuosity ~ as well as the form-
factor F of the porous media according to the self-diffusion data. For the system "Vycor"-
water we observe the peculiarities which can be explained only by structure ordering
caused by the strong hydrogen bonds. We suggest that our results can be used to study the
association phenomenon of low molecular weight liquids by NMR PFG method.

ACKNOWLEDGMENTS
This work was carried out within the framework of projects of the RFBR (N 00-03-33071)
and CRDF (REC 007)

REFERENCES
I. Steiskal, E.D. and Tanner, J.E. (1965) Self-diffusion measurements: spin-echoes in presens of a time
dependent field gradient, J. Chem. Phys., 42, 1,288-292.
2. Galvosas, P., Stallmach, F., Seiffert, G., Karger, J., Kaess, U., and Majer, U. (2001) Generation and
Application of Ultra-High-Intensity Magnetic Field Gradient Pulses for NMR Spectroscopy, Journal of
Magnetic Resonance. 3, 1-9
3. Maklakov, A.I., Skirda, V.D., and Fatkullin, N.F. (1990) Self-Diffusion in Polymer Systems, in N.P.
Cheremisinoff(eds.), Encyclopedia ofFluid Mechanics V.9. Polymer Flow Engineering, Gulf-Publishing.
CO.Houston, London, Paris, Zurich, Tokio, pp.705-745.
4. Karger, J. (1969) Zur massbarkeit von diffusionkoeffizienten in zweiphase system mit hilfe der methode
der gepulsten feldgradienten, Ann. Physik, 24,1-7.
5. Skirda V.D. (1992) Self-diffusion in polymer systems, Doctorate (Phys.) Dissertation. Kazan State Univ.,
Kazan .
6. Sevryugin, VA, Skirda, V.D., and Skirda, M.V., (1998) Exchange processes in water solutions of
saccharides, Russia~ Journal of Physical Chemistry, 72, 869-874
7. Dvoyashkin, N .K., Skirda, V.D., Maklakov, A.I., Belousova, M.V., and Valiullin, R.R. (1991) Peculiarities
of self-diffusion of alkane molecules in kaolinite, Appl. Magn. Reson., 2, 83-91 .
8. Valiullin, RR, Skirda, V.D., Stapf, S., and Kimmich, R. (1997) Molecular exchange processes in partially
filled porous glass as seen with NMR ditTusometry, Phys. Rev. E: 55, 3, 2664 - 2671.
9. De Gennes, P.-G., (1971) Reptation ofa Polymer Chain in the Presence of Fixed Obstacles, J.Chem. Phys..
55,572-579.
10. Skirda,V.D., Sundukov,V.I., Maklakov,A.I., Zgadzai, O.E., Gafurov, I.R., and Ya'siljev, G.1. (1988) On
the generalized concentration and molecular mass dependens of macromolecular self-diffusion in polumer
solutions, Polymer, 29, 1294 -1300.
II. Sagidullin, A.I., Krykin, M.A., Ozerin, A.N., Muzafarov, A.M., Skirda, V.D. (2001) "Generalized"
Concentration Dependence of Self-Diffusion Coefficients for Macromolecules in Solutions of
Polyallylcarbosilane Dendrimer, Abstr. at N'h Europ. Symp. on Polymers Spectroscopy. Dresden.
p.129
12. Nesmelova, I. V., Skirda, V.D., Fedotov V.D. (in press) Generalized concentration dependence of globular
protein self-diffusion coefficients in aqueous solutions, Biopolymers
13. Skirda, V.D., Doroginizkij, M.M., Sundukov, V.I., Maklakov, A.I., Fleischer, G., Housler, K.G., and
Straube, E. (1998) Detection of spatial fluctuations of segments in swollen polybutadiene networks by
nuclear magnetic resonance pulsed field gradient technique, Macromol. Chem .. Rapid. Commun., 9, 603-
607.
14. Gaphurov I.R, Skirda V.D., Maklakov A.I., Perevezentseva S.P., Zimkin E.A. (1989) The study of self-
diffusion in aqueous solutions of gelatins and process of gel formation, Vysolwmol. Soed. (in russian). 31A.
269-275.

15. Skirda, V.D., Aslanyan, I.Yu., Philippova, O.E., Karybiants, N.S., Khokhlov A.R. (1999) Investigation of
translational motion ofpoly(ethylene glycol) macromolecules in poly(methacrylic acid) hydrogels,
Macromol. Chem. Phys.. 200. 2152-2159.
16. Valiullin, R, Skirda V. (2001) Time dependent self-diffusion coefficient of molecules in porous media, 1.
Chem. Phys .. 114,452-458.
129Xe NMR OF ADSORBED XENON USED AS A PROBE TO STUDY
MICROPOROUS SOLIDS

M.-A. SPRINGUEL-HUET
Laboratoire des Systemes Inter/adaux aI 'Echelle Nanometrique,
CNRS-FRE2312,
Universite P. et M Curie, 4 place Jussieu, 75252 Paris Cedex 05,
France

1. Introduction

The use of 12~e NMR to study the properties of porous solids started at the end
of the 70s. Fraissard's group began to study zeolites [1], while Ripmeester
studied clathrate hydrates [2]. Since that time a great number of papers (more
than 300) dealing with zeolites and other solids, and several reviews [3-9] have
been published.
Xenon is an ideal probe because it is an inert gas, with a large spherical
electron cloud. From the NMR point of view, the I~e isotope has a spin of one-
half, its natural abundance is 26 % and its sensitivity of detection relative to the
proton is 10.2• The xenon atom diameter is 4.4 $., which allows it to access sites
of catalytic interest.
The high polarizability of the electron cloud makes the chemical shift an
especially sensitive measure of local atomic interactions. This is reflected in a
wide chemical shift range (about 1000 ppm) for physical interactions of Xe. The
whole range goes from -40 ppm for Xe adsorbed in AgX zeolite to about 7500
ppm in the chemical compound, xeO/ [6].
The interest of I~e NMR has grown significantly with the hyper-
polarization technique. Optical pumping from an alkaline element being itself
electronically excited under laser irradiation dramatically enhances the I~e
nuclear magnetization [6]. Solids with low surface area can be studied. The highly
polarized Xe can also be used to magnetize other nuclei of low sensitivity or to
perform NMR imaging. This latter technique is mainly applied in biology and in
the medical field. The hyperpolarization technique and its applications are
discussed elsewhere in this book.
255
1. Fraissard aruJ o. Lapina (eds.), Magnetic Resonance in Colloid aruJ Interface Science, 255-266.
© 2002 Kluwer Academic Publishers.
256
2. Basic ideas ofXe NMR

Fraissard and co-workers have characterized the shift of adsorbed Xe as a


sum of various interactions:
(1)
8s is due to the interaction ofaXe atom with the pore surface; 8xe to Xe-Xe
interactions inside the pores, and depends on the Xe loading. 8SAS arises when
there are strong adsorption sites (metal particles, highly charged cations ... ). 8E
and ~ come from the electric and magnetic field, respectively, created by highly
charged, possibly paramagnetic cations. Xe atoms are very mobile in the pore
structure and average these interactions more or less; the spectrum usually
consists of a single line.
An empirical relationship (Figure 1) has been obtained between the term 8s
and the pore structure characterized by the mean free path, f, ofaXe atom which
is defined as the average distance travelled by Xe between two successive
collisions against the pore wall [7]. The hyperbolic shape of the curve has been
explained using simple models based on calculations of Lennard-Jones potentials
between Xe and the oxygen atoms of the pore surface [10,11]. In the region of
low f values the points are dispersed, since Xe is also sensitive to the chemical
properties of the surface, e. g. the SilAI ratio in case of zeolites.

i lSO

--.
t
00
I:

~.-
~ 150
~
8
=
; 100

1 2 3 4 5 6 7 8
o
Mean Free Path 1 (A)

Figure 1. Empirical relationship between the term 0 S


and the mean free path of xenon adsorbed in zeolites
and related materials with low SiiAl ratio (D), high
SiIAI ratio ( 0 ) (from ref. 7, with permission)
257

The intra- and intercrystallite diffusion of xenon may also play a role and
must be taken into account. If, on the NMR time scale (typically of the order of
ms), Xe diffuses within several crystallites, it reports not only on its sampling of
the inside environment of a single crystal but also of that of many crystallites as
well as that of the interparticle space. The chemical shift is then averaged and is
not truly characteristic of the solid studied. The influence of intercrystallite
diffusion depends on the size and morphology of the crystallites, the pore size, the
crystallite packing, the Xe gas pressure and the temperature [12-14].
On the contrary, in case ofNaA zeolite, even intracrystallite diffusion is slow.
The spectrum is then resolved according to the different Xe populations in the
NaA cages [15]. The XeINaA system constitutes an excellent model for
theoretical calculations (Monte Carlo or dynainic simulations). Many papers
dealing with the distribution of xenon among the cages under various conditions,
possibly in the presence of another adsorbed gas, as well as the dynamics of the
Xe atom motion between the cages have been published [16-20).

3
4
2
5

I I I I I I I , ! , I I , I

300 200 100


Chemical Shift (ppm)

Figure 2. 12~e NMR spectrum ofXe adsorbed in NaA


zeolite. Each peak corresponds to a defmite number of
Xe atoms (1, 2, 3, 4, and 5) in the cavities (from ref.
15, with permission)

If the influence of intercrystallite diffusion can be neglected (large crystals,


slow intracrystallite diffusion), the number of signals is determined by the number
of different types of voids, provided that they are accessible to Xe and there is no
exchange between them. There are several examples, such as ferrierite, mordenite
258

and rho. The spectra depend on the nature of the cation and on the temperature
[21-23].

3. General Applications

12'>xe NMR was first used to study zeolites and c1athrates, but it has been rapidly
extended to other porous solids like clays [24-26], amorphous silicas [27,28],
polymers [29-37], carbons [38-42] and semi-conductors [43] The studies mainly
concern the pore structure, the surface properties, the nature and distribution of
extra-framework species.

3.1. PORE STRUCTURE

As has been seen in Section 2, there is an empirical relationship between the


chemical shift and the pore size. It has been used to estimate the pore size of
unknown zeolite structures, of clays and other solids.
The presence of defects, after dealumination of zeolites, for example, or of
structure intergrowths can be detected and quantified under favourable conditions
[44]. Polymer blends can be treated in the same way [33,34].
12'>x:e NMR is very useful for the study of pure polymers [45,46]. The
chemical shift is characteristic of a given polymer and ranges over about 150-250
ppm. The chemical shift decreases with increasing temperature, the plot exhibiting
a change of slope at the glass transition temperature, T g [29,47,48]. The linewidth
dramatically increases below T g, indicating that motions of the polymer chains are
suppressed. Xe is then trapped in various sites. Xe diffusion coefficients are often
determined by Pulsed Field Gradient NMR experiments or 20 exchange NMR in
the case of blends and interpreted in terms of mean pore size or size of
monodomains for blends. IH_ l 2'>x:e NMR polarization transfer experiments give
information on chain mobility.

3.2. SURF ACE PROPERTIES

The Xe chemical shift depends on the chemical properties of the pore surface.
Obviously, the sensitivity of Xe to the nature of the surface is all the more
important as the pore is small. For example, we have shown that the term Os
increases with the AI content of the ZSM-5 framework with a marked
discontinuity for a concentration of 2 AI atoms/u.c. [49]. Similar results are
obtained with borosilicalites.
The grafting or deposition of chemical species can also be studied, such as
polymer grafting on silica [50] or heteropolyacid supported on silica for examples
[51].
259


••I
115
~~
•:.-..
.. -
........cr.-
-~
S 110
Q,
Q,
••
... I
f.Q -.~"'.I

100
1 1 3 4
A1Iu.c.

Figure 3. Os variation versus framework Al content:


(0) NaZSM-5 and (A) NaZSM-ll (from ref.. 49, with
permission)

3.3. CATIONS

F or most zeolites, in the acid or sodium form, the influence of cations on the 12'0ce
signal is negligible at room temperature. The chemical shift is given by the two
terms Os and OXe of equation 1. However, when the cations are more highly
charged, the 0 = f([Xe]) curves shows an increase as [Xe] tends to zero,
characteristic of the OSAS term. There is a specific interaction with the cations
responsible for a higher chemical shift especially at low loading, when the Xe
atoms interact mainly with these cations (Figure 4a). This interaction has been
described by simple models ranging from high polarization of the electron cloud
[52] to an electron transfer from the xenon atom to the cation [53].
Many studies of X and Y zeolites containing cations like Mi+, Ca2+, Zn2+,
Cd + and even rare earth cations (y3+, La3+, Ce3+) have been carried out with
2

different degrees of cation exchange, and thermal treatment of the zeolites. Since
Xe can only interact with cations located in the faujasite supercages, the
dependence of cation location on the exchange rate or of their migration on their
hydration state has been studied [54-57]. The migration of cations between
crystals during thermal treatment has also been observed in the case of a mixture
of dehydrated RbNaX and NaY zeolites [58]. The problem is more difficult for
paramagnetic cations (Ne+, Co2+). One must consider the ~ term, which may be
very large, leading to 0 values over 1000 ppm [59,60].
260
Another interesting example is the unusual behaviour of Ag+ and Cu+. The
chemical shifts observed for AgX and CuX zeolites are small (compared to the
Na+ form) and even negative (AgX) [61-64]. The adsorption ofXe in dehydrated
AgX is much greater than in NaX. Most remarkably, the shifts decrease with
increasing Ag+ content down to negative values, -40 ppm for [Xe]~O, (Figure
4b). These results have been attributed to specific interactions of xenon with Ag+
cations in the supercages, especially with Ag+ in SIll sites [65]. This location of
Ag+ allows close contact with Xe that favours electron donation involving the Ag+
4d IO and Xe 5do orbitals. This process is considered responsible for the observed
low-frequency signal. For CuX and CuY zeolites, during zeolite dehydration, the
Cu2+cations initially present in the zeolite are partially autoreduced to Cu+. These
Cu+ cations behave like Ag+ and participate in 3d 1o_ 5do donation [64].

l~.-------------~

llS ,---- -------- ------,-,...

_~ +-r-.~,_r_r__r_._;rr_r_,..,..,~r-r'
SE+28 lE+21 o 1 2 3
XeAtomsIg
XeoonlSupereage

Figure 4. Chemical shift variation versus Xe concentration for (0) HY, NaY with
different SilAI ratios: (L\) 1.35; (<» 2.42; (X) 54.2; (0) 1.28 and MgY zeolite for
various magnesium contents: (+) 47%; (A) 53%; (e) 62%; (_) 71% (A) (from
ref. 56, with permission); and AgX zeolite for different cation contents: 0 %; <-)
(e) 20 %; (A) 40 %; (+) 80 %; (0) 100 % (B) (from ref. 62, with permission)

3.4. EXTRA-FRAMEWORK SPECIES

Xe has been much used to detect the presence and in certain cases the
concentration or the location of extra-framework species such as AI, coke, water,
hydrocarbons, metal particles.
261
3.4.1 . Extra-framework Al and Coke
Zeolites are often dealuminated in order to increase the acid strength. Extra-
framework Al species may then be deposited in the pores. Their presence is easily
detected if they interact with Xe [66]. They also play a major role in coke
formation during hydrocarbon cracking reactions [67]. The coking of various
zeolites, HY, HZSM-5 and NaA, has been studied. The distribution of coke inside
the crystal porosity or on the external surface as well as the distribution of coke
over the crystallite bed has been followed as a function of the coke content or
coking conditions.

3.4.2. Supported Metal Particles


12'1ce NMR has been widely used to study supported metal particles, mostly on
zeolites but also on Ah03 or Si02. Such studies concern Pt metal and other
transition metals like Pd, Ir, Ru, Rh, Mo and bimetallic alloys. Metal particles are
strong adsorption sites for Xe, and the Xe shift may be very high at low Xe
loading (Figure 5). The ~SAS (=~ in this case) has been estimated, by
extrapolation to zero Xe concentration, to be 1000-1300 ppm for PtlNaY [68].
The chemical shift increases with the metal concentration and the spectrum
depends on the particle distribution in the sample. In particular, a second signal
appears with a chemical shift characteristic of pure NaY, if there are large zones
of the crystallites without metal particles [69]. The location of metal clusters and
metal precursors can be followed as a function of calcination conditions.

500
~ <> O%Pt
'. o O.8~. Pt
400 \ I>.
3%Pt
\.

."
c 14-1. Pt
'''-.
E 300 "-
=-=- ,,

00 l\ ,,
. '<l...._
200 ,
"- ,
.....

100 "'-
""-
"'-<>...-
---
0
0 100 200 300 400

PXe Torr

Figure 5. 12'1ce chemical shift of xenon adsorbed in PtlNaY zeolite (from ref. 68
with permission)
262

The chemisorption of hydrogen has been used to determine the mean size of
the metal particles. The adsorption of H2 on a part of the metal particles of the
zeolite crystallite changes the Xe-particle interaction, and a second line appears in
the spectrum, at lower chemical shift, its intensity increasing with the amount of
H2 adsorbed. Assuming a given chemisorption stoichiometry (number of H2
molecules per metal particle), the total number of metal particles can be
determined [68].

3.4.3. Co-adsorbed Molecules


Co-adsorbed molecules can also be detected. The location of water molecules in a
HY zeolite in function of their concentration has been studied in detail [70].
The distribution of other organic molecules, such as n-hexane, benzene, tri-
and hexamethylbenzene, in NaY zeolite has been investigated [71-73]. The
aromatic molecules do not diffuse rapidly in the crystallites, due to strong
interaction with the N a + cations, and the distribution remains heterogeneous over
long periods. It is necessary to heat the sample at high temperature (>500 K) to
obtain homogeneous distribution of the aromatics.

3.5. DIFFUSION OF COADSORBED MOLECULES

Since the NMR parameters (chemical shift, linewidth and intensity) depend on the
nature and on the concentration of a co-adsorbed molecule, Xe can be used to
probe the concentration variation of a co-adsorbed molecule during an adsorption
(or desorption) process. The evolution with time of spectra (Figure 6) makes it
possible to describe the diffusion process, to determine concentration profiles and
to estimate diffusion coefficients in systems out of equilibrium (mass transport
diffusion) [74].
This approach requires a model of diffusion usually involving macropores
(intercrystallite space) and micropores (intracrystallite pores), and the analytical
solution of the diffusion equations expressing the mass balance of the diffusant in
each type of pore [75]. In a model consisting of two types of diffusion,
intercrystallite diffusion and intracrystallite diffusion characterized by the
diffusion coefficients pinter and Dintra, respectively, the solutions of the two
corresponding diffusion equations give the concentration gradients of the diffusant
in the bed and in the crystallites in function of spatial coordinates, time, the
crystallite size and the diffusion coefficient Dintra (Figure 7). The NMR spectra
are simulated assuming Lorentzian lines whose parameters are previously
determined in function of the diffusant concentration under equilibrium
conditions.
The benzeneIHZSM-5 system has been studied by this method [76]. The
influence of compression of the sample, the length of the catalyst bed and of the
263

addition of a binder to the zeolite was been analysed. The diffusion coefficient
obtained from the NMR spectra simulation (Dintm ~7xl0-15 m2s-l) is in good
agreement with the literature. Its value is the same for compressed samples.
Results obtained with n-hexane (Dintm ~ 1x 10-13 m2s-l) and paraxylene in
HZSM-5 (Dintm ~3xI0-16 m2s-1) are also consistent with the literature.

1.0 , - - - - - - = - 1.0 r--;;~:::--.., 1.01======9


CUi

8.4

6.2
o
o 6.2 8.4 8.6 o.s to 6.2 8.4 8.6 o.s to
Z X
Figure 6. Benzene concentration profiles in the intercrystalline space, C(Z,t),
along the normalized bed length, Z = zit, and in the intracrystalline pores, Q(Z,
X, t), along the normalized crystallite radius, X = rlR, for crystallites located
either at the bottom of the bed (Z = 0) or at a height of Z = 0.8; bed lengths, t = 5
mm (from ref. 76, with permission)

2.5 2.3

2.1 2.0

1.4 to
0 0
i i i i
180 150 no 90
o(ppm)
Figure 7. Experimental and simulated l~e NMR spectra for benzene adsorption
in HZSM-5 (from ref. 76, with permission).
264
4. References
J. Ito, T. and Fraissard, J. (1980) in L.V.c. Rees (ed.), Proceedings of the Fifth International
Conference on Zeolites, Heyden, London, pp 510-515
2. Ripmeester, J. A. and Davidson, D. W. (1981) J Mol. Struct. 75(1), 67-72
3. Fraissard, J., Ito, T. and Springuel-Huet, M.-A. (1988) J Chim. Phys. 85, 747-757
4. Dybowski, C., Bansal, N. and Duncan, T. M. (1991) Ann. Rev. Phys. Chem. 42, 433-464
5. Barrie, P. J. and Klinowski, J. (1992) Progr. NMR Spectr. 24, 91-108
6. Raftery, D. and Chmelka, B. F. (1994) NMR Basic Principles and Progress 30, 111-158
7. Springuel-Huet, M.-A., Bonardet, J.-L. and Fraissard, J. (1995) Appl. Magn. Reson. 8, 427-
456
8. Bonardet, J-L., Fraissard, J., Gedeon, A. and Springuel-Huet, M.-A. (1999) Catal. Rev.-Sci.
Eng. 41, 115-225
9. Ratcliffe, C. I. (1998) Annu. Rep. NMR Spectrosc. 36, 123-221
10. Ripmeester, J. A. (1990) J Phys. Chem. 94, 7652-7656
I J. Cheung, T. T. P. (1995) J Phys. Chem. 99,7089-7095
12. Chen QJ. and Fraissard, J. (1992) J Phys. Chem. 96, 1815-1819
13. Ripmeester, J. A, Ratcliffe, C. I. (1993) Anal. Chim. Acta. 283, 1103-1112
14. Jameson, C. J., Jameson, A. K., Gerald, R. E. and Lim, H. M. (1997) J Phys. Chem. BIOI,
8418-8437
15. Samant, M. G., de Menorval, L. c., Dalla Betta, R. A. and Boudart, M. (1988) J Phys.
Chem. 92, 3937-3938
16. Jameson, C. J., Jameson. A. K. Gerald, R.E. and de Dios, A. C. (1992) J. Chem. Phys. 96,
1676-1689 and 1690-1697
17. Larsen, R. G., Shore, J., Schmidt-Rohr, K., Emsley, L., Long, H., Pines, A., Janicke, M. and
Chmelka, B. F. (1993) Chem. Phys. Lett. 214, 220-226
18. Jameson, C. J., Jameson, A. K. Lim, H. M. and Baello, B. I. (1994) J Chem. Phys. 100,
5965-5976
19. Cheung, T. T. P. (l993)J Phys. Chem. 97, 8993-9001
20. Nivarthi, S. S. and McCormick, A. V. (1994) Microporous Mater. 3, 47-53
2 J. Ripmeester, J. (1984) J. Magn. Reson. 56, 247-253
22. Ito, T., de Menorval, L.-c., Guerrier, E. and Fraissard, J. (1984) Chem. Phys. Lett. 3, 271-
274
23. Tsiao, c.-J., Kauffman, J. S., Corbin, D. R., Abrams, L., Carrol, E. E. and Dybowski, C.
(1991)J Phys. Chem. 95, 5586-5591
24. Fetter, G., Tichit, D., de Menorval, L-C. and Figueras, F. (1990) Appl. Catal. 65, Ll -L4
25. Barrie, PJ., McCann, G.F., Gameon, I., Rayment, T. and Klinowski, J. (1991) J Phys.
Chem. 95, 9416-9420
26. Tsiao, CJ., Carrado, K.A. and Botto, R.E. (1998) Microporous and Mesoporous Materials
21,45-51
27. Terskikh, V. V., Mudrakovskii, I.L. and Mastikhin, V.M. (1993) J Chem. Soc., Faraday
Trans. 89,4239-4243
28. Cros, F., Korb, J.P. and Malier, L. (2000) Langmuir 16, 10193-10197
29. Stengle, T.R. and Williamson, K.L. (1987) Macromolecules 20,1428-1431
30. Miller, J.B. (1993) Rubber Chemistry and Technology 66,455-461
3 J. Miller, J.B., Walton, I.H. and Roland, C.M. (1993) Macromolecules 26,5602-5610
32. Walton, 1.H. (1994) Polymers & Polymer Composites 2,35-41
265
33. Tomaselli, M., Meier, B. H., Robyr, P., Suter, U. W. and Ernst, R. R. (1993) Chem. Phys.
Lett. 205, 145-152
34. Mansfeld, M. and Veeman, W. S. (1993) Chem. Phys. Lett. 213, 153-157
35. Mirabella, F.M. and McFaddin, D.e. (1996) Polymer 37,931-938
36. Mansfeld, M., Flohr and Veeman, W.S. (1995) Appl. Magn. Reson. 8, 573-579
37. Yang, C.Y., Wen, W.Y., Jones, A.A. and Inglefield, P.T. (1998) Solid State Nucl. Magn.
Reson. 12, 153-164
38. Shibanuma, T, Asada, H., lshi, S.1. and Matsui, T. (1983) Japan. 1. App/. Phys. 22, 1656-
1658
39. Neue, G. (1987) Zeitschriftfiir Physikalische Chemie 152, 271-280
40. Suh, D.J., Park, T.J., lhm, S.K. and Ryoo, R (1991) 1. Phys. Chem. 95,3767-3771
41. Bansal, N., Foley, H.e., Lafyatis, D.S. and Dybowski, e. (1992) Cata/. Today 14, 305-316
42. McGrath, K.J. (1999) Carbon 37(9), 1443-1448
43. Ago, H., Tanaka, K., Yamabe, T., Miyoshi, T., Takegoshi, K., Terao, T., Yata, S. Hato, Y.
Nagura, S. and Ando, N. (1997) Carbon 35,1781-1787
44. Ito, T., Springuel-Huet, M.A. and Fraissard, 1. (1989) Zeolites 9, 68-73
45. Simpson, 1.H., Wen, W.Y., Jones, AA, Inglefield, P.T. and Bendler, 1.T. (1995) Appl.
Magn. Reson.8,349-356
46. Urban" e., McCord, E.F., Webster, O.W., Abrams, L., Long, H.W., Gaede, H., Tang, P. and
Pines, A (1995) Chem. Mater. 7, 1325-1332
47. Kentgens, AP.M., van Boxtel, H.A., Verweel, R.I. and Veeman, W.S. (1991)
Macromolecules 24,3712-3717
48. Simpson, J.H., Wen, W.Y., Jones, A.A. and Inglefield, P.T. (1996) Macromolecules 29,
2138-2142
49. Chen, Q.J., Springuel-Huet, M.A., Fraissard, 1., Smith, M.L., Corbin, D.R., and Dybowski,
C. (1992)1. Phys. Chem.96, 10914-10917
50. Ferrero, M.A., Webb, S.W., Bonardet, J.-L., Conner, W.e. and Fraissard, J. (1992) Langmuir
8, 2269-2274
51. Terskikh, V.V., Mastikhin, V.M., Timofeeva, M.N., Okkel', L.G., Fenelonov, V.B. (1996)
Catal. Lett. 42, 99-104
52. Fraissard, 1., Ito, T. (1988)Zeolites 8, 350-361
53. Cheung, T. T. P., Fu, e. M. and Wharry, S. (1988)1. Phys. Chem. 92, 5170-5180
54. Ito, T. and Fraissard, J. (1987) 1. Chem. Soc., Faraday Trans. 183,451-462
55. Gedeon, A and Fraissard, J. (1994) Chem. Phys. Lett. 219, 440-444
56. Chen, Q. 1., Ito, T. and Fraissard, J. (1991) Zeolites 11,239-243
57. Kim, J.-G., Kompany, T., Ryoo, R., Ito, T. and Fraissard, 1. (1994) Zeolites 14, 427-432
58. Chen, Q. 1. and Fraissard, 1. (1990) Chem. Phys. Lett. 169,595-598
59. Gedeon, A, Bonardet, J.-L. and Fraissard, J. (1989) 1. Phys. Chem. 93, 2563-2569
60. Bonardet, 1.-L., Gedeon, A and Fraissard, J. (1995) Stud. Surf Sci. Cata/. 94,139-146
61. Gedeon, A, Burmeister, R, Grosse, R., Boddenberg, B. and Fraissard, 1. (1991) Chem. Phys.
Lett. 79, 191-195
62. Grosse, R, Burmeister, R, Boddenberg, B., Gedeon, A and Fraissard, 1. (1991) J. Phys.
Chem. 95,2443-2447
63. Gedeon, A, Bonardet, J.-L. and Fraissard, J. (1993) 1. Phys. Chem. 97, 4254-4255
64. Gedeon, A, Bonardet, J.-L., Lepetit, e. and Fraissard, 1. (1995) Solid State NMR 5,201-212
65. Grosse, R., Gedeon, A, Watermann, J., Fraissard, 1. and Boddenberg, B. (1992) Zeolites 12,
909-915
66. Chen, Q., Guth, J.L., Seive, A, Caullet P. and Fraissard, J. (1991) Zeolites 11, 799-803
67. Barrage, M.-e., Bonardet, J.-L. and Fraissard, J. (1990) Cata/. Lett. 5, 143-154
68. de Menorva1, L.-c., Fraissard, 1. and Ito, T. (1982) 1. Chem. Soc., Faraday Trans. 78, 403-
406
266
69. Pandya, K., Heald, S., Hrildac, 1, Petrakis, L., and Fraissard, 1 (1996) J. Phys.Chem. 100,
5070-5075
70. Gedeon, A., Ito, T. and Fraissard, 1 (1988) Zeolites 8, 376-380
71. Chmelka, B., Ryoo, R., Liu, S.B., de Menorval, L.-c., Radke, C.J., Petersen, E.E. and Pines.
A. (1988) J. Am. Chem. Soc. itO, 4465-4467 .
72. de Menorval, L.c., Raftery, D., Liu, S.B., Tahegoshi, K., Ryoo, R. and Pines, A. (1990) J.
Phys. Chem. 94, 27-31
73. Chmelka, B.F., Pearson, J.G., Liu, S.B., Ryoo, R., de Menorval, L.-C. and Pines, A. (1991)
J. Phys. Chem. 95, 303-310
74. Karger, J., Pfeifer, H., Wutscherk, T., Ernst, S., Weitkamp, 1 and Fraissard, J. (1992) J.
Phys. Chem. 96, 5059-5063
75. N'Gokoli-Kekele, P., Springuel-Huet, M.-A. and Fraissard, J. Adsorption submitted
76. N'Gokoli-KekeIe, P., Springuel-Huet, M.-A., Bonardet, 1-L., Dereppe, J.-M. and Fraissard J.
(2001) Stud. Surf Sci. Cata/. 133,375-382
THE MECHANISM FOR IONIC AND WATER TRANSPORT IN NAFION
MEMBRANES FROM RESONANCE DATA

VITALY l.VOLKOV·, EVGENY V.VOLKOV··, SERGE F.TIMASHEV·


• Karpov Institute of Physical Chemistry) 0, Vorontsovo Pole, Moscow,
103064, Russia
"Physics Department, Moscow State University, Moscow 1l7234, Russia

The mechanism of ionic and molecular transport in nanosystems can be understood on


the basis of magnetic resonance studies of the transport interconnection channel structure. the
ionic and molecular state and the translational mobility in different spatial scales.
Ion-exchange polymer membranes were studied as model systems. The ion exchange
membranes bear acid or basic charged groups; these groups along with mobile ions and water
molecules constitute a transport channel network.
Nafion, a perfluorinated ion-exchange membrane (from DuPont) is a well known regular-
structured ion-exchange system. The membrane structure consists of perfluorinated polymer
chains and an ethereal side-chain with sulfonic acid groups capable of exchanging positive
counter ions (protons or alkali and other metal ions). Fig. 1 shows the structures of the polymer
chains and transport channels.

Figllre I . Schematic representation of the amorphous


fragment of a sulfonated cation exchange
membrane. I - main polymer chain; 2 -
hydrated counterions and ionogenic groups at
low water content; 3 - ion and water transfer
"channels" at high water content; LI=4 nm from
low-angle X-ray scattering data; L2=IO nm from
Mossbauer spectra; II and 12 from ENDOR and
NMR relaxation data; h rrom porosimetry and
ENDORdata.

267
1. Fraissard and O. Lapina (eds.i, Magnetic Resonance in Colloid and Interface Science, 267-275.
© 2002 Kluwer Academic PubLishers.
268

The structure of transport channels was studied using porosimetry, X-ray scattering,
Mossbauer spectroscopy, EPR, ENDOR, NMR-relaxation and PFG NMR techniques [I] . The
transport channels were found to occupy almost a quarter of the polymer volume and to be very
regularly structured. They are about 3-5 nm in width, sulfonic acid groups being ca. 7 nm from
one another and the charged groups being rather evenly distributed.
The diameter of ionogenic groups, including hydrated metal ions, is about 1.0 nm. These
structural characteristics depend strongly on the concentration of water, sulfur groups and on the
polymer pretreatment.
Magnetic resonance.is a powerful tool for studies in the field.

EPR [2-4]
If counter-ions are paramagnetic or are to be used as a paramagnetic probe, EPR is the
preferred technique. Analysis of EPR spectra provides information on the charged group
distributions, the probability of arranging these groups, coordination bonding between charged
groups and metal ions. Application of the END OR technique makes it possible to study not only
the nearest neighbours but also remote coordination spheres. Detailed information on the
neighbourhood of the paramagnetic centre (I to 1.5 nm) can be obtained by combining EPR and
ENDOR studies.

High resolution NMR. [5-7]


Subjects of NMR studies were internal water and Li', Na·, Cs+ cations sorbed by the
Nafion membrane. The aim was to understand the behaviour of water and the ions. by
simultaneous measurement of both the chemical shift and the line width.
We have investigated the dependence of chemical shifts and line widths on humidity. The
line width is reciprocally proportional to the spin-spin relaxation times, and thus it is
proportional to the residence times of ions on the sulfonic acid group.
High resolution W NMR studies allow the number of water molecules ho in the first
hydration shell of ions to be determined [5]. The value of ho is smaller than that for aqueous
solutions of acid and salts containing the same cations (Fig. 2, 3).
For protons, ho is always equal to 2. Hence, the proton in aqueous solution usually exists
in the form ofH 5 0 z·, not as H30+.
The ho value is maximal for Na +, as can be seen from the dependence of ho on the
crystallographic radii of cations (Fig. 3).
It is also possible to determine the fraction of broken H -bonds against that in pure water.
The probability of breaking the hydrogen bond decreases from Li· to Cs + for alkali metal ions
and from Mg2+ to Ba2+ for alkaline earth metal ions (Fig. 4) .
High resolution alkali metal NMR studies allow the mechanism of ionic interaction with
the ionogenic sulfo group to be established and, consequently, the nature of membrane
selectivity to be understood [6, 7].
The type of cation-sulfo group interaction depends on the water content. When · the
number of water molecules per charged group (n) is less than ho, the cation and the charged
group interact directly to form a contact ionic pair.
lfn > ho, there are isolated ionic pairs.
The number of contact ionic pairs and the diffusional ionic mobility, which is determined
by the residence time 'd, depend on n. The 23Na and I33Cs NMR data allow us to calculate the
fraction of contact ionic pairs and to estimate a value proportional to 'd [6].
269

9
3

1 1I(2n +5)
0~__4 -_ _~_ _~0,_12___0~,1_6___
0,6 0,8
1/ n
8
-0,5

6
-1,0

2
-1,5
1
Figure 2. Proton chemical shifts of water as a function of moisture content in perfluorinated sulfocationite
membranes. Points stand for experimental values, curves I through 8 for data calculated by equation (7);
curve 9 for data calculated by equation (5); I - Lt, 2 - Na+, 3 - K+, 4 - Rb+, 5 - Cs+, 6 - Bal +, 7-
Cal., 8 - Mg l ., and 9 - W (5].

ho __ Na+(~.8)
/ " ....
I
....
4
I
I K+Q.5)
I
I Rb+(3.3)
Li+ (3.3) Cs. +
. (3.0)
3
I ""
I "
H+(2.0)
2 ,
0

0.57 0.76 l.01 1.34 1.49 l.68 r,A

Figure 3. Hydration number ho for H+ and alkali metal ions for Nation·type membranes [5).
270

a
'Li+ ,Mg2+
0.6
JVa , C
'i;r +
, a,
2+

,,
...
0.4 +, 2+
K ,Ba Rb+

0.2 ... ...


o

1.0 1.5 2.0 r,A

Figure 4. Proportion of broken H-bonds, a, for alkaline earth metal ions in perfluorinated Nation type membranes[5]
It is seen from Fig. 5, that for CST the fraction of contact ionic pairs is greater than that for
Na', but the residence time of Cs +ions on the SO) - group is smaller. This is because of different
crystallographic radii, which make Na+ positively hydrated but Cs+ negatively hydrated . These
data explain why Cs + ions penetrate the membrane easier than Na" ions: the Cs + ion interacts
more strongly with SO) - groups than Na + and is more readily sorbed, but the residence time for
Cs+ is shorter and Cs+ moves faster than Na+.

T I, T 2 relaxation measurements. Self-diffusion coefficient measurements.

Spin-lattice and spin-spin relaxation processes were studied for protons of water
molecules and for 7Li of Lt ions in Lt ionic form of the Nation membranes to obtain the
correlation times for water and Lt mobilities.
The correlation times and the motion activation energies were calculated from the
temperature dependence of the relaxation rates r l, r2 using models based on the Bloembergen,
Purcel, Pound theory.
The method of calculation is given elsewhere [8]. The next step was to understand what
kind of motion is characterized by these correlation times.
From IH relaxation data [8], tH20 is the time of water molecule rotation by one radian,
this being about the time for the elementary jump of a water molecule to the distance
comparable to its size, 0.3 nm.
From the 7Li relaxation data [9], th is the time of water motion oscillation about LiT ions;
this is really the time for a hydrogen atom of a water molecule to move to the distance of the
hydrogen bond between water molecules. td is the residence time of Lt ion on the SO) - group.
These parameters characterize the elementary diffusion processes in the membrane completely.
It is possible to calculate the macroscopic self-diffusion coefficients. The simplest way is to use
the Einstein equation.
For water, the self-diffusion coefficient will be

DC
H,O
=~
6 (1)
- 'f Hp
271

Pc a) Ard b)
1,0

Cs+ --<>- Cs+


0,5 Na+ -ij- Na+

°o 1 2 3 4 n o
Figure 5. The fraction of contact ionic pairs, Pc' (a) and the residence times
(b). (Obtained from 2JNa and IllCS NMR data [6]).
1
td
2 3 4 n
ofNa+ or Cs· ions on the sulfo group

where d is the water molecule size equal to ca. OJ nm.


For Lt ions
2
DC =~ (2)
u+ 6-rd
where r is the distance between the sulfo groups.
The self-diffusion coefficients De H20, De Lt can also be measured directly using the PFG
NMR technique. The values of 'tH20, 'th, 'td, DC H2O, D\t, D\120, D\t are shown in Table 1 for
different n, where n is the number of water molecules per SO)- group.
TABLE 1. Values Of'tH20, 'tho 'to, DeH2o , D\t, DeH20 , D\t for Nafion membrane.

D CH20, D\j+, De H20, D\t,


n 'th, S 'td, S 'tH20,S
m 2 /s m 2/s m 2 /s m 2/s
4 8'10,\0 10,8 10'9 5·10,12 2·10,12 10,12 10,12
20 7'10,11 10,9 10,10 3·10,10 4·10'\0 8·10,11 10,11

As can be seen from Table 1, the self-diffusion coefficients calculated from the
microscopic data are in a good agreement with the experimental macroscopic self-diffusion
coefficients measured independently by PFG NMR. This is a very important result: ion and
molecule transfer in the Nafion membrane is controlled by the times of ions or molecules
jumps. These jump times are controlled by the interaction of ions or molecules with the charged
groups as well as by the membrane channel structure.
The contribution of the interaction to self-diffusion increases if the water content in the
membrane decreases, because cations of metals interact directly with the charged groups at low
n < ho to produce cation-anion contact pairs.
Thus, there is a relationship between the water proton chemical shift and water molecule
272

self-diffusion coefficients for the Nafion type membranes in alkali metal ionic species (Fig. 6)
[5].

(42+

11

0,1 0,5 1,0


Figure 6. Self-diffusion coefficient of hydration water VS. its proton chemical shift for Lt, K+, Rb+, Cs+, Sa 2" Ca 2+
and Mg2+ ions in perfluorinated sulfocationite membranes for n = 2.5 . For W species (Oc -tSmol = 4 ppm
and D = 10- 10 m2/s for n = 2.5.

10-10

10-11

0,5 1,0
mg-eq
Ion - exchange capacity - - -
ml
Figure 7. Dependence of water self-diffusion coefficient on the ion-exchange capacity for Nation-type ion-exchange
membrane.

The relationship between Ds and 0 was investigated at h=2.5 < ho; 0 characterizes the
bond of water molecules with the ionogenic groups including the (S03-Me+)(H20)n system. As
shown by high resolution NMR, an increase in the chemical shift corresponds to an increase in
the water molecule-charged group bonds. Apparently, the possibility of jumps of ions or water
molecules should depend on the distance between the charged groups and on the size of
273
hydrated charged species [1,2).
For example, the self-diffusion coefficient is strongly dependent on the ionogenic group
concentration (see Fig. 7). At high water content and high charged group concentration, the
hydrated shelJs of charged groups overlap and an infinite network of hydrogen bonds appears
(Fig. 8).

TT
Figure 8. Situation corresponding to high sulfo group concentration and high water content.

Under these conditions the influence of the interaction of water molecules and mobile
ions with sulfo groups not very specific and therefore, the selectivity of ion-exchange
membranes is not very high.
R - SO;-------. H 0

TT H20~OH2 OH 2

Figure 9 Model corresponding to low concentration of su lfo groups and low water content.

At low water content and low charged group concentration, the hydration shells of
charged groups do not overlap and the main reason for ion and molecule transfer is the entropy
factor. Under these conditions transfer becomes very sensitive to the nature of the ion and water
content n. This is the region of selective transfer where high membrane selectivity can be
observed (Fig. 9).
The self-diffusion coefficient can be obtained from percolation theory. For the entropy
contribution to the self-diffusion coefficient [I] (Fig. 10):

D = DO exp[ p(a - n1/ 3 )'1 ] (3)


274

where p and Il characterize the probability of molecules or ions migrating between neighbouring
functional groups, a is determined by the density of water packing within the counter-ion
hydration shells. These are exactly the parameters determining the membrane selectivity.
In(Dsxl013)
8

4 3 2 1
Figure 10. Dependence of water self-diffusion coefficients on moisture content in perfluorinated sulfocationite
membranes in the K+ form for p=-I.S9±O.OS and ,u=2.9±O.1 and in the Lt form for p=-I.5S±0.45 and
,u=2.3±O.S.

D, m2/s
••
10-10
0
0
10- 11

10-12

10 -13 '---_-'--_.........._-'-_--'-_---'
o
Figure II . Dependences of water and Lt self-diffusion coefficients on water content for Li- form of Nafion-type
membranes. Lines are for the percolation equation (3). Points are for the experiment.
The dependence of the self-diffusion coefficients on the water molecule content is
reproduced in Fig. II for water and Lt ions in Nafion membranes.
It is clear that the self-diffusion process is described by percolation theory at low water
content and that the parameters p, ).l and a are the same as those for water and Lt ions. This
275

means that at low water content the translational motions correlate for water molecules and for
Lt ions. (The lines are parallel, see Fig. 11.)
In conclusion, the elementary stages of water molecule and hydrated ion diffusion transfer
can be assumed to be as follows .
For example, T=300 K, n=20
I) "Oscillation" of H20 in the first hydrated Lt shell. 'th= 1O.los ; E = IS kJ/mol eLi spin-
lattice relaxation data [9]).
2) Jumps of water molecules at a distance equal to the water molecule diameter (OJ nm)
'tH20=0.9·10·los ; E=20.5 kJ/mol(IH relaxation data[8]).
3) Multistage diffusion of Lt ion from one sulfonic acid group to another
DLt =r(n)/6'td(n), 'td=6.6·10-9 s, E=40 kJ/mol, r is the distance between S03' 'groups eLi spin-
spin relaxation data [9]). For low water content the situation is very similar but stages 2) and 3)
are correlated. .

References
I. Timashev S.F.( 1991), Physical Chemistry 0/ Membrane Processes (phys. Chem. Series), Ellis Horwood, New
York,246.
2. Volkov V.1. and Timashev S.F.(1989) Magnetic resonance methods in the investigation of a perfluorinated ion
exchange membrane, RussJ Phys. Chem., 63, I08.
3. Volkov V.I., Gladkinh S.N ., Timashev S.F. etc.(1983) The investigation of pertluorinated sulfocation exchange
membrane structure by NMR relaxation and paramagnetic probe methods, RussJChem.Phys., 2, 49.
4. Volkov V.I., Muromtsev V.I., Pukhov K.K. etc. (1984) The investigation of pefluorinate sulfocation-exchange
membrane structure by matrix ENDOR technique, Doki. Akad. Nouk. S.S.S.R, 276, 395.
5. Volkov V.I., Sidorenkova E.A., Timashev S.F. etc. (1993) State and diffusion mobility of water in perfluorinated
sulfocationate membranes according to proton magnetic resonance data, Russ J.Phys. Chem., 67, 914.
6. Volkov V.I., Sidorenkova E.A., Korotchkova etc. (1994) The nature of the selectivity perfluorinated sulfocation
exchange membranes to the alkali metals on high resolution NMR data for 7 Li , Na+, Cs' data,
RussJPhys.Chem., 68, N2 2,275-281.
7. Volkov V.I., Sidorenkova E.A., Korotchkova S.A. etc. (1994) The influence of non exchange sorbed electrolyte
on the state and diffusion mobility of water and Na+, CST ions in perfluorinated sulfocation exchange membranes
on high NMR 'H, Na+, CST date, RussJPhys.Chem., 68, N23, 500-505.
8. Volkov V.I., Nesterov LA., Sundukov V.1. etc. (1989) The diffusion transfer of water in perfluorinated sulfocation
exchange membranes as studied by pulse NMR, Russ-J.Chem.Phys., 8, 209.
9. Nesterov LA., Volkov V.I., Pukhov K.K. etc. (1990) The magnetic relaxation of nuclei Li+ and litheum
counterions and water molecules motions in perfluorinated su Ifocation exchange membranes.
Russ.J.Chem,Phys., 9, 1155.
AN AGGREGATION NUMBER-BASED DEFINITION OF THE IONIZATION OF A
MICELLE

Demonstration with TRFQ, SANS, and EPR

BARNEY L. BALES
Department of Physics and Astronomy and the Center for Supramolecular
Studies, California State University at Northridge, CA 91330 USA

Abstract

The degree of ionization of a micelle, a, is defined by asserting that the aggregation


number, N, is dependent only on the concentration of counterions in the aqueous
pseudophase whether supplied by the surfactant or added salt. Employing different
combinations of surfactant and salt that give the same value of N leads to a straightforward
definition of a. The value of the aggregation number is not needed in the method; the
experiment need only assure that the value of N is the same for two combinations. The
definition is demonstrated with time-resolved fluorescence quenching, (TRFQ) small-
angle neutron scattering (SANS), and electron paramagnetic resonance (EPR).

1. Introduction

It has been well accepted for many years[l] that a fraction, a, of an ionic surfactant's
counterions are dissociated from the micelles leaving them charged[2, 3]. A quantitative
measure of the value of a is crucial in order to understand many aspects of the behavior of
micelles. Nevertheless, there is not a satisfactory consensus as to the value of a for any
surfactant, not even the often studied surfactant sodium dodecylsulfate (SDS). To
illustrate, Table I of Romsted's review[ 4] compiles 30 determinations of a for SDS near
the cmc. These values vary from a low of a = 0.14 to a high of a = 0.70. Romsted[4]
noted that the variability in the values of a using a particular technique was less than the
variability from technique to technique. That a particular technique would lead to
distinct values of a is not surprising in view of the fact that each technique is likely to
measure a different definition of what is meant by "dissociated" counterions. A decision
must be made about the location of the counterions--are they in the micelle or not?
Further, an assumption must be made concerning whether counterions take part in a
particular experiment as micelle-associated or not.
It would clearly be an important step forward to define a in a way that is
independent of any particular experimental method. In a recent paper; we proposed[5] such
a definition based on the value of N and demonstrated the method with EPR. The purpose
of this paper is extend the demonstration to TRFQ and SANS using data from the
277
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 277-284.
© 2002 Kluwer Academic Publishers.
278

literature. We also report another EPR study using a different spin probe in order to
establish that the method does not depend on any particular features of the spin probe.

2. Theory

The fundamental hypothesis is that N is uniquely given by the concentration of


counterions in the aqueous pseudophase, Caq , which, according to the conventional
pseudophase ion exchange mass balance relationship[2, 6] as modified by Soldi[7] is
given by

C aq = F(StJ{aSt + (J-a)Sf+ Cad), (I)

where St, Sf, and Cad are the molar concentrations of total surfactant, surfactant in
monomer form, and added common counterion in the form of salt respectively. Equation
1 expresses formally the concentration of counterions associated with the micelle in the
sense that governs the value of N. It makes no statement about their physical location.
The factor within the brackets would give the concentration of counterions in the aqueous
phase if that phase occupied the entire sample; however, at higher surfactant
concentrations the excluded volume effect becomes important[7]. Following Soldi et
al.[7], we correct for this excluded volume effect by including the factor F(St)

1
F(StJ = 1 - VSt ' (2)
where V is the molar volume of the anhydrous surfactant in moles per liter assuming that
the density of the surfactants is approximately 1.0 g/mL[7]. For the two surfactants
reported here, V = 0.272 and 0.288 liters/mole for lithium dodecylsulfate and sodium
dodecylsulfate, respectively. .
A value of a is measured by preparing two samples yielding the same value of the
aggregation number, but with different combinations of St and Cad. For these two
combinations the hypothesis asserts that the value of Caq is the same as follows:

F(StJ{aSt + {l - a}Sf+ Cad} = F(St'} {aSt' + {l - a}SI + C'ad}· (3)

As previously discussed [5], the terms involving Sf and Sf' cancel in eq 3, thus for an
equivalent value of N, we have

F(StJ{aSt + Cad} = F(St'}{aSt' + C'ad}· (4)

A rearrangement of eq 4 yields the value of a.

F(St)Cad - F(St')C'ad
(5)
a = F(St')St' - F(St)St .
279

fl f T rTT
ff
ff
T
:<;
if
.5 f f
t.)
·2 yf ~ T TT !!
£
§ ff~ ~
~ ~ +! !
J t
tp a

t
p'
c

F{O.16S+C
I ad
I

Figure 1. Procedure to determine the value of a from an experimental technique that


detects a quantity monotonic in the value of the aggregation number, the ordinate in each
case. The abscissa is the quantity F(St) faSt + Cad} where a is adjusted in search of a
common curve. Samples are prepared so that the surfactant concentration St is varied in
the absence of salt 0 or the salt concentration Cad is varied •. A coincidence of the
data shows that a = 0.33 is the correct value.

Any property that varies monotonically with N could, in principle, be used. In


principle, a could vary with Caq ; however, we previously showed that a is constant for
SDS up to [SDS] = 600 mM. This is accord with theory[3] and is certainly a reasonable
assumption for the typically encountered surfactant concentrations up to 0.2 M or so.
Thus, rather than computing a for each pair of samples, we illustrate a simpler approach
of requiring that all values of the measured property fall on a common curve when plotted
versus the variable F(S,} faSt + Cad} where a = constant.
One method is illustrated in Figure I. One series of samples is prepared in the
absence of salt, by varying the surfactant concentration and another by holding the
surfactant concentration constant and varying the salt concentration. The ordinate in each
case in Figure 1 may be any experimental quantity that varies monotonically with the
aggregation number. The abscissa is the quantity F(S,}{aSt + Cad}. The value of a is
adjusted in the Figure la - c in search of a common curve. Figure 1 was prepared with a
hypothetical value of a = 0.33, thus, at this value, a common curve is obtained, Figure
1b. Other combinations of surfactant and salt concentrations may be used as well.
Both TRFQ and SANS measure N directly, so this is ordinate for these two
experimental methods. The EPR method rests upon the fact that the hydration of the
polar shell of SDS micelles varies monotonically with N for aggregation numbers smaller
than about N = 130 above which the sphere-to-rod transition occurs[8]. The surface
hydration is detected by a spin probe residing in the Stem layer using an electron
paramagnetic resonance (EPR) technique detailed in recent papers[8-1 0].
280

110 6.0
2
N X
100

,
5.0

90.0

,
4.0
&0.0

70.0
~ 3.0

60.0 !~ a b
~ 2.0

50.0 ~

40 .0 11w..-'-'-'-u'"'"'-'-'.......L..........................~L.......'-'-'-uu.l 1.0 w......-'--'-'~.........L~-'-'-~...........~'-'
o 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.15 0.2 0.25 0.3 0.3 5 0.4

F{ [LiDS]+[LiOI]}, M F{O.27[LiDS]+[LiCI]} , M a

Figure 2. Aggregation numbers for LiDS derived from TRFQ measurements. The
symbols have the same meaning as in Figure I. (a) Data plotted assuming a = I and (b)
the best fit value a = 0.27. (c) variation of the mean square deviation from the power law
trial function resulting in a best-fit value of a = 0.27.

The method utilizes the difference in resonance fields of the low- and central-field
line of a nitroxide spin probe, denoted by A+ . Thus, the hyper fine spacing A+ is a
monotonic function of N for N < 130. The experimental details are identical to those
recently published[S] except that 16-doxylstearic acid methyl ester (16DSE) was used
instead of(SDSE).

Results and Discussion

Figures 2 - 4 are organized as suggested by Figure I with a =1 being presented in part a


and the best fit value of a used in part b . Figure 2 shows aggregation numbers for
lithium dodecylsulfate (LiDS) taken from the literature[lO]. The best value of a was
determined by calculating the mean square deviations of the data points from a trial
function, in this case a power law[ 10] A plot of the mean square deviations is plotted
versus a in Figure 2c showing a minimum at a = 0.27.
Thus, the general scheme of Figure 1 yields a = 0.27 from TRFQ results.
Obviously, the statistics are not sufficient to critically analyze this result; nevertheless, it
is in line with values found in the literature[ 10 - 12]. From variations of the critical
micelle concentration (cmc) with salt concentration, a value of a = 0.32 has been
derived[lO], showing that eq 4 leads to reasonable results when implemented with TRFQ.
It seems worthwhile to pursue the method using TRFQ, improving the statistics. TRFQ
might not be the method of choice to measure a, but it is important to establish that the
definition eq 4 is independent of the experimental method.
281

95 1 1 1
0 t .1 0

•• • •
N 0 0

90 C-

••
O 0
85
• • C-

• •
80
0 ••• 0 •• •
75
b
• a

• •
0 0
70

1 I
65
0 0.1 0.2 0.3 0.4 0.5 0.1 0.2 0.3 0.4 0.5 0.6 0.7
F {[LiDS]+[LiCI)}. M F( .65(LiDS}+[LiCI]}. M

Figure 3. Aggregation numbers for LiDS derived from SANS measurements. The
symbols have the same meaning as in Figure 1. (a) Data plotted assuming a = I; (b)
assuming the best fit value a = 0.65.

We note that although there are a number of uncertainties in determining the


value of N using TRFQ, eqs 4 and 5 have the tremendous advantage of being rather
insensitive to systematic errors as long as these errors do not correlate with St or Cad..
Figure 3 shows values of N taken from the literature for LiDS derived from
SANS measurements[13]. A plot of the type shown in Figure 2c (not shown) yields a
minimum at a =0.65, however this value is clearly out of line with literature values[lO]
and the TRFQ technique.
Thus, the application of the technique does not seem to work as well for SANS
as it does for TRFQ and EPR. One reason for this could be that micelle interactions enter
into the calculations for SANS whereas these are to a very high degree of approximation
absent in TRFQ and EPR.
Turning to the EPR results, three-line narrow EPR spectra of 16DSE in SDS
micelles typical of nitroxide free radicals undergoing approximately isotropic motion in
the motional narrowing region were observed for all samples. See, for example, Figure
I(a) ofref[8].
Figure 4 shows values of A+. In addition to the types of samples suggested by Figure I,
one other series was prepared such that the value of A+ was approximately constant.
These latter samples were prepared by mixing two samples in various proportions: one
with [SDS] = 0.500 M and [NaCl] == 0 and the other with [SDS] = 0.025 M and [NaCl]
= 0.136 M. The standard deviations in 3 - 5 runs using the same sample are smaller than
the symbols. A value of a = 0.29 ± 0.02 was found by minimizing the mean square
difference between the experimental points and a cubic trial function similar to Figure 2c
(not shown). Previously[5], we found a = 0.272 ± 0.017 using a different spin probe,
5DSE, thus, the value of a determined by EPR is not probe dependent.
282

A+, G

15.35

~ -
15.30
4IQ a b
0

• CO 0
15.25

o
15.20
• 0 ~
15. 15 Q] Q]
• •
15.10

• •
15.05
..
.
<> <> <> <>
15.00 <>

14.95 ••
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.05 0.1 0.15 0.2 0.25

F([SDSj+[NaCljl. M F{0.29[SDSj+[NaCljl. M

Figure 4. Hyperfine spacing A+ measurements derived from EPR. The symbols have
the same meaning as in Figure 1 with an addition series denoted by the crosses which is
due to various mixtures of two samples: one with concentration [SDS] = 0.500 M and
[NaCl] = 0 and the other with [SDS] = 0.025 M and [NaCl] = 0.136 M. (a) Data plotted
assuming a = 1; (b) assuming the best fit value a = 0.29.

The values derived from EPR compare well with the value a = 0.27 obtain by
Sasaki and coworkers[ll] from EMF measurements in a concentration cell which is likely
to be reliable because the measurements are almost free of assumptions and
approximations[ 11]. Values of a obtained from measurements of the dependence of the
cmc on the concentration of added common counterion salt, a= 0.322[12] are similar to
the values for LiDS. Light scattering techniques lead to values of a that are consistent
with one another, but lower than those derived from other measurements. For example,
from Table 1 of Romsted[4] five determinations of a from light scattering yielded a =
0.17 ± 0.02 where the uncertainty is the standard deviation in the five measurements. For
the other 25 determinations in the same table, using various techniques, a = 0.23 ±
0.05[4] which is within experimental error of the two determinations found using spin
probes.
Due to the simplicity of the method, it becomes feasible to investigate large
numbers of systems. In addition to studying other surfactants, investigation of the
variation of ex with a number of experimental parameters such as temperature, chain
length, counterion, and added non electrolytes becomes attractive.
The EPR method ought to be applicable to any micelle that fulfills the general
criterion of growing with increasing Caq and that possesses water in its polar shell that is
expelled as the micelle grows. We believe this to include most micelles in the slow
growth region where they are approximately spherical; i. e., below the sphere-rod
transition.
There are likely to be other experimental techniques that are capable of
implementing the defmition in eq 4. It is reasonable to assume that the size dispersion of
283

the micelles is dependent only on N. Thus, when matching values of N in order to apply
eq 4, the dispersions will also be matched which allows the method to be applied to
either a number- or weight-averaged aggregation number. Therefore, any technique that
consistently detects the same statistically weighted average aggregation number could be
utilized. Any method that must make significant corrections due to the concentration of
micelles will be severely handicapped, because to get meaningful measurements, one
must extend them over a rather large range of surfactant concentrations.

Conclusions:

An unambiguous definition of a results from the hypotheses that the aggregation number
and a depend only on Caq . In effect, the aggregation number becomes an internal
measure of the value of Caq . TRFQ and SANS results have been applied to the
aggregation number based definition of a, eq 4 for LiDS. The TRFQ results are
consistent with other techniques while the SANS results are not. Results for SDS
consistent with those recently published[5] and in agreement with literature values were
found by EPR using a different spin probe.
284

Acknowledgments. This work was supported by the NIH S06 GM48680-08, the CSUN
Research and Grants Committee, the College of Science and Mathematics and the CNRS.

References

1. Hartley, G. S. (1936) Aqueous Solutions of Paraffin-Chain Salts. Hermann and Cie,


Paris.
2. Bunton, C. A., F. Nome, F. H. Quina, and L. S. Romsted. (1991) Ion Binding and
Reactivity at Charged Aqueous Interfaces. Acct. Chem. Res. 24,357.
3. Jonsson, B., B. Lindman, K. Holmberg, and B. Kronberg. (1998) Surfactants and
Polymers in Aqueous Solution. John Wiley, Chichester.
4. Romsted, L. S. 1975. PhD Thesis. Indiana University.
5. Bales, B. L. (2001) A Definition of the Degree oflonization of a Micelle Based on its
Aggregation Number. J. Phys. Chem. In Press, summer 2001.
6. Chaimovich, H. , R. M. V. Aleixo, I. M. Cuccovia, D. Zanette, and F. H. Quina.
1982. , p. 949. In L. B. M. a. E. J. Fendler (ed.), Solution Behavior of
Surfactants, vol. 2. Plenum Press, New York.
7. Soldi, V., J. Keiper, L. S. Romsted, I. M. Cuccovia, and H. Chaimovich. (2000)
Arenediazonium Salts: New Probes of the Interfacial Compositions of Association
Colloids, 6 . Relationships between Interfacial Counter ion and Water
Concentrations and Surfactant Headgroup Size, Sphere-to-Rod Transitions, and
Chemical Reactivity in Cationic Micelles. Langmuir 16, 59-71.
8. Bales, B. L., L. Messina, A. Vidal, M. Peric, and O. R. Nascimento. (1998)
Precision Relative Aggregation Number Determinations of SDS Micelles Using a
Spin Probe. A Model of Micelle Surface Hydration. J. Phys. Chem. 102, 10347-
10 358.
9. Bales, B. L., A. M. Howe, A. R. Pitt, J. A. Roe, and P. C. Griffiths. (2000) A Spin-
Probe Study of the Modification of the Hydration of SDS Micelles by Insertion of
Sugar-Based Nonionic Surfactant Molecules. J. Phys. Chem. 104,264-270.
10. Bales, B. L., A. Shahin, C. Lindblad, and M. Almgren. (2000) Time-resolved
Fluorescence Quenching and Electron Paramagnetic Resonance Studies of the
Hydration of Lithium Dodecyl Sulfate Micelles. J. Phys. Chem. 104,256-263 .
11. Sasaki, T., M. Hattori, J. Sasaki, and K. Nukina. (1975) Studies of Aqueous
Sodium Dodecyl Sulfate Solutions by Activity Measurements. Bull. Chem. Soc.
Japan 48, 1397-1403.
12. Williams, R. J., J. N. Phillips, and K. J. Mysels. (1955) Trans. Faraday Soc. 51,
728.
13. Chen, S. H. (1983) Determination of Intra- and Inter- Particle Structures of Ionic
Micelles by Small-Angle Neutron Scattering. International School of Physics
Enrico Fermi, 281 - 301.
USE OF IH NMR IMAGING TO STUDY COMPETITIVE ADSORPTION OF
HYDROCARBONS IN ZEOLITES

J.-L. BONARDET, P. N'GOKOLI-KEKELE, M.-A. SPRINGUEL-HUET,


J. FRAISSARD

Laboratoire S.I.E.N - Chimie des Surfaces, CNRS FRE 2312, Universite P.


et M. Curie, 4 place Jussieu, 75252 PARIS Cedex 05, France

IH NMR imaging has been used to study the adsorption or desorption of pure or
mixed hydrocarbons (benzene, n-hexane,) on or from a fixed bed of HZSM-5 zeolite
crystallites. This technique makes it possible to visualise the progression of the
diffusing molecules in the bed of the catalyst and to determine their intracrystallite
diffusion coefficient. In the case of competitive adsorption, it gives the instantaneous
distribution of the two components before thermodynamic equilibrium is reached.

1. Introduction

Magnetic Resonance Imaging (MR!) is a technique developed in the 70s mainly in the
medical field. The development of more and more powerful equipment and increasingly
sophisticated numerical treatment of images and the fact that it is a non-invasive method
have made this technique a routine tool for human body investigation.
However, medicine is not the only field of application of MR!, and the 90s saw
considerable progress in this technique for the study of porous media. For example,
MRI has been successfully used to study solvent penetration and the dynamics of water
in polymers [1,2], the structure and dynamics of polymer gels [3], the segregation of
grains in blends [4], the permeation resistance of cement [5], pore size distribution
mapping [6], the drying kinetics of gels [7], the diffusion and molecular mobility of
water in 4A zeolites [8] and hydrocarbon diffusion in deactivated Y zeolites [9].
We present here a new application of this technique to the study of the diffusion at
room temperature of pure or mixed hydrocarbons (benzene, n-hexane) in HZSM-S
zeolite In particular, we show for the first time that it is possible to visualise the
distribution of several gases adsorbed competitively.
285
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 285-295.
© 2002 Kluwer Academic Publishers.
286

2. Experimental section.

The application of pulsed field gradients during the NMR pulse sequence used for the
detection of the resonance signal leads to concentration profiles of resonating spins in
the direction(s) of the gradient(s). It is therefore possible to obtain images in one, two or
three dimensions. In our studies we used two sequences: (i) The field gradient is applied
during the 1[/2 pulse and throughout the signal acquisition time. In this case the signal
only depends on the density of the spins along the direction of the gradient (Figure I-a).
(ii) Spin echo detection (Figure I-b) in which the profile obtained also depends on the
relaxation time of the spins. This sequence can decrease the signal intensity but gives
information about the distribution of the spin relaxation times T2 in the sample, as has
been shown in a study of Y zeolite deactivation during catalytic cracking [9]. The
spectra were recorded on a Bruker MSL 300 apparatus using a gradient of 4T m· l . Each
scan lasts 250 ms and each profile requires 100 scans.

r
1[ 1[
Spin echo

RFJi m 1!
a) 1[/2 b)

RF
I I
Gz ~ L I I
AQ IL Gz
I f

I4
300 ms
J
Figure. 1. Schematic representation of the 1D Imaging MR sequence: a) Direct
detection; b) Spin echo detection.

A HZSM-5 zeolite sample consists of spherical crystallites with average diameter


D = 40 !-lm, used in the form of a powder or compressed (at 103 bar) into cylindrical
pellets 7 mm in diameter and between 5 and 15 mm long, either pure or with a binder
(amorphous silica-alumina,).
287

The experimental cell (Figure 2) is


t
Va c uum
divided in two parts: the lower part
contains the sample and allows its pre-
treatment under vacuum at 673 K
overnight; the upper part is a reservoir
containing liquid hydrocarbon(s) in

vapour pressure. At time t = °


equilibrium with its (their) saturation

sample is put into contact with the


the
Tank
supply of liquid hydrocarbon (n-
Hydrocarbon hexane or (and) benzene) in
gas --~ equilibrium with the gas phase by
opening the stopcock S2 and is then
placed quickly in the magnet.
To distinguish the hydrocarbons when
they are mixed, one is perhydro and
the other perdeutero.
Sam pie

Figure 2 Schematic representation of the cell


containing the sample.

3. Results and Discussion

3.1 DIFFUSION OF PURE GASES

The hydrocarbon mass balances are given by the following equations [10, 11]:
-In macropores:

CinlerTt
a2 c -
ac = D interC inter -;--T 3(1 - cinler
)Dint ra
R (aarq ) (1)
uZ r =R
-In micropores:

(2)
288

c and q are the hydrocarbon concentrations in the macropores and micropores,


respectively; z and r the linear and radial coordinates in the bed (from the bottom) and
in the crystallite (from the centre); R the radius of the crystallites, Dinler and Dintra the
inter- and intracrystallite diffusion coefficients and Einter the macroporosity of the
sample.
These experiments were conducted at pressures where intercrystallite diffusion is
much faster than intracrystallite diffusion. This can be verified from the rectangular
concentration profile of the images. In this case only equation 2 is used, with the
following boundary conditions:

q(r, t=O) = 0; q(r=R, t) = qoo [1 - exp( -kt)]; oq(r = O,t) = 0


or
where k is the transfer constant of molecules to the external surface of the crystallites,
and qoo the hydrocarbon concentration in each crystallite at adsorption equilibrium.
Under these conditions, the solution of equation 2 is given by the Cranck equation (eq
3)[12].

M, is the total amount of adsorbed gas at time t (given by the integral of the NMR image
signal) and Moo the corresponding value at equilibrium.
Due to the small amount of hydrocarbon in the gas phase compared to that in the
adsorbed state, the hydrocarbon concentration in the macropores is always negligible
and the signal observed is mainly due to adsorption in the micropores.

3. 1. 1.Diffusion of Pure Benzene


At the beginning of adsorption, (0 < t < 0.1 h) the 1D images of benzene adsorbing in
the powder HZSM-5 bed present an adsorption front with a concentration gradient from
the top towards the bottom of the sample (Figure 3a). When t becomes higher than 0.2 h
we observe rectangular profiles for the loose as well as for the compressed powder
(Figure 3b) proving that after this time, under these experimental conditions (high
pressure of the gas phase) the benzene concentration is the same in all parts of the bed
289

and that the diffusion of hydrocarbon is then controlled by the micropores, as discussed
by Heink et al. [l3]. In the case of compressed pellets, the density of the sample
increases, increasing the SignallNoise ratio, (better NMR signal) and the profiles are
rectangular from the very beginning of adsorption. For a loose or compressed sample,
the signal intensity increases with the adsorption time. The experimental kinetic curves
M(t) = f(t) are simulated using e~uation 3, by fitting Diotra and k; for a powder sample,
this leads to values of about 10- 1 m2 S-I and 1.2 x lO-3 S-I, respectively. In the case ofa
pellet, the intracrystalline diffusion coefficient (necessarily!), but the k value (0.8 x 10-3
S- I) decreases slightly. The value of Diotra agrees well with those reported [14] in
literature or obtained by 129Xe NMR [15].

a) b)
Time (b)

Time (b)

z=o 7.=( z=o


gas flow gas flow

Figure. 3: 1D-NMR profiles of adsorbed benzene: a) on powder; b) on pellet

When the zeolite in powder form is mixed with silica-alumina (60% weight of
zeolite, length 25 mm), the benzene concentration profiles (Figure 4-a) present an
adsorption front which lasts more than 30 min. The intercrystallite diffusion rate is
decreased by the presence of the meso porous silica-alumina, but benzene also adsorbs
on the binder. In the case of a compressed mixed sample (length 12 mm) the profiles are
rectangular as of the first spectrum (Figure 4-b). The difference between the two cases
is mainly due to the decrease in the macroporosity and the length of the sample. These
290

results show that mass transport in industrial catalysts depends greatly on the binder, on
the dimensions and on the compression of the sample.

a)Pow~ Time (h) b) pellet Time (b)

~
12.00
8.00
5.40 11.00
4.00 8.00

~
0.60 5.33
4.83

~
0.50 2.50
0.30 0.42
0.25 0.33

~
0.20 0.25
0.15 0.17
0.08 0.08
,
j
z~f z;"O z::'f z=O
~
~

gas flow gas flow

Figure. 4: lD-NMR profiles in a mixture (zeolite/silica-alumina) during benzene


adsorption.

3.1.2. n-Hexane
For n-hexane adsorption on a HZSM-5 powder sample, the concentration profiles
(Figure 5) become perfectly rectangular after only t = 0.03 h. The value obtained for
Dintra ('''''10 013 m2 sol) is of the same order of magnitude as that reported in the literature
using other techniques (zero length column, frequency response, etc.) [13,14].

3.2. COMPETING DIFFUBION OF GAS MIXTURES (BENZENE + n-HEXANE)

We report the results concerning: i) their competitive adsorption, ii) the variation of the
distribution of a pre-adsorbed hydrocarbon during the adsorption of a second one.

3. 2.1. Competitive adsorption.


When n-hexane is adsorbed from a gas mixture of C6Hl4 + C6D6 both at the same partial
pressure ( 60 mbar ) the intensity of the profiles due to the IH nuclei of hexane first
291

decreases from the top to the bottom of the bed (Figure 6-a). Such a profile indicates a
strong hexane concentration gradient in the z direction of the sample, which persists up
to a time of 0.13 h. Several reasons can be proposed to explain this evolution:
intercrystallite diffusion slowed, surface barrier lower for n-hexane than for benzene,
etc. At intermediate time (0.22< t< 1.02 h) there is a slight excess of n-hexane in the
bottom of the bed before reaching equilibrium which is characterized by rectangular
profiles (t ~ 5 h).

Time (h)

17.80

~~~~3.87 1.01
0.11
--,-__ _._ , 0.04

7.=0
~

gas flow
Figure.S. I D-NMR profiles of n-hexane
adsorbed in HZSMS zeolite in powder
form.

The following explanation can be advanced: n-hexane, having the higher


intracrystalline diffusion coefficient, preferentially adsorbs in the first layers before its
pressure becomes equal throughout the bed. Benzene, whose intracrystallite diffusion is
slower than that of n-hexane, progresses along the bed and adsorb then on the
subsequent layers whose crystallites are free of any adsorbate. As time increases,
benzene also adsorbs on the first layers, displacing n-hexane towards the bottom where
n-hexane displaces in turn the benzene molecules to finally obtain a distribution
governed by thermodynamics. This scenario is confirmed by the evolution of the C6H6
concentration profiles recorded during an identical experiment in which benzene and n-
perdeuterohexane were used (Figure 6-b). These profiles clearly show that benzene first
292

adsorbs preferentially in the bottom layers (signal intensity higher at the bottom of the
bed) before displacing the n-hexane adsorbed in the top layers (increase of the signal at
the top). The intermediate states of the systems (0.25 < t < lh) are the result of
competing kinetic and thermodynamic effects, the diffusion of n-hexane being faster
than that of benzene while the latter is more strongly adsorbed inside the crystallite.

a) b)
Time (b) Time (b)
5.88
8.70 4.58
-..........~_
5.08 2.53
4.33 0.83
1.02 _ _ _- 0.57
0.47 0.32
0.22 0.25
0.13 0.08
0.12 0.02
0.07

Figure. 6. lD-NMR profiles of n-hexane a) and benzene b) during their competitive


adsorption in HZSM-5 zeolite.

3.2.2. Distribution ofpre-adsorbate during the adsorption of another gas.


Figure 7-a shows the evolution of the distribution of n-hexane C6HJ4 pre-adsorbed at 2
mbar during the adsorption of C6D6 at its saturation vapour pressure. At time t = 0 the
concentration profile of pre-adsorbed n-hexane is rectangular, proving a homogeneous
distribution of the adsorbate all over the catalyst bed. Immediately upon contact with
benzene, profiles become asymmetric, signal intensity increases from the top to the
bottom, showing that hexane desorbs preferentially from the upper layers, as is shown
in Figure 7-b. However, a practically rectangular concentration profile is obtained
rapidly, in about 0.2 h. It should be noted that the amount of hexane desorbed between
the beginning of the experiment and the attainment of a new equilibrium is about 40%.
293

140 ,
120 '
l00 ~
80 ,, $
~
6O~i==
40 +
Time (h) 20
o o 't-----....------
0.06
0.30 0.17 0.3 0.47 0.72
3.90
9.00

zkf z:bO
gas flo;
Figure. 7. a) lD-NMR profiles of pre-adsorbed C6HI4 during C6D6 adsorption. b)
Signal intensity versus time for different values ofzJi in the bed of the catalyst:. (t = 0
*
h), D(t = 0.02 h), 6 (t = 0.30 h), 0 (t = 1.16 h), • (t = 3.90 h), (t = 9.00 h)

Figure 8-a corresponds to the case of benzene C6H6 being pre-adsorbed (90%
saturation) and n-hexane C6DI4 adsorbing at its saturation pressure. As for pre-adsorbed
n-hexane, we initially observe a quasi-rectangular profile con-esponding to a uniform
distribution of benzene in the sample. After contact with the reservoir of perde utero n-
hexane, the total signal area decreases slowly with time. The decrease in the benzene
concentration between the beginning and the end of tae experiment is about 20%.
However its distribution in the sample is particularly inhomogeneous, as is shown in
Figure 8-b. The intensity decreases at the top and increases markedly at the bottom of
the bed, showing that perdeutero n-hexane adsorbs first in the upper layers, "pushing"
the benzene towards the bottom of the tube. The local partial pressure of benzene and,
in parallel, its concentration in the bottom of the bed increase. As the "wave" of C6DI4
reaches the bottom, this latter gas adsorbs on the lower layers; the two partial pressures
become uniform along the sample, and the benzene molecules can adsorb again in the
upper layers until the thermodynamic equilibrium is obtained. The distribution of
adsorbed gases is first determined by kinetics and then the system is governed by
thermodynamics.
294

The difference in the extent of displacement of one gas by another in the two
experiments confirms, if this were necessary, the greater affinity of HZSM-5 zeolite for
benzene.

Time (b) b)
o 60 ·

0.02
0.09
~~-- 0.17
0.31
1.03
_ _ _ 1.51
0.16 0.33 0.5 0.8 0.9
----_ 4.77
i - - - 1 6.06
i
z";'O
gas now

Figure. 8. a) ID-NMR profiles of pre-adsorbed C6H6 during C6DI4 adsorption. b)
Signal intensity versus zjl for different times in the bed of the catalyst. 1'1 (t = 0 h), 0 (t

*
= 0.02 h), • (t = 0.09 h), 0 (t = 0.17 h), • (t = 0.31 h), X (t = 1.03 h), ... (t = 1.51 h), •
(t = 4.77 h), (t = 16.06 h)

4. Conclusion

IH ID-Magnetic Resonance Imaging is a powerful tool for investigating hydrocarbon


diffusion in microporous structures. We can obtain two types of information: The first
pertains transport coefficients determined by simulation of the kinetic curves obtained
from the variation of the surface area of the signal with time during adsorption. The
intracrystallite diffusion coefficients of hexane and benzene in HZSM-5 determined by
this technique are 10,13 and 10,14 m2 S'I, respectively, in good agreement with data in the
literature. The second, which is very important in catalysis studies, comes from the
295

shape of the instantaneous concentration profiles reflecting the variation of the local
adsorbate concentration, and reveals competition between kinetics and thermodynamics.
In the case of competitive adsorption of several gases, this technique appears to be
the only one able to show the relative distribution of each of the gases in the adsorbent
and the variation of this distribution with time.

5. References

1. Blackband S. and Mansfield P. J., Phys. C: Solid State Phys., 19 (1986) L49.
2. Tanaka N., Matsukawa S., Kurosu H. and Ando I., Polymer, 39 (20) (1998) 4703.
3. Ando I., Kurosu H., Matsukawa S., Yamasaki A., Hotta A. and Tanaka N., Wileys
Polym. Networks Group Rev., Ser. I (1998) 331, John Wiley & Sons Publ Inc.,
New York.
4. Porion P., Sommier N. and Evesque P., Europhysics-Letters, 50 (3) (2000) 319.
5. Papavassiliou G., Milia M., Fardis M., Rumme R., Laganas E., Sepe A., Blinc R.
and PintarM.M., J. Am. Ceram. Soc., 76 (1993) 2109.
6. Strange J.H., Webber J.B .W. and Schmidt S.D., Magn. Res. Imaging,7-8 (1996) 803.
7. Koptyug 1., Fenelonov V.B., Khitrina M.Y., Sagdeev R.Z. and Parmon V.N., J.
Phys. Chern. B, 102 (1998) 3090.
8. Halse M.R., Magnetic Res. Imaging, 7-8 (1996) 745.
9. Bonardet J.-L., Domeniconi T., N'Gokoli-Kekele P., Springuel-Huet M.-A. and
Fraissard J., Langmuir, 15 (1999) 5836.
10. Ruckenstein E., Vaidyanathan A.S. and Youngquist G.R., Chern. Eng. Sci., 26
(1971) 147.
11 Lee, L.K., AIChE. J. 24 (1978) 531.
12. Cranck, J., in "The Mathematics of Diffusion" Clarendon Press, Oxford, 1956.
13. Heink W., Karger J. and Pfeifer H., Chern. Eng. Sci., 33 (1978) 1019.
14. Karger J. and Ruthven D.M. (1992) Diffusion in Zeolites, John Wiley & Sons Pub!
Inc, New York.
15. N'Gokoli-Kekele P., Springuel-Huet M.-A., Bonardet J.-L. and Fraissard J., Stud
Surf. Sci. Catal. 135 (2001) 93.
EPR STUDY OF PHOTO-INDUCED SURFACE MODIFICATIONS OF
NANOCRYST ALLINE Ti01 SAMPLES

J. M. CORONADO, A. J. MAIRA, J. C. CONESA AND J. SORIA


lnstituto de Catalisis y Petroleoquimica, CSIC, Campus de
Cantoblanco, Madrid 28049. Spain. E-mail: ;mcoronado@icp.csic.es.

1. Abstract

The effect of UV -laser irradiation in dynamic vacuum on the surface


characteristic of the nanostructured Ti0 2 is studied. For these purpose two titania
samples, with primary particle size of respectively 5 and 16 nm were outgassed at
RT, either in the dark or under UV-Iaser irradiation. Radicals formed by additional
UV -illumination at 77 K in static vacuum or in the presence of O 2 were monitored
by means of EPR spectroscopy. The obtained spectra reveal that the studied
samples are differently affected by laser illumination. In the case of the Ti02
materials with the smaller particle size some Te+ centers are detected upon
irradiation at 77 K in static vacuum, and the laser pretreatment greatly increases the
concentration of photogenerated radicals formed in the presence of oxygen,
especially subsurface 0' and O 2' species. However, in case of the Ti0 2 sample with
larger crystalline diameter no reduced titanium sites are detected and the effect of
irradiation in dynamic vacuum is not so remarkable; oxygen adsorption and
subsequent UV -illumination give rise to the formation of 02H, 0' and 0 3' radicals
are observed with different proportions depending on the pretreatment conditions.
These results are discussed in connection with structural characteristics of these
nano-sized materials, along with the possible photoinduced modifications of their
surfaces.

2. Introduction

The anatase phase of titanium dioxide is widely used as photocatalyst for


elimination of pollutants [1]. These processes are based on the reactivity of the
photogenerated charge carrier pairs (e' and h+) formed upon irradiation of a
semiconductor with light having wavenumber such that hV2!EBandgap' Other
technologically interesting properties of Ti0 2, like the formation of a highly
amphiphilic surface [2], also depend on the interaction of the UV light with this
oxide. Early reports revealed that UV -illumination induced the adsorption of
297
I Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 297-306.
© 2002 Kluwer Academic Publishers.
298
oxygen and the simultaneous removal of adsorbed water from Ti02 powders [3].
Several studies have shown that UV -irradiation of Ti02 samples results in the
formation of Te+ centers [2,4]. These altered areas of the surface are prone to
dissociate water and eventually increase the hydrophilicy of the Ti02 [2]. However
a more thoroughly understanding of the mechanism of these processes could aid to
effectively control the photo-induced modifications. Since many of the photo-
generated surface species are radicals, the EPR spectroscopy is particularly
adequate to gain information about these aspects.
The aim of the present work is to determine the effect of UV light on the
surface characteristics of Ti02 powders. Nanostructured anatase specimens,
prepared by either thermal or hydrothermal methods, have been used in this study
because of their excellent photocatalytic performance [5]. Parallel experiments
were undertaken with samples preconditioned in dynamic vacuum at room
temperature (RT), in the dark or under UV-laser irradiation. Simple outgassing
does not alter significantly the Ti02 surface, as it removes exclusively weakly
bound molecules. However, the combined action of vacuum and photoactivation
can shift the equilibrium toward the elimination of surface groups. These processes
presumably lead to restructured Ti02 surfaces, as the loss of adsorbed species
should be accommodated. Radicals formed by additional UV -irradiation at 77 K in
static vacuum or in the presence of O2 were monitored by means of EPR
spectroscopy. The analysis of the spectra obtained reveals photo-induced
differences, which are discussed with regards to the surface and structural
characteristics of the Ti02 samples.

3. Experimental

3.1. MATERIALS

Amorphoustitania gel aggregates (100 nm) were prepared by a sol-gel process


[5] using titanium isopropoxide (Aldrich). Subsequent hydrothermal treatment of
the amorphous material in a water-isopropanol mixture at 423 K produced a
crystalline sample labeled as P5. Another aliquot of the gel was fired at 773 K in
air for 3 hours yielding another sample of crystalline Ti02, named P16. XRD
measurements showed that both materials are pure anatase with a crystalline size of
respectively 5.2 nm and 16 nm for P5 and P16. In accordance to these data, the
BET area of the hydrothermal sample (204 m2.g. l ) is considerable larger than that
of the calcined material (34 m2.g- I ).
299
3.2. EPR MEASUREMENTS

EPR measurements were carried out in a Bruker ER200D instrument operating in


the X-band. Aliquots of the catalyst (~20 mg) were placed into a special cell made of
spectroscopically pure quartz and with greaseless taps for treatments. All the spectra
were recorded at 77K. in a TE104 type cavity. The frequency of the microwave was
calibrated for each experiment using a standard of DPPH (g=2.0036) located in a
second cavity. Computer simulations were used when necessary to check spectral
parameters and to establish the contribution of each signal.
Samples were pretreated in dynamic vacuum (continuous pumping, residual
pressure p<10 4 N.m-2) at RT (295 K) for I hour either in the dark or under
simultaneous irradiation with a pulsed N2 laser (LSI, VSL-337ND), emitting at 337.1
nm with a pulse energy of 250 J.1l. Despite the low frequency of the pulsed emission
(1Hz) a slight increment of the temperature (up to :::::305 K) was detected during this
treatment. Subsequent UV-irradiation under static vacuum (cell detached from the
pumping system) were carried out at liquid N2 temperature, placing the cell in a quartz
Dewar flask without any metallic coating. For these experiments the UV sources were
three fluorescent lamps (Osram Eversun L40WI79K), which emit their maximum
intensity at 350 nm. Additional measurements were carried out after irradiation at 77
K with O2 adsorbed on the Ti02. For these experiments a fixed a mount of O2
(~1211mols dose) was contacted with the sample at 77 K. Subsequently, a short
outgassing treatment was carried out in order to remove the excess of physisorbed O2,
which could result in dipolar broadening of the EPR signals.

4. Results and Discussion.

4.1 . EPR STUDY OF THE Ti02 SAMPLES OUTGASSED IN THE DARK

The two nanostructured Ti02 samples were submitted to a vacuum treatment in


the dark at RT and their EPR spectra w~re measured at 77 K. At this stage only
very weak signals at g~ge are observed for both studied materials (not shown). The
spectrum of sample P5 after UV irradiation at 77 K for 15 min. is displayed in
Figure la, it is constituted by a sharp axial signal T with g..t.=1.990 and gn=1.959.
This signal corresponds to Te+ centers in anatase [3], formed by trapping of
photogenerated electrons in Ti4+cations. Some poorly resolved features are also
observed at ca. g=2.012 and g=2.003. Subsequent oxygen adsorption at 77 K in the
dark results in the spectrum of Figure 1b, which shows the almost complete
depletion of signal T. This fact indicates that the sites that originate this feature are
located at or close to the Ti02 surface. Simultaneously to the removal of signal T a
new feature, A, with components at gl=2.027, g2=2.009 and g3=2.003, is formed
with almost the same amplitude. Signal A can be assigned to superoxide radicals
on the anatase surface [4] formed according to the reaction
T1o3 + + 0 2-+ T 14·+- 0 2- (1)
300

1.990....

a
b

/
2.013
2.035 ,

2.003
I
3350 3400 3450 3500

H/Gauss
Figure 1.- EPR spectra of the P5 sample outgassed in the dark at RT and a) subsequently UV-
irradiated at 77 K, b) contacted with O2 at 77 K and c) UV-irradiated again at 77 K

The presence of adsorbed oxygen while sample P5 is irradiated again at 77 K


originates a marked increase of the spectrum intensity, as shown in Figure Ic. This
overall increment of the amplitude indicates that oxygen, by acting as an electron
scavenger, hinders the recombination of the charge carriers. Spectrum of Figure Ic
can be fairly well reproduced by computer simulation considering three
contributions: signal A; signal B, with g)=2.024, g2=2.013 and g3=2.003; and signal
C, with g,=2.035, g2=2.009 and g3=2.002. Signal B can be related with holes
stabilized in Ti02 as subsurface 0- centers [6]. The influence of surface OK in the
stabilization of these 0 - species has been suggested, although in the case of P5
sample, that presents a very small crystal size, the possible existence of structural
defects could also play a significant role. Finally, signal C is attributed to
hydroperoxide radicals generated in the presence of acidic hydroxyls [7] according
to
Ti 4+-02- + H+ ~ Ti 4+-02H· (2)
The spectrum of sample P16 obtained upon irradiation in vacuum at 77 K, Figure
2a, shows a reduced intensity, and only poorly resolved features at g=2.016 and
301
g=2.003 are present, which very likely correspond to some of the signals observed
later with larger amplitude in Figure 2c. In contrast with the results obtained with
P5, signal T is not detected for this sample. This result suggests that the
stabilization of discrete Ti 3+ centers is favored in Ti02 samples with lower crystal
size. Oxygen adsorption after irradiation, Figure 2b, modifies slightly the shape of
the spectrum, due to the formation of a new signal D with a component at g~2 .007.
As in the case of P5 sample, UV irradiation in the presence of adsorbed oxygen
results in a considerable increment of the overall intensity of the spectrum, Figure
2c. Computer simulation shows that this spectrum is constituted by the overlapping
of at least three components, signal C; signal D with g)=2.011, g2=2.007 and
g3=2.002; and signal E with g)=2.029, g2=2.016 and g3=2.002. In addition, some
contribution of signal A cannot be discarded. Parameters of signal D are similar to
those reported for 0 3- species (0 2-0-) [7]. These radicals can be generated by hole
trapping on the Ti02 surface in the presence of O2 according to
Ti4+_02- + O2 + h+ - t Ti4+-0 3- (3)
Signals similar to E have been assigned to surface 0- species formed upon
irradiation of Ti02 samples with relatively large crystalline size [6]. Table I
summarizes the signals observed in the present study.

2.003
/
3350 3400 3450 3500
H/Gauss

Figure 2.- EPR spectra of the Pl6 sample outgassed in the dark at RT and a) subsequently
UV -irradiated at 77 K. b) contacted with O2 at 77 K and c) UV -irradiated again at 77 K
302

TABLE 1. Assignment of the EPR signals obtained.

Signal Components of the g tensor Assignment


A g)=2.027 g2=2.009 g3=2.003 T1·4+- 0-
2
B g)=2.024 g2=2.013 g3=2.003 Ti 4+-0" - Ti4+-OH-
C g)=2.035 g2=2.009 g3=2.002 Ti4+-02H
D g)=2.011 g2=2.007 g3=2.002 Ti 4+-03 -
E g)=2.029 g2=2.016 g3=2.002 Ti4+_0 2-- Ti 4+-0-
R g)=2.024 g2=2.009 g3=2.003 Ti4+-02- or ROO·
T g.t.=1.990 gll=1.959 Te+-

I
1.959
c

/
2.009

t 5x10-4 a.u

2.003
/

3350 3400 3450 3500

H/Gauss
Figure 3.- EPR spectra of the sample P5 a) laser-irradiated in dynamic vacuum at RT, and
b) subsequently UV-irradiated at 77 K, c) contacted with O 2 at 77 K and d) UV-irradiated
303

4.2. EPR STUDY OF THE Ti02 SAMPLES OUTGASSED UNDER LASER


IRRADIATION

The EPR spectrum of sample P5 laser irradiation in dynamic vacuum at RT for


hour is displayed in Figure 3a; it shows a weak signal, R, with parameters
gl=2.024, g2=2.009 and g3=2.003. Additional UV-Iamp irradiation at 77 K results
in the increment of the intensity of signal R, and the generation of signal T, with
slightly lower intensity that in the case of Figure lao Although the parameters of
signal R are rather close to those of the superoxide species on Ti02, the fact that it
appears in absence of oxygen and simultaneously with signal T, suggest that it may
correspond to species different than O2-, Considering that organic peroxide radicals
show g tensor components similar to those of R feature [7], this signal could
alternatively be ascribed to ROO· formed by hole trapping on organic groups,
which have not been entirely removed from the Ti02 surface after hydrothermal
treatment and drying. Additional experimental work should be necessary in order
to clarify this point. Oxygen adsorption at 77 K upon UV -irradiation at the same
temperature, Figure 3c, produces a relatively intense signal A, with gl at lower
field than the corresponding of R. At the same time signal T disappears almost

/
2.008

!2X10" a.u.

2.003
I
3350 3400 3450 3500

H/Gauss
Figure 4.- EPR spectra of the sample Pl6 laser-irradiated in dynamic vacuum at RT. and b)
subsequently UV-irradiated at 77 K. c) contacted with O 2 at 77 K and d) UV-irradiated
again at 77 K
304

completely. In contrast with the results obtained with the P5 sample outgassed in
the dark (Figure I), double integration of the spectra reveals that the amount of Ti-
O2• adducts formed by O2 adsorption is considerably larger than the number ofTe+
centers depleted. This fact indicates that processes different than (1) are operative
during oxygen adsorption. Consequently, additional EPR silent electron sources
must be available. A possible explanation is that under UV -laser irradiation two-
electron diamagnetic entities are generated in anionic vacancies. In the presence of
oxygen these centers may become destabilized, since O2- species are readily
formed due to the electronic affinity of oxygen. This increment of the
concentration of radicals indicates that laser irradiation of P5 leads to the reduction
of the e--h+ recombination rate. In any case, UV-illumination at 77 K in the
presence of adsorbed oxygen greatly enhances the overall intensity (Figure 3d) of
the spectrum, which shows contributions of signals A, Band C. This observation
points out the importance of adsorbed O2 as electron scavenger [I], that after the
laser treatment may access to the photogenerated electrons more easily. As in the
case of the P5 sample pretreated in the dark, signals A, Band C contribute to the
spectra. However, the proportion of superoxide species formed is larger when P5 is
irradiated with the laser during the outgassing treatment.
In the case of sample P16, the spectrum obtained after laser irradiation in
dynamic vacuum consists of a very weak signal with features at g~2.003 (not
shown). Subsequent UV -irradiation at 77 K leads (Figure 4a) to the formation of a
new signal E, which shows a broadening of the low field component. As in the case
of outgassing in the dark, Figure 2, there is no indication of the presence of Ti 3+
centers, suggesting that photo-generated electrons are delocalised in the conduction
band. A considerably increment of the spectrum amplitude is observed following
oxygen adsorption at 77 K. This is mainly due to the rise of the intensity of signal
E, along with the appearance of signals of the type D; some additional minor
contributions cannot be precluded. Comparison with the results obtained with P16
material outgassed in the dark indicates that laser treatment favors radicals
stabilization. The route that leads to the generation of 0 - and 0 3- surface species is
not obvious, because the formation of these species it is usually assumed to require
the participation of photogenerated holes. A feasible explanation can be proposed if
UV -irradiation results in the formation of a significant amount of hydrogen
peroxide according to
Ti 4 +-OR + h+ ~ Ti 4+-OH' (4)
2HO· ~ H20 2 (5)
Hydroxyl radicals are highly reactive and consequently they are not usually
detected by EPR, unless spin-trapping compounds are used. This fact could
account for the absence of signals attributable to these species. In any case,
subsequent oxygen adsorption in the dark could lead to the following processes
H20 2 + O2 ~ 0 3- + 0- + 2 H+ (6)
Although other reactions may yield the observed radicals, any conceivable
mechanism should consider EPR-silent species like hydrogen peroxide. Subsequent
irradiation at 77 K in the presence of adsorbed oxygen results in a rise of the
305

spectrum intensity, Figure 4c, whicn is mainly formed by the signals E, D and C.
This fact indicates again the crucial role of oxygen in hindering charge carriers
annihilation. In contrast with the results obtained with the sample P5, the
comparison of Figures 2c and Figure 4c shows that differences induced by the laser
treatment after UV-illumination in the presence of oxygen are rather limited. The
most remarkable variation is the lower proportion of 0 3- produced after the laser
pretreatment.
The expected effect of laser irradiation during evacuation is the elimination of
certain adsorbed species, resulting in the modification of the surface structure,
and/or the formation of peroxide species, reactions (4) and (5), in larger amount
than when illuminated with the UV-lamp. Although the present results does not
inform about the nature of these changes produced by the high photon flux of the
laser, EPR measurements provide evidences of an enhancement of the stability of
the charge carrier on Ti02 samples treated with laser.

5. Conclusions

The two studied samples, P5 and P16, are distinctly affected by the laser-
irradiation in dynamic vacuum. In the case of Ti0 2 material crystallized by the
hydrothermal method, P5, laser treatment produces centers where EPR-silent
electrons are stabilized. A feasible explanation is the formation of coupled
electrons pair in anionic vacancies. In any case these spec.ies are very effectively
captured by adsorbed oxygen. In the case of the calcined Ti02 sample, P16,
differences induced by the laser irradiation can be justified by considering the
accumulation of hydrogen peroxide on the surface. These results remark the
different characteristics of both Ti02 samples, that may related to their electronic
behavior. Thus, electrons can be found in discrete states in P5, whereas in the case
ofP16 they remain delocalised

6. Acknowledgment

This investigation has received financial support from ECSC (Project 7220-
EB/004) J.M.C. wants to thanks the Comunidad Aut6noma de Madrid by the award of
a postdoctoral grant.

7. References

1. Hoffman, M.R., Martin, S. T., Choi, W., and Bahnemann, D. W. (1995).


Photocatalysis over Semiconductors Chern. Rev. 95, 69.
2. Wang, R., Hashimoto, K., Fujishima, A., Chikuni, M., Kojima, E., Kitamura,
A., Shimohigoshi, M., and Watanabe, T. (1998) Photogeneration of Highly
Amphiphilic Ti0 2 Surfaces Adv. Mater. 10, 135.
3. Munuera, G., Rives-Arnau, V., and Saucedo, A.(1979) Photo-adsorption and
Photo-desorption of Oxygen on Highly Hydroxylated Ti0 2 Surfaces J. Chern.
306

Soc. Faraday Trans. 175, 136 (1979).


4. Howe, R.F., and Gratzel, M. (1985) EPR Observation of Trapped Electrons in
Colloidal Ti02, J Phys. Chern . 89,4495.
5. Maira, AJ., Yeung, K.L., Soria, l, Coronado, 1 M., Belver, e., Lee, e. Y., and
Augugliario, V. (2001) Gas-Phase Photo-Oxidation of Toluene Using
Nanometer-Size Ti02 Appl. Catal. B, 29, 327.
6. Nakaoka, Y., and Nosaka, Y. (1997) ESR Investigation into the Effects of Heat
Treatment and Crystal Structure on Radicals Produced over Irradiated Ti02
Powder J Photochem. Photobiol. A, 110,299.
7. Coronado, J. M., Maira, AJ., Conesa, le., Yeung, K.L., Augugliario, V., and
Soria, J. EPR Study of the Surface Characteristics of Nanostructured Ti02 under
UV -Irradiation Langmuir, 17,5368.
DOUBLE RESONANCE NMR STUDY OF MESOSCOPIC
INTERACTION BETWEEN SURFACTANT AND SILICA-
ALUMINA DURING THE DIRECT SYNTHESIS OF AISBA-15
MESOPOROUS SOLIDS

lB. D'ESPINOSE DE LA CAILLERIEa, E. HADDADb , A. GEDEON b

Laboratoire S.f.E.N. CNRS-FRE 2312


a Physique Quantique, ESPCI. 10 rue Vauquelin, 75231 Paris Cedex OS,
France
bUniversite P. et M. Curie,4 place Jussieu, 75252 Paris Cedex OS, France

1. Introduction

We have recently synthesized acid AlSBA-15 mesoporous solids with regular channels
and very high thermal and hydrothermal stability [I]. AlSBA-15 materials retain the
hexagonal order and physical properties of purely siliceous SBA-15. They present
higher thermal stability and catalytic activity in cumene cracking reaction than
AIMCM-41 solids. To better understand the origin of these improved properties,
textural results from N2 porosity measurements are confronted with molecular scale
double resonance MAS NMR results in order to discuss the incorporation of Al and the
interpenetration of the organic/inorganic phases during synthesis.
Indeed, recent publications have evidenced the significant occurrence of a
microporous "corona" around the internal surface of the mesopores in SBA-15 [2,3].
Considering that micropores result from the calcination of an incompletely hydrolyzed
silicate precursor, it is of primary importance, if one wants to be able to control the
extent of the microporosity, to understand at the molecular scale why the silicate
network did not fully condense: Is it because of the interpenetration of the hydrophilic
part of the surfactant with the forming inorganic phase [2]? Or is it because the
organometallic TEOS precursor was not fully mineralized prior to calcination?
To address this question, samples of different hydrolysis levels were prepared by
varying the maturation time. NMR established the speciation of aluminum as being part
of the silicate framework. It was then possible to investigate phase separation and
307
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 307-315.
© 2002 Kluwer Academic Publishers.
308

ordering in the parent material by NMR double resonance between the organic protons
and the aluminum of the solid, the results were then related to the structural properties
of the calcined final mesoporous AISBA- lS. This protocol was deemed preferable to
comparing samples prepared by varying the temperature, as this would have affect
simultaneously the hydrolysis kinetics and the hydrophilicity of the PEO fragments.

2. Experimental

2.1 MATERIALS AND SYNTHESIS

AI-containing SBA-15 mesoporous solids were synthesized by using tetraethyl


orthosilicate (TEOS), aluminum tri-tert-butoxide, and triblock poly( ethylene oxide)-
poly(propylene oxide)-poly(ethylene oxide) (E02oP07oE02o) Pluronic 123 copolymers.
In a typical synthesis, 9 mL tetraethyl orthosilicate (TEOS) and the calculated amount
of aluminum tri-tert-butoxide (in order to obtain a well defined Sil Al ratio equal to 10)
were added to 10 mL of HCI aqueous solution at pH=1.5. This solution was stirred for
over 3 h and then added to a second solution containing 4 g of the copolymers in 150
mL of HCI aqueous solution at pH=1.5 at 313 K. After being stirred for 3 hours the gel
solution was transferred into a Teflon bottle and heated at 100 °C for different reaction
times 0,6, 16 and 48 h. The solid products were filtered (parent composites) and finally
calcined (calcined samples) in air flow (9 L hoi) at 823 K for 4 h with a heating rate of
24 K hoi. In what follows, the samples are denoted AISBA-15.

2.2 MAS-NMR

Magic angle spinning nuclear magnetic resonance (MAS NMR) experiments wele
performed on a Bruker ASX500 spectrometer at 11.7 T. 27 Alone-pulse experiments
were performed at 14 kHz with a selective pulse «1t/6) duration of O.S j.l.s, 1 s recycle
time, and SOOO acquisitions. IH one-pulse experiments were performed at 13 kHz with a
1t/2 pulse duration of 5.5 Ils, 20 s recycle time, and 8 acquisitions.
The IH to 27 Al Cross-Polarization (CP) pulse scheme (Figure la) used to establish
2D heteronuclear correlation (HETCOR or WISE) [4] was the usual one. The
experimental conditions for CP to an abundant quadrupolar nuclei were determined by
using the same considerations developed in our study of alumina surfaces by IH to 27 Al
CP [5]. The spectra were obtained with radio frequency magnetic field strengths (Q /
21r) of approximately 44 kHz for IH and 90 kHz for 27 AI , and a spinning frequency (vr )
of 4 kHz. Other acquisition parameters are given in the legend of the figures together
with the one-pUlse spectra as inserts along the F, and F2 dimensions.
309

The IHe 7Al} TRAPDOR pulse scheme is given in Figure Ib [6]. The dipolar
dephased IH spin echo (So) is recorded after evolution at different integer multiples of
the rotation period and compared to the spin echo obtained under the same conditions
but with additional dephasing due to irradiation of 27 Al during half of the evolution
period (S). The 27 Al TRAPDOR dephasing of the IH signal (So-S)/So is then plotted as a
function of the evolution time. The radio frequency fields were 45 kHz and 140 kHz for
IH and 27 AI.

nl2 IH

O. nl2 n

.0
'H

--
t, CP

27Al
• D~ t Aq


27AI
CP Aq
\4 t2
••
Figure laoScheme of the HETCOR pulse Figure lb. Scheme of the TRAPDOR
sequence. pulse sequence.

2.3 STRUCTURAL CHARACTERIZATION

N2 adsorption/desorption isotherms were measured on a Micromeritics ASAP-2000


instrument at 77 OK. Specific surface areas were calculated from the adsorption
isotherms by the BET method, and pore size distributions from the desorption isotherms
by the BJH method. To assess the presence of micropores, we used the modified us-plot
analysis [7].

3. Results and discussion

3.1 STRUCTURAL PROPERTIES

The structural properties of AlSBA-15 samples have been analyzed in earlier studies by
Synchrotron XRD and transmission electron microscopy [8,9]. The five well-resolved
XRD peaks associated with 2D p6mm hexagonal symmetry and the well ordered
hexagonal arrays of mesoporous channels from TEM images indicate the high structural
quality of these materials.
The nitrogen adsorption-desorption isotherms have been used to obtain
310

infonnation about their mesoporosity. These isothenns, illustrated in figure 2, are


similar to those reported earlier [8]. The HI type hysteresis loops are situated in the
range 0.45<PlPo<0.9. The sharpness of the adsorption branches is indicative ofa narrow
mesopore size distribution. As the maturation time decreased, the N2 adsorption
isothenns preserved the same shape with a significant decrease of the hysteresis loop
towards lower PlPo values. This led to smaller surface area and mean pore diameter. The
pore size distributions (PSD) for the studied samples are shown in figure 3. They
indicated that the mesoporous channels were very regular with a narrow gaussian
distribution centered at 4.5, 7.0, and 8.2 nm respectively for samples at 0, 16 and 48 h.
of maturation time. The textural characteristics of the AlSBA-I5 samples are listed in
table!.

48h
0.10
48h
1000

0.08
800

16 h

no maturation

0.02
200

0.00 .J-,,....I.,--..L,......-,-.l..-.,..:;:::;::=,,;;;;;9"~
O+-~.-~'-~-r~-r~-' o 20 40 60 60 100 120 140
0.0 0.2 0.4 0.6 0.8 1.0
pore diameter (A)
relative pressure (P/Po)

Figure 2. N2 adsorption-desorption Figure 3. BJH pore size distribution of


isothenns for AISBA-I5 samples for AlSBA-I5 samples for different maturation
different maturation times. times.

The adsorption-desorption isothenns have been further used to evidence an


additional microporosity. The specific mesoporous volume and the micropore volume
were calculated using modified us-plots. The micropore volume evolved inversely to
the maturation time.
311

TABLE I. Structural-textural parameters of the AISBA-15 samples studied.

maturation S8£T Pore Pore micropore Mesopore


time (m2/g) a volume diameter volume b volume b
(cm 3/g) (om) V"'i (cm 3/g) Vme (cm 3/g)

none 525 0.60 4.5 0.045 0.536

16 h 648 0.96 7.0 0.010 0.696

48 h 908 1.41 8.3 0.002 0.972

a) total pore volume estimated from the amount adsorbed at PlPo = 0.99
b) micropore and mesopore volumes evaluated from the modified a -plot method.

3.2 SPECIATION OF ALUMINUM: ONE-PULSE AND 21AL{IH} HETCOR OF


CALCINED ALSBA-15

In one-pulse 27 Al NMR, three resonances appeared: IV Al at 52 ppm, vAl at 38 ppm, and


VIAl at 0 ppm (not shown). However, only the tetrahedral sites were part of the
aluminosilicate framework as the other resonances disappeared upon mild washing by
NH4Ci solutions after calcination (Figure 4, above the F2 axis).
On the same NH4 + exchanged sample, two correlations were established by
HETCOR between the IV Al resonance and F/ resonances at 4-5 ppm and 7 ppm. The
former corresponded again to the protons of physisorbed water and possibly of
unresolved hydrogen bound silanols, but the latter corresponded to a correlation with
the newly introduced protons of the NH4 + surface moiety. The proximity of tetrahedral
alumina to the surface charge balancing ammonium cations proved that IV Al occurred as
A1 3+ substitutions of Si 4+ in the oxide framework. Furthermore, it confirmed that at
least part of this IV Al is located near the surface. Incidentally, ammonium exchange
also allowed for the identification of the octahedral aluminum species. The octahedral
resonance almost totally disappeared from the quantitative one-pulse spectra. This
signified that it belonged to an extra-framework charge balancing cationic species and,
therefore, that no alumina (with IV Al as well as VIAl) phase was formed apart from the
(Al)IV containing aluminosilicate.
312

I
I

~ 1
a ~
·5

~
.., :
I

i I

} ---::------ ---------~GJ !'

.
5 E
Co
w ~ ---- ----------~ Co

~
10
II
a
15
Q
o
100 50 ·50
ppm:'AJ

Figure 4. 27 Al eH } WISE and one-pulse spectra of AISBA-lS matured 48 h, calcined,


and washed by NH 4 Cl. Contour lines are drawn at 15,30,45,60, 75, and 90% of the
maximum. Contact time (tcp): 1.S ms; number of acquisitions in t]: 1352; t, increments:
61 times IS ~ts .

3.3 RIGIDIFICATION OF THE TEMPLATE: ONE-PULSE IH NMR AND


DIPOLAR DEPHASING OF PARENT ALSBA-15

While the 27 Al spectra were identical irrespective of maturation time, the IH resonances
broadened during maturation (Figure S). This suggested an increased homonuclear
interaction that was indeed confirmed by IH Dipolar Dephasing experiments (Figure 6).
Analysis of the time dependence of the dipolar dephasing of the CH and CH 2 protons of
the surfactant revealed a common bi-exponential decay with two time constants (T2) of
about 10 ms and 0.7 ms. This reflected different segmental mobility for the PO and EO
backbone. More surprisingly, the relative weight of the faster relaxing component of
the 3.S ppm resonance increases from 60% to 7S% for the samples with no maturation
and with 48 h maturation respectively. Everything being equal, this meant that a larger
part of the backbone became rigid as maturation proceeded, thereby increasing
homodipolar interaction.
313

maturation time: • no rraturation


- - 011
. --- 48 h
~ 6h
... 16h
• : 48h

Z. 10
'1)5

c:
ppm'H ~ °O+~~--0~.5~--1'~~--1'.5~--2r.O~--2r.5~-'3.-0
't(ms)

Figure 5. IH one-pulse spectra Figure 6. IH spin echo amplitude So as a function


of AISBA-15 PEO-PPO-PEO of dipolar dephasing time of AISBA-15 parent
parent composites synthesized composites synthesized with different maturation
with different maturation times. Amplitudes were obtained from integration
times. of the two resolved lines of the IH spectra (Figure
5). Best bi-exponential fits of the T2 decay are
shown (see text).

3.4 MOLECULAR SCALE INTERACTION WITH THE TEMPLATE: IH eAL}


7
TRAPDOR OF PARENT ALSBA-15

As the template rigidified, the mesoporous surface area of the forming aluminosilicate
increased. The rigidification therefore reflected the larger interaction of the PEO blocks
with the larger aluminosilicate mesoporous surface. However, this could also be due to
an increased interpenetration between the organic polymer and the aluminosilicate
network as the t cations condense around the S+ chains interacting through the s+cn+
mechanism [10]. This would translate eventually also in a larger organic-inorganic
interaction but necessarily in a stronger interaction as the PEO fragments become
imbedded within the forming aluminosilicate.
To evidence such an increase in the strength of the interaction, the TRAPDOR effect on
the CH and CH 2 protons of the surfactant due to dephasing by the 27 Al of the
aluminosilicate was quantitatively followed for samples at different maturation time. It
appeared that the TRAPDOR effect and thus the strength of the organic-inorganic
314

interaction remained the same irrespective of maturation time and of the microporous
volume of the final AlSBA-15 material (Figure 7 and Table 1).

0.7

"•
0.6 •
(/)0
• •'"
&;- 0.5
••
if • •
"
E Figure 7 IHf7 Al } TRAPDOR

i
0.4 • • no maturation dephasing (see Figure 1b) as a function
'" % " 6h of the 27 Al irradiation time of the CH
• • 16h
and CH 2 protons AlSBA-I5 PEO-
0.3
:Ja • 48h
..." PPO-PEO parent composites
f- 0.2
~
."
•• "
synthesized with different maturation
times. Amplitudes were obtained from
0.1 • inter,ration of the two resolved lines of
the H spectra.
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3,0
l(ms)

This last result ruled out the possibility that the microporous "corona" in SBA-IS
material originates in the penetration of the hydrophilic PEO fragments of the template
within the forming silicate framework .

4. Conclusion

The occurrence of a dipolar interaction of constant strength between the CH and CH z


protons of the surfactant and the 27 Al of the aluminosilicate in the composite parent
material irrespective of the microporous volume of the final AISBA- IS solid precluded
the possibility that the microporosity ofSBA-15 originates from the occlusion of part of
the template in the oxide. The increased rigidity of the PPO-PEO-PPO backbone
evidenced by the increasing fraction of protons in strong homonuclear interaction must
be attributed to the forming mesoporous surface and thus to a larger amount of PPO
fragments interacting with it. Actually, the microporous volume decreased drastically
with maturation time as hydrolysis proceeded therefore suggesting that it resulted
instead from incomplete condensation due to the TEOS not being fully hydrolyzed. In
other words, the appearance of the microporous "corona" is not related to the surfactant
behavior but is only due to the hydrolysis kinetics ofTEOS under acidic conditions.
315

5. References

1. Yue, Y., Gedeon, A., Bonardet, J.-L., Melosh, N., d'Espinose, J.-B. and Fraissard, J.
(1999) Direct synthesis of AlSBA mesoporous molecular sieves: characterization and
catalytic activities, J. Chem. Commun. 19, 1967 - 1968.
2. Kruk, M., Jaroniec, M., Ko, C. H. and Ryoo, R. (2000) Characterization of the
porous structure ofSBA - 15, Chern. Mater. 12, 1961 -.1968.
3. Imperor-Clerc, M., Davidson, P. and Davidson, A. (2000) Existence ofa
Microporous Corona around the Mesopores of Silica - Based SBA - 15 Materials
Templated by Triblock Copolymers, J. Am. Chern. Soc. 122, 11925 - 11933.
4. Schmidt-Rohr, K., Spiess, H. W. (1994) Multidimensional Solid-State NMR and
Polymers, Academic Press, San Diego, 213 - 226.
5. Mertens de Wilmar, D., Clause, O. and d'Espinose de la Caillerie, J.-B. (1998)
Adaptation of 27 Al CPIMAS NMR to the Investigation of the Adsorption of Molybdate
Ions at the y - AhO) / Water Interface, J. Phys. Chem. B 102, 7023 - 7027.
6. Grey,C. and Vega, A. J. (1995) Determination of the Quadrupole Coupling Constant
of the Invisible Aluminium Spins in Zeolite HY with I H / 27 Al TRAPDOR NMR, J.
Am. Chern. Soc. 117,8232 - 8242.
7. Fenelonov, V. B., Romannikov, V. N. and Derevyankin, A. Y. (1999) Mesopore Size
and surface area calaculations of hexagonal mesophases (types MCM - 41, FSM - 16,
etc.) using low - angle XRD and adsorption data, Microp. Mesop. Mater. 28,57- 72.
8. Yue, Y., Gedeon, A., Bonardet, J.-L., d'Espinose, J.-B. and Fraissard, J. (2000)
Textural properties, stability and catalytic activity of acidic mesoporous AlSBA
materials obtained by direct incorporation of aluminium, Stud. Surf Sci. Catal. 130,
3035 - 3040.
9. Yue, Y., Gedeon, A., Bonardet, J.-L., Melosh, N., d'Espinose, J.-B. and Fraissard, J.
(2000) Direct incorporation of Al in SBA mesoporous materials: characterization,
stability and catalytic activity, Stud. Surf Sci. Catal. 129,209 - 218.
10. Zhao, D., Huo, Q., Feng, J., Chmelka, B. F. and Stucky, G. D. (1998) Nonionic
Triblock and Star Diblock Copolymer and Oligomeric Surfactant Synthesis of Highly
Ordered, hydrothermally stable, mesoporous silica structures, J. Am. Chem. Soc. 120,
6024 - 6036.
MR PROFILING OF DRYING IN ALKYD EMULSIONS:
ORIGINS OF SKIN FORMATION

J.-P. GORCE, J.L. KEDDIE, AND PJ. McDONALD


University of Surrey
Department of Physics, University of Surrey, Guildford,
Surrey GU27XH, UK

1. Introduction

Film formation from waterborne colloidal polymers is a complex phenomenon [1] that is
becoming better understood as a result of recent efforts employing magnetic resonance
profiling and imaging. The first reported use of MR microscopy to study drying (which is
the first step of the film formation process) in such a system was in 2000 [2]. This work
showed that in an alkyd emulsion cast on a substrate to form a convex drop, water flowed
laterally to replace the water depleted at the edges. The data were consistent with a model
that considered the lateral flow. MR microscopy has more recently been used in a
systematic study of the factors that influence the drying of colloidal films. Theory based
on .the lubrication approximation [3] predicts how particle size, film thickness, and
evaporation rate, among other factors, determine the time it takes for water to recede from
the edges of a wet film. Analysis of MR images [4] supports the prediction that uniform
lateral drying is favoured by larger particles, slower evaporation rates and lower film
thickness.

A permanent magnet with a large purposely-designed magnetic field gradient [5], known
as GARField (Gradient At Right-angles to the Field), has been a significant technical
development for studying waterborne coatings. It is ideally-suited for one-dimensional
profiling through planar samples, offering a pixel resolution of the order of 10 11m. The
GARField magnet has been employed successfully to determine the spatial variation of
photo-initiated crosslinking in latex films [6]. Significantly, the films were found to
crosslink faster in the central layers of the film. Near the air surface, the reaction of the
free radicals was inhibited by ingressing oxygen, whereas near the substrate, there was
initially insufficient light intensity to initiate the reaction. The results were consistent
with numerical modelling that took these factors into account. More recently, GARField
has been used to determine the influence of driers on the auto-oxidative crosslinking
profile in alkyd films [7] (polymer derived from natural oils). Important new insight was
gained into the crosslinking mechanisms. The drier combination affected the rate of
crosslinking in the early stages only, after which they were ineffective. Gradients in
crosslinking density emerged as a direct consequence of the oxygen concentration profile
that developed. Elsewhere, 'H and 2H solid-state NMR spectroscopy has been used to
317
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 317-326.
© 2002 Kluwer Academic Publishers.
318

examine latex dispersions [8,9]. In films dried homogeneously and having less than 16
wt% remaining water, three types of water were identified: external water, water at the
polymer/water interface and water inside the polymer itself.

Despite these advances, there is nevertheless much that is not fully understood about the
drying of waterborne colloids [10]. The drying of alkyd emulsions, in particular, raises
many questions. There is speculation in the literature [11,12] that there is an inversion
from an alkyd-in-water to a water-in-alkyd emulsion in the later stages of drying.
Although lateral drying is of clear scientific interest, in consumer products (e.g. paints,
coatings and adhesives), drying in the vertical direction is of greatest interest because non-
uniformity in this direction is usually not wanted in these applications. There is
uncertainty as to when non-uniformity develops in the water concentration in drying films
in the vertical direction. Although it has been proposed that this should depend on the
value of the Peclet number [13], experimental evidence is lacking. A closely related
question centres around the origins of the so-called "skin" layer (i.e. a continuous dry
surface layer above a wet layer) that sometimes develops in alkyd emulsions. Studies [14]
of drying in emulsions of silicone polymers, which are less viscous than the alkyds, have
shown the existence of a coalesced skin on macroscopic scales. Here we consider the
drying of alkyd emulsions in the vertical direction using the GARField magnet to address
these issues.

2. Experimental section

2.1.MATERIALS

The alkyd resins were obtained from an esterification reaction based on tall-oil fatty acid,
polyfunctional alcohols and acids. An alkyd emulsion was prepared via a phase-inversion
procedure using a commercial non-ionic surfactant (known as Atsurf E-3969) as
emulsifier. The solid content of this alkyd emulsion was 50 wt%. It is estimated from
chemical analysis that up to 1 wt% xylene, which is the solvent used during the
preparation of the alkyd resin, is present in the alkyd phase of the emulsion. The average
droplet size is of the order of 350 nm.

2.2.FILM PREPARATION

Thin glass coverslips (surface area of 20 mm by 20 mm), soaked in ethanol to improve


the wettability, were used as substrates. The alkyd emulsion was cast on these covers lips
using a 250 !lm applicator. Wet coatings produced in this way were initially between 200
and 250 !lm thick. After casting, the wet films were immediately placed in the magnet.
Similar films were cast on 10 cm by 10 cm glass plates and placed in a precision balance
for mass loss experiments.
319

2.3.NMR DEPTH PROFILING WITH GARFIELD

GARField [5] is a specially designed small permanent magnet used to investigate thin
films and thin slices of larger samples (thicknesses ranging from a few hundred
micrometers down to a few tens of micrometers). Its main advantage over more usual
solid state NMR techniques is its ability to study samples containing both solid and liquid
components. Therefore, GARField is particularly suitabl~ for the investigation of
waterbone coatings during their several drying stages. Wet coatings can be placed
horizontally between the pole pieces of the magnet above the radio-frequency coil used to
excite and detect the NMR signal. The main magnetic field , Bo' is parallel to the sample
plane. The magnetic field due to the radio-frequency coil, B., and the magnetic field
gradient, Gy ' are perpendicular to the sample plane (see Figure 1). At the position
corresponding to a main magnetic field of 0.7 T, the gradient strength is 17.5 T.m-I.
GARField is installed in a temperature controlled room fixed at 20 +/_ 1 0c. The relative
humidity (RH) was around 40 +/. 5 %. Within the instrument, the temperature is kept at 30
+/. 1 °c and the relative humidity is the same as the room.

Bo Bl Gy

Gravity
Film SamEle

r ~,
Coverslip RF Coil
~
Fi g ure I : Sche matic diagram of the GARField dc .. ign

The IH NMR signal was obtained from an excitation using a multiple echo sequence :
90.--r-(90y--r-echo--r-). with n=32 echoes and a pulse gap of -z=95fJ.S. The NMR signal was
averaged over 256 scans. The number of points per echoes was 128. The dwell time was
chosen to be 1.2 J..l.s. The repetition time was set to 0.3 s in order to obtain in a relatively
short amount of time a sufficiently high signal-to-noise ratio (proportional to the chosen
number of scans). Thus, "snap-shots" of the profile shapes as a function of time,
characteristic of the several stages of the film formation, were obtained. For each one of
the 32 echoes recorded, a profile of the sample was obtained by Fourier transform. For
the chosen parameters, the pixel resolution was 8.7 J..l.m. The decrease of the intensity at
each point of the 32 profiles is directly related to the effective T2 spin-spin relaxation time.
A single exponential function was used to estimate the T2 relaxation time for each position
along the profiles. The amplitude obtained from this single exponential fitting procedure
was used to estimate the magnetisation at each location, which is a measure of the IH
density. Profiles were normalised by an elastomer standard in order to correct for
sensitivity decline over the film thickness.
320
3. Results and Discussion

3 .I.MASS LOSS

Mass loss experiments provide a useful insight into the drying mechanism, but do not of
course, provide spatially resolved information. Over the first 50 minutes the rate of the
mass loss per unit area of the alkyd emulsion, as shown in Figure 2, is constant. This
mass loss is attributed to water evaporation. After about 50 minutes, the mass loss rate
starts to slow down. At later times, after about 80 minutes, the mass loss is very slow.
The origin of this slow mass loss is not known, but it could be due to the loss of residual
solvent, such as xylene, or trapped water. Hence, MR profiling was used to investigate
this observation.

Time (minutes)
o 50 100 150 200 250 300
O~----~----~----~----~------~--~

7 .................................. . .

Figure 2 : Mass loss per unit area for an alkyd emulsion film as a function of time, 23°C and 37% RH

3.2.MR PROFILING

Figure 3 shows a plot of tH density profiles over the entire range of film formation (from
time t =0 until the time when no more change is observed, t "" 103 min). During the first
13 minutes, the thickness of the wet coating decreases linearly with time. This decrease is
estimated to be near to 9x1O·8 m s·t. In similar conditions, Hisatake et al [15] found that
the value of the evaporation of water into air (at 29°C with a relative humidity of 55% and
air flow velocity of 1.5 m s·t ) is 0.715xl0·3 g cm·2 min· t. This value is equivalent to a
thickness variation of 12x1Oog m sot, in agreement with our observation, considering that
the air was static above the sample in the magnet. After about 18 minutes, the thickness
does not decrease significantly, indicating that most of the water has evaporated.
321

• <> .0 minute
-9-7 minutes
. ... ·13 minutes
-+- 58 minutes
G ' ..,.. . 128 minutes
__ 1014 minutes
Q

" ~ Evaporative drYi~g

\ ~
. ;.

-25 0 25 50 75 100 125 150 175 200 225 250 275


Height (J.1m)

Figure 3 : Proton density profiles obtained from an alkyd emulsion film over time, as indicated

Figure 4 shows T2 profiles obtained from the same alkyd emulsion as in Figure 3. The
signal-to-noise level is not sufficient to perform a two-component (alkyd and water) fit to
the data. Instead, a single T2 relaxation time was obtained. We judge this to be acceptable
because diffusion attenuation of the NMR signal of the very much more mobile water in
the strong gradient leads to an effective relaxation time comparable to that of the more
viscous alkyd.
2.75
..... 18 minutes
2.5 ........... ..... ....... .•.. .... --a.- 38 minutes
-;;- 2.25 .' .......... . . 8
~.. .., /\ .•... .~'.. ..,. .. 5 minutes
§ 2 !f"'-'~-::"'" \._~ . ""*-73 minutes I
,, 0 ·, 1014 minutes I
e<>
.;:)
1.75 \ .... .... .• • •
..
\
i
c 1.5
0
.~ 1.25 ~" I
~ 1
'E
~
0.75
0.5
.'1 I
0.25
0
·25 o 25 50 75 100 125 150
Height (J.1m)

Figure 4 : T, profiles obtained from the alkyd emulsion film as a function of time

At a time of 18 minutes, T2 is approximately uniform with depth into the film. Soon
thereafter, the T2 at the air surface decreases abruptly from a value of about 2.2 ms to a
value of 0.9 ms. A relatively-sharp step in T2 develops. We will refer to the surface layer
with a lower T2 value as a "skin" . The thickness of the skin layer increases with time. At
later times (longer than 190 minutes), the T2 is uniform with depth and stabilises in value.
The skin has effectively grown to the entire depth of the film . In the next two sections,
the origins of the skin layer are considered.
322

3.3.RESIDUAL XYLENE

One interpretation of the T2 profile development is that the molecular mobility of the
alkyd decreases near the air surface. It is known from independent analysis that the alkyd
contains trace amounts of xylene, and perhaps other solvents. Solvents are known to have
a plasticising effect on polymers and hence to increase T 2. One hypothesis is that the skin
consists of neat alkyd from which solvent has evaporated that lies in a layer above alkyd
that still contains trace amounts of solvent. The skin layer would then grow as solvent
evaporates. In this mechanism, the solvent has to diffuse through the increasingly thick
top layer of neat alkyd. Simultaneously with the solvent evaporation, the mass of the film
would be expected to gradually decrease. This can be observed in the mass loss
experiment. Indeed, over a time period of 80 minutes, the alkyd emulsion film was found
to decrease in mass by around 4.5 %, as shown in Figure 2. However, this value is more
than four times greater than the maximum estimated initial xylene concentration (i.e. 1
wt%).

Further IH NMR experiments were conducted to determine whether the loss of remnant
solvent could explain the observed decrease in T2 as shown in Figure 4. A Maran
permanent magnet spectrometer (0.5 T tuned to 21 MHz) from Resonance Instruments
Ltd was used to systematically determine T2 as a function of xylene concentration in the
alkyd resin . Experiments used a CPMG sequence having 32 echoes, a pulse gap of 1=64
f.1s, 16 scans, a dwell time of 1 f.1s, 32 points and a repetition time of 2 s. Alkyd resins
were placed in a 1 cm diameter test tube. Careful examination of the data found that it
was best described by a two-component exponential decay, yielding two values of Tr
The higher value is typically on the order of 2.1 ms (volume fraction of 32%) , whereas
the lower component was 0.4 ms (volume fraction of 68%). At the same time, TI' the
spin-lattice relaxation time, was determined using an inversion recovery pulse sequence.
Using a single exponential fitting procedure, TI was found equal to 111 ms. In order to
compare the T2 results obtained from the bench top instrument to the GARField T2 profiles,
the data were also described by a single component fit. Figure 5 shows that the effective
T2 values vary approximately linearly with xylene concentration over the range studied.
Extrapolation of these data predicts that the T2 of the neat alkyd is 0.9 ms. This predicted
value is comparable to that obtained in the alkyd emulsion films after 190 minutes, which
indicates that if xylene had been present in the emulsion it has evaporated from the film at
that time.

Comparison of the results in Figure 4 and Figure 5 and extrapolation to the higher
concentration can provide an estimate of the change in concentration of xylene that would
be required to induce the observed decrease in T2 from 2.2 to 0.9 ms. The results suggest
that the xylene concentration would need to decrease from 4.8 % to nearly 0 % in order to
explain the development of the skin layer. Once again, it is highly unlikely that the
original emulsions contained such a high fraction of solvent.
323

1.5 -.--.............-.~ ..................~

0.9 f-----..----,r----,------r--.------..----,-----I
o 0.25 0.5 0.75 1 1.25 1.5 1.75 2
Xylene concentration (wt%)

Figure 5 : T, relaxation time as a function of xylene concentration present in the alkyd resin using single
component analysis

As a further test, an emulsion was prepared in which 5 wt% xylene was purposely added
to the alkyd phase. After evaporation of most of the water at 15 minutes, the alkyd film T2
obtained with GARField was uniform with depth and equal to about 2.5 ms. This value is
approximately the same as obtained in the previous experiments. This result suggests that
even if xylene is present in the alkyd emulsion, it evaporates quickly during film
formation and therefore is no longer present by the time when most of the water has
evaporated.

Finally, an unsuccessful attempt was made to estimate the concentration of xylene in an


alkyd film over time using confocal Raman Microscopy. Raman spectra obtained from a
day-old film showed strong bands at around 1603 cm'! and 1030 cm'! characteristic of the
fundamental modes of vibration of phenyl groups [16]. Because xylene is not expected to
still be present in a day-old film, this result indicates the presence of phenyl groups within
the structure of the alkyd resin itself. Specifically, phthalic anhydride is likely to be
present. The detection of less than 1 wt% variation in xylene content in a newly cast
alkyd film was therefore not achievable in such a chemical system.

3.4.RESIDUAL WATER

An alternative hypothesis for the origin of the T2 profile evolution is that the assumption
of a single T2 component when fitting the data is incorrect. If, in fact, a small amount of
water was trapped in the emulsion, it might contribute disproportionately to the signal.
The single T2 could be influenced by the presence of the alkyd, any trapped water, and
even the surfactant. The 4.5 % mass loss observed after 80 minutes could thus be due to
the slow evaporation of trapped water.

A recent process model of film formation, developed by Routh and Russel [13], supports
the idea that a skin layer will develop in the alkyd emulsion films . Skin formation results
324

when there is non-uniform drying coupled with particles that undergo wet sintering. Non-
uniformity in water concentration in the vertical direction is expected when the rate of the
surface recession is greater than the rate at which particles re-distribute themselves by
Brownian motion. This condition is met when Pe, the Peclet number, given as
61rpRHElkT is »1. Here ).l is the viscosity of the continuous medium (i.e. water in our
case with ).l "" 1 mPa s), R is the particle radius (about 175 nm), H is the initial film
thickness (210 /lm) , E is the evaporation rate (given in units of velocity: 9xlO·8 m S·I), k is
the Boltzmann constant, and T is the absolute temperature (303 K). In the alkyd
emulsions studied here, Pe is therefore about 15, predicting that there will be non-uniform
drying profiles. At room temperature, the alkyd resin is well above its glass transition
temperature of -35°C, and can be observed to flow over macroscopic distances.
Consequently, the predicted mechanism of particle deformation is by wet sintering [13],
and skin formation is expected.

Figure 6 shows a plot of the thickness of the skin layer as a function of the square root of
time. The origin of the time scale was chosen from the onset of the "skin" layer
formation. The thickness varies linearly with the square root of time. This may suggest a
diffusion process by which the residual water is diffusing through the skin layer and then
evaporating.
90

80

S 70
~60
'"~ 50
12
. ~ 40
.c
';; 30
:.i2
CI) 20

10
O&-~.--.~-.---.--.---.--'r--.---.--.-~

o 2 3 4 5 6 7 8 9 10 II
Time ln (minll2)

Figure 6 : Thickness of the "skin" layer as a function of the square root of the time since skin formation

As a preliminary analysis, we assume that the time required for water to diffuse through
a skin layer of thickness x is proportional to ilD, where D is an effective diffusion
coefficient for water in the alkyd. The diffusion coefficient is thereby estimated to be
about 1.5xlO·9 cm2 S· I. This value can be compared to the values ranging between 8.57 and
0.52xlOo9 cm 2 sol determined at room temperature for water diffusion through PET films
[17] . The estimated D is therefore not an unreasonable value for water diffusing through a
hydrocarbon polymer. The sharp front of the T2 profiles delimiting the wet (trapped
water) and dry layers in the alkyd film suggests a strongly concentration dependent
diffusivity.
325

We therefore conclude that the abrupt decrease in T2 near the film surface from 18
minutes onwards is likely to be attributed to the development of a skin of neat alkyd. As
water diffuses through the skin and evaporates, the thickness of the skin layer increases.

4. Summary and conclusions

For the first time, NMR profiling has been successfully used to study vertical profiles CH
density and T2 relaxation time) during alkyd emulsion film formation with better than 10
11m pixel resolution.

From a basic mass loss experiment, we have shown that in the later stage of the film
formation, evaporation of a few percent of either solvent (initially present in the alkyd
resin) or water trapped below a skin layer occurs. MR profiling of alkyd emulsions found
that a skin layer having a lower T2 developed during the later stages of drying. It grew in
thickness over time.

Our first hypothesis was that these developing profile shapes were related to the
evaporation of xylene from successively deeper regions of the film. Xylene is known to
plasticise the alkyd, therefore it could lead to more mobile polymer chains and so to
higher T2 values. Experiments using a NMR benchtop spectrometer indicated that the
large difference in T2 between the skin and the rest of the film would require an
unrealistically high xylene concentration (4.8 wt%). An experiment on an emulsion
enriched with xylene found that xylene evaporates from the film quicker than first
expected.

Our second hypothesis was that a small amount of water was trapped below a skin layer.
This hypothesis is supported by the predictions of the process model of Routh and Russel
[13]. From the mass loss experiment we conclude that up to 5 wt% of water may be
trapped in the film once a skin is formed. The water has to diffuse through this skin layer,
the thickness of which was found to increase linearly with the square root of time. The
data are consistent with a diffusion coefficient of water in alkyd of 1.5xlO-9 cm2 S-I. This
interpretation is more probable than the xylene hypothesis.

These encouraging results suggest ideas for further investigation. GARField experiments
could be performed in order to gain additional information on the initial stages of alkyd
film formation. Alkyd emulsions prepared using Dp could strongly reinforce the
hypothesis of water diffusion through an increasingly thick skin layer. Indeed, profiles
obtained from such alkyd films drying in controlled atmosphere should be free of any
non-uniformity. On the other hand, if such films were cast and dried in the normal
atmosphere, the uptake of protons from the water vapour of the air could help to estimate
the time scale and the amount of water trapped in the alkyd film. Infrared ellipsometry
could also be used to confirm the presence of trapped water. Finally, by varying
experimental parameters such as temperature, relative humidity, air flow, solid content
and initial film thickness, we could use GARField profiles to test the theoretical models of
film formation from colloidal dispersions.
326
5. Acknowledgements

We thank Drs. Amanda Barry, Eli Ciampi, Graham Bennett and Jacky Mallegol
(University of Surrey), Dr. Alexander Routh (University of Bristol) as well as Drs. Gary
Jefferson, David Bovey, David Taylor and Peter Palasz (ICI Paints, Slough, UK) for
experimental assistance and useful discussions. We are grateful to ICI Paints as well as to
the EC Framework V Programme MARWINGCA (ref: QLKS-1999-01587) for financial
support.

6. References

I. Keddie, J.L. (1997) Film formation of latex, Mater. Sci. Eng. Rep. R2113, 101-170
2. Ciampi, E., Goerke, V., Keddie, 1L. and McDonald, PJ. (2000) Lateral transport of water during drying of
alkyd emulsions, Langmuir 16,1057-1065.
3. Routh, A.F. and Russel, W.B. (1998) Horizontal drying fronts during solvent evaporation from latex films,
A1ChE J. 44, 2088-2098.
4. Salamanca, 1M., Ciampi, E., Faux, D.A., Glover, P.M., McDonald, PJ., Routh, A.F., Peters, A.C.I.A.,
Satguru, R. and Keddie, J.L. (2001) Lateral drying in thick films of waterborne colloidal particles, Langmuir 17,
3202-3207.
5. Glover, P.M., Aptaker, P.S., Bowler, J.R., Ciampi, E. and McDonald, PJ. (1999) A novel high-gradient
permanent magnet for the profiling of planar films and coatings, Journal of Magnetic Resonance 139,90-97.
6. Wallin, M., Glover, P.M., Hellgren, A.c., Keddie, J.L. and McDonald, PJ. (2000) Depth profiles of polymer
mobility during the film formation of a latex dispersion undergoing photoinitiated cross-linking,
Macromolecules 33, 8443-8452.
7. Mallegol, J., Barry, A.M., Ciampi, E., Glover, P.M., McDonald, PJ., Keddie 1L., Wallin, M., Motiejauskaite,
A. and Weissenborn, P.K. (submitted) Influence of drier combination on through-drying in waterborne alkyd
emulsion coatings observed with magnetic resonance profiling, J. Coat. Technol.
8. Rottstegge, 1, Landfester, K., Spiess, H.W. and Heldmann, C. (2000) Different types of water in the film
formation process oflatex dispersions as detected by solid-state nuclear magnetic resonance spectroscopy,
Colloid and Polymer Science 278. 236-244.
9. Baumgart, T., Cramer, S., Jahr, T., Veniaminov, A., Adams, J., Fuhrmann, J., Jeschke, G., Wiesner, V .,
Spiess, H.W., Bartsch, E. and Sillescu, H. (2000) Film-forming colloidal dispersions studied by tracer methods,
Macromolecular Symposia 151,451-457.
10. HolI, Y., Keddie, J.L., McDonald, P.1 and Winnik, M.A. (2001) Film formation in coatings: mechanism,
properties and morphology, ACS symposium Series 790, Provder, T. and Vrban, M.w. Editors, pp. 2-26.
II. Beetsma, 1. (1998) Alkyd emulsion paints: properties, challenges and solutions, Pigment & Resin
Technology 27, 12-19.
12. Beetsma, 1. and Hofland, A. (1996) Alkyd paints - from the ease of organic solvents to the difficulties of
water, Paint & Resin International 1, 15-17.
13. Routh, A.F. and Russel, W.B. (1999) A process model for latex film formation: limiting regimes for
individual driving forces, Langmuir 15, 7762-7773.
14. Guigner, D., Fischer, C. and HolI, Y. (2001) Film formation from concentrated reactive silicone emulsion. I.
Drying Mechanism, Langmuir 17,3598-3606.
15. Hisatake, K., Tanaka, S. and Aizawa Y. (1993) Evaporation rate of water in a vessel, J. App/. Phys. 73,7395-
7401.
16. Wilmshurst, 1.K. and Bernstein, H.J. (1957) The infrared and Raman spectra Of toluene, tOluene-a-d" m-
xylene, and m-xylene-aa'-d" Can. J. Chem. 35,911-925.
17. Sammon, c., Yarwood, J. and Everall, N: (2000) A FfIR-ATR study of liquid diffusion processes in PET
films: comparison of water with simple alcohols, Polymer 41, 2521-2534.
CATCHING A FALLING DROP BY NMR: CORRELATION OF POSITION
AND VELOCITY

SONG-I HAN, SIEGFRIED STAPF and BERNHARD BLUMICH


Institute for Technical Chemistry and Macromolecular Chemistry
RWTH, Worringerweg 1,52074 Aachen, Germany

The spin density profile and the statistical probability density of displacements of a drop
free falling through a vertical-bore magnet was measured in all three spatial directions,
both separated and in a correlated way employing NMR imaging methods and in
particular pulsed field gradient (PFG) NMR techniques. The falling motion was
monitored by double encoding of vertical position at successive times separated by D.
within a single two-dimensional (2D) PFG NMR experiment i.e. position exchange
spectroscopy (POXSY). 2D images along the cross-section of the drop, mapping the z-
velocity component in each pixel, verifies the constant z-velocity throughout the total
drop volume. 2D images encoded with transverse velocities contain structures
characteristic of velocity distributions which can be attributed either to vortex-motions
inside the drop or coherent oscillations of the whole drop. The fact that the net
transverse velocity over the cross-section is zero proves that no net motions of the drop
itself contribute to the results. The essential challenge has been to implement one- and
multi-dimensional NMR imaging techniques within the residence time of the falling
drop inside the resonator of about 10 ms. The solution was to accumulate information
from many drops by triggering every single acquisition to another falling drop. The
regularity and uniformity of the dripping process was clearly demonstrated, which was
the absolute requirement for signal accumulation as well as for multidimensional NMR
experiments. Given that the experimentally determined velocity profile of the drop by
NMR imaging methods was reproducible and reliable, the basis for mapping motion of
and inside a drop in a direct, three-dimensional and non-invasive way by NMR flow
imaging methods is given.

1. Introduction

Although a falling water drop is a common sight, the mechanism which leads to its
generation and determines its shape and dynamics remain a challenge for scientists. The
motion inside a falling drop cannot be described by simple hydrodynamics. This can be
attributed to the special feature of a falling drop: the continuous building of new and
disappearing of old interfaces and the rate .of deformation being dominated by the
interfacial tension [I]. The study of such drop dynamics is motivated not only from the
327
1. Fraissard and O. Lapina (eels,). Magnetic Resonance in Colloid and Interface Science. 327-336.
© 2002 Kluwer Academic Publishers.
328

point of view of fundamental research, but also in the light of a multitude of technical
processes, where detailed knowledge of the inner convection motion is of vital
. importance. Examples of applications
a) b) are the production of sprays and
emulsions, as well as the function of
inkjets [2, 3].
One of the most important
applications are the chemical extraction
[4], which is a fundamental process in
I jet
the purification of chemical
1/ compounds. The extraction efficiency is
photo sensor
I /
I
significantly influenced by the inner
dynamics of the moving drops. The
single free falling drop -although not
: resonator obeying simple hydrodynamics at all-
. IV is the simplest model for studying

-, extraction phenomena. The fundamental


difficulty to predict and analyze
quantitatively the rheology of interfaces
with standard optical methods is given
by the fact that conclusions about
complex and microscopic process must
be drawn from a macroscopic property,
Figure 1. The experimental set-up for the drop
apparatus contains a glass jet plugged in a long glas in this case the surface contour of the
tube hanging inside the magnet and a pair of photo falling drop [5,6]. More directly but
sensors placed 30 mm below the tip of the jet, which is invasively, convective motion can be
connected to the trigger input of the spectrometer. The studied by particle tracer techniques [7-
glass apparatus is held by a teflon-spacer in the
resonator, the center of which is set 200 mm below the 9]. Here NMR microscopy combined
tip of the glass-jet. a) Photograph; b) Schematic with pulsed field gradient spin echo
drawing of the drop apparatus. (PGSE) methods [10-12] come into the
perspective as the only method being
able to observe the shape and to investigate the motion of and inside the falling drop in
a non-invasive, direct and correlated manner, as long as identical drops can be
generated repetitively over a long period.
Any field gradient G(r) superimposed to the static magnetic field Bo has the effect
of producing a spatially dependent magnetic field offset which leads to an encoding of
position via the relation of the Larmor precession frequency, roL(r) = Bo + yG(r)rwhere
y is the gyromagnetic ratio. This information can be exploited directly by signal
acquisition in the presence of the gradient, where the spin density is obtained from a
Fourier transformation ofthe signal from the time domain into the frequency domain -
the frequency encoding method-, or indirectly by applying short gradient pulses which
lead to a phase shift of the signal of each position-dependent spin isochromat -the phase
encoding method [13-15].
329
2. Experimental Part

Experiments were performed on a vertical 4.7 T, 150 mm bore magnet equipped with a
Bruker DSX-200 spectrometer. The setup consists of a glass jet, mounted into a long
glass tube, which is hanging inside the magnet, and an infrared photo sensor placed
30 mm below the tip of the jet, which triggers the spectrometer each time the drop
obstructs the light path. A high-precision pump (Pharmacia P-500) maintains a closed
circuit and a constant filling height in the reservoir above the jet (Fig. 1).
The drop consisting of 2.8 % aqueous surfactant (betain) solution moves in free
fall through the air-filled open structure of the resonator. For the experiments presented
in this work, the center of the birdcage resonator lies about 200 mm below the tip of the
jet; hence a falling velocity of about 2 mls results and the drop remains only for about
10 ms within the sensitive volume of the birdcage ·resonator. As a consequence, any
ultra-fast NMR imaging method will fail to complete the measurement during this time;
it remains not much more time than to acquire one echo per drop. Clearly, echo and
acquisition times as short as possible are desirable for all measurements. Using a
triggered cycle was the solution for signal accumulation and for multi-dimensional
experiments. Because every single scan had to be collected from different drops, the
a) ..... 900 1800 d) ......OJJ 90 0 1800
OJJ
OJ>
'5
echo
..
OJJ
.;:
echo

G read CJ Gphase

I I
n n
b) G G II III II
8 !:l
Gphase
II M II
e) ....
OJJ
OJJ
900 1800
echo
8 6. '5
...

I
900 1800 tt
C) "OJJOJJ echo
'E
Gphasel G
'

G phase

I I
II III II
G phase2

.f
0
IIf
I II
G
'

8 6. 6.

Figure 2. Pulse sequences used in the present work: a) Frequency encoding method for position encoding b)
Bipolar gradient pulse pair with strength G, separated by time .d and with a pulse duration of 0 in order to
encode the NMR phase by velocity along the gradient direction. c)-d) PGSE sequences for double encoding
of positions, where the gradient pulses are stepped simultaneously as indicated by the arrows. c) Phase
encoding method to measure the average position during.d. d) PGSE methodfor measuring the propagator
of displacements or the distribution of averaged velocity during .d. e) POXSY pulse sequence, where two
gradients are applied along the same spatial direction with the strength G I, G2, pulse duration 0 and time
separation .d. The gradient pulses are stepped independently of each other resulting in a 2D map.
330

j
I
f~qUency ~ncoding b) 1.0 x-dlre ion - phase encoding ,
- lscan . 0.9 - frequenC\' encoding -j
- 2scan ne ~
1
-- 3 scan
4scan Q7 i
U ~

Q5 lj
0.4
0.3
0.2
Ql
°o~~~~~~---L---J--~~~~
i
posit... [mm)4
c) 1.0 f y-diredlon seencoding d) 1.0 z-dire ion - p se encOding .
0.9 , - frequenC\' encoding 0.9 - frequency encoding j
08 r
f
0.8
]
0.7 [-, 0.7
0.6 ~
0.5 ~
0.6
0.5
]
i
r
0.4 f 0.4
03 ~
02 ~
0.3
0.2
j
0.1 ~ 0.1
O~o~~~--~--~--~--~~~~ O~~~L-~--~--~--~~~
P05iJ... [mm) 4 6
o

Figure 3. 1D profiles along the X-, y- and z-directions. a) 1D profiles along y obtained by the frequency
encoding method with 1-4 accumulations. b)-d) 1D profiles obtained by the frequency encoding method
(dashed lines) superimposed with those obtained by the phase encoding method (solid lines) along x, yand
z.

first aspect to prove was the periodicity and unifonnity of the dripping process. The
reproducibility of the drop fonnation and the unifonnity in the shape of each drop are
closely linked to each other and are indispensable conditions for reliable NMR
experiments. Otherwise random contributions would lead to a partial extinction of the
accumulated signal.
The conventional NMR frequency encoding method (Fig. 2a) to measure one-
dimensional (lD) projections along any spatial axis was repeated with different
numbers of accumulations and the results are presented in Fig. 3a. In one shot, 32 data-
points per projection were acquired with a total acquisition time of 650 J.1s. The I D
profiles along the y-direction resulting from 1, 2, 3 and n accumulations, respectively,
i.e. collected from as many drops as the number of accumulations, differ only in their
amplitude, which is found to be strictly proportional to the number of accumulations.
This excludes random variation in the dripping process. The variance calculated
between the profiles from several experiments was less than 1%. In the next step I D
profiles were measured employing the phase encoding method (Fig. 2c), where each of
the 65 points in the image dimension is obtained by incrementing the amplitude of the
gradient pulse in a triggered cycle, so that as many drops as the number of image pixels
constitutes the signal, giving the spin-density profile after Fourier transfonnation. The
331

a) 2.Mo6,.....,..........~,......,......,.~...,.....,........,,.............,....,~..................~.,.............,.,............,.....,........,,.............,....,........,
a [ms] A[ms] G[T/m] fof[m/s]
0.2 1.2 1.0 9.0
~ I.5xlo6 - ---
0.2 2.4 1.0 4.5 • -
::j
"
.!!t 0.2 4.3 1.0 2.5 ~
~ 1.Oxlo6-
'iii z-direction
-
c
.2l
.s 5.Oxlfrl f- -
o ~ j
-4.52 -3.S7 -3.23 -2.5S-1.94 -1.29 -0.65 0.00 0.65 1.29 1.94 2.5S 3.23 3.S74.52
velocity [m/s]
b) 2.Oxlo6r--..........-.,..............,....,...........,................,...~.,.........,..........- ......................-..,...............~................".........,
alms] A[ms] G[T/m] fof[m/s]
0.4 2.4 1.0 0.83
- - x-direction
- .. _ .. y-direction
- - _. self-diffusion

?1~00~-~S6~--~71~--5~7--4-3~~---1~4--:wO~-1"'4-~':-"4~3-~5~7-"'7"'1- ....S6-...l100
velocity [mm/s]
Figure 4. Propagators along the x, y and z direction. 65 data-points per propagator and 8 accumulation
for signal averaging were acquired. a) z- propagator with three different ranges of the fo! b) x (solid
line), y propagator (dashed line) and the computed propagator of self-diffusionfor comparison.

profiles obtained by the methods of phase encoding and frequency encoding are
compared in Figs. 3b-d. The agreement of the profiles with the drop diameter of
3.5 mm in all three spatial directions as also presented in Fig. 3b-d proves the uniform
dripping behavior. The equivalence of the results further demonstrates the complete
interchangeability of both methods despite their inherently different susceptibility to
motion during the imaging sequence. It can thus be concluded that the dripping process
is highly uniform and that distributions of position information due to motion artefacts
can safely be neglected.
Measurements of the probability density of displacements, or the propagator [10]
(Fig. 2d), in x, y and z directions were realized employing PGSE methods (Fig. 4). Here
it is even more important to have the stability of the drop under control, otherwise the
macroscopic fluctuations of the drops themselves would dominate the results. The z
velocity distribution confirms the falling velocity of 2.0 mls and its full width at half
maximum (FWHM) is found to be ± 0.04 mls (Fig. 4a). The transverse velocity
distribution has a FWHM of ± 17 mmls and a net velocity of zero (Fig. 4b). This
observation is an indirect proof that on the one hand the drop falls .vertically , resulting
in a net displacement of zero in transverse direction and on the other hand convective
332

motion caused by small vortices inside the drop is present. The rms displacement due to
self-diffusion amounts to <X2> = (2.D·~fr, ~ 3J.lm and would contribute only with
"Vdiff "= ± 1.3 mmls in average to the propagator under the given experimental

:O:: ~~"~~f" ' 'I " ~0


7 .00 ::: • ~ bility fu nction 3
~ < ~
§ 3
6 . 00 . 3
5.00 -
= ~
~

~
to j
3 . 00 ~/--+----+ ---0-----'---!-----:---.;-- ~
~
2.00 E '-'-'_"..1.U~"""Jii
0 . 00 1 .00 2 . 00 3.00 4.00 5 .00 6 .00 7.00 8 .00
position Zt [mm]
Figure 5. 2D POXSY //lap along z. Each pixel along Ihe vertical position correlales to a displacement
of Z = 2 . 7111111 over Ll = 1.3 illS. Th is means all averaged vertical velocity of 2. 1 /Ill s throughout all
vertical positions of the drop.

condition, as shown in Fig. 4b by the self-diffusion propagator superimposed to the


others. D is equal to the self-diffusion coefficient of water, and

" "rms
v diff =-;;- =-.J2Dil
il- . (1)

Figure 4a shows z propagator obtained with three different ranges of the field-of-flow
(fo!) . Here, n denotes the number of steps for the pulsed gradients, and

n-l 1 (2)
/0/=-'---
2 y·G·o · il

One of the propagator has been measured with a too small fof (Eq. 2) of 2.5 mis, not
covering the given velocity range and therefore the signal folds back into the velocity
window. In this case, because the approximate falling velocity is known beforehand, the
exact velocity could be obtained by adding the resulting averaged velocity of -0.5 mls
to the value of the faf= 2.5 m/s. Once the approximate vertical velocity is determined
by a pilot scan measurement, data acquisition can be focussed to a restricted portion of
the faJ, which allows the determination of the velocity distribution with much higher
resolution in a given time. This possibility to measure such high velocities with a too
small faf and nevertheless to be able to retrace the real velocity value could be very
helpful in overcoming some problems in the PGSE technique in the high velocity range.
One is the long measurement time for high velocities (Eq. 2) and another is the low
sensitivity.
333

As a means to directly correlate the spin densities at two successive times rather than to
obtain an averaged propagator, POsition eXchange SpectroscopY (POXSY) was
recently introduced by the authors [16,17]. It is realized by independent stepping of two
gradient pulses G I and G2 applied in the same spatial direction and separated by the
number k is defmed by k = yoG (Fig. 2e). Fourier transformation of S(kI,k2) results in a
2D correlation map between two position coordinates at successive times II = 0 and
t2 = f)., which is the joint two-time probability density W(ZI,0;Z2,f).) to find a particle at
position ZI at 1=0 and at position Z2 at I = f). [16].

W(ZI,0;Z2,f).) can be transformed into W(ZI,O;ZrZI,f).) and W(ZI,O;(zrzl)1 f).,f).) representing


the probability density to fmd a particle at position ZI at t=0 either carrying out a
displacement of Z = Z2-ZI during f). or possessing the averaged velocity v = (Z2-ZI)/f). over
f). respectively. Choosing the vertical axis for the gradients in the case of the falling
drop, a correlation image between the average vertical position during f). and the
corresponding falling velocity results (Fig. 5). In the experiment, gradient pulses G I and
G2 with a maximum
c) spin density image (zy) strength of 0.94 Tim and
a pulse duration 0 of
0.1 ms were chosen. The
two gradient pulses were
separated by f). = 1.3 ms
and incremented in
n = 65 steps each,
resulting in a
field-of-view (jov) of 8.0
b) mm and a field-of-flow
(joj) of 6.0 mmls, where

n-l 1
f o v = - · - - (4).
2 y·G · o
o The plot of Fig. 5
can be understood as
.4 1 follows: those spins in
the drop appearing at an
Figure 6. 2D spin-density images (a, c) and velocity encoded images (b, d)
arbitrary position at
along the cross-section of the drop. The spin density image along the xy t = 0, say ZI = 4 mm,
plane (a) and zy plane (c) prove the spherical shape of the drop. The z- correlate with the highest
velocity distribution along the xy plane shows an homogeneous
distribution of about 2 mls (b), whereas the x velocity distribution bears
marked structures along the zy plane. The net transverse velocity sums to
about zero, which excludes any net motion along the transverse direction.
334

probability density to the position Z2 = 6.7 mm at t = /1. Thus the spin had been
displaced by Z = ZrZI = 2.7 mm during /1. Because the distribution of the probability
density runs parallel to the main diagonal, the corresponding displacement amounts to
2.7 mm during /1 = 1.3 ms, irrespective of the initial positions, which is equivalent to an
averaged velocity of 2.1 mIs, for all parts of the drop. This velocity agrees with the
value resulting from the propagator measurements (Fig.4a). This consistency is
inherently given, because the sub-diagonal of the POXSY image corresponds to a
simultaneous stepping of the gradient pulse pair with same effective area and opposite
direction, which is nothing else than the conventional PGSE experiment (Fig. 2d) [12].
The same is true for the main diagonal of the POXSY image representing an averaged
position coordinate equivalent to that obtained by the phase encoding technique
discussed beforehand (Fig.2c) [12,17]. The projection of the image on either the
position axis at 11= 0 as well as at 12 = /1 reproduces the drop diameter of 3.5 mm as has
been determined independently by both 1D phase encoding and frequency encoding,
respectively. The falling drop turned out to be the ideal application for POXSY because
it represents a non-continuum flow bearing constant net velocity, where back folding in
flow direction (z) due to the limited sensitive volume of the resonator cannot take place
and the interpretation of displacement and its correlation to the initial position are well
defined.
The spatial distributions of velocities within the drop have been analysed by 2D
images along the cross-section of the drop by encoding the NMR phase $ of the image
with either vertical or transverse velocity v in the presence of a standard PGSE
encoding pair with amplitude G, duration 0 and time separation /1, in addition to the
frequency encoding gradients (Figs. 3a and 3b), where

(5)

Figure 6a presents a 2D spin-density image


along the xy plane and Fig. 6b the
corresponding Z velocity encoded image. In
Fig. 6a the liquid wetting the inner surface of
the glass tube is visible. The drop is seen to fall
almost in the center of the tube. The fact that
two orthogonal cross-sectional images, the 2D
xy spin-density image (Fig. 6a) and the 2D zy
spin-density image (Fig. 6c), are circular
proves the spherical symmetry of the drop. The
z velocity distribution appears to be
homogeneous along the xy cross-section and
the resulting value of 2.1 mls agrees with the
result obtained by the POXSY experiment. The Figure 7. 3D spin-density image of a falling
x velocity encoded zy image in Fig. 6d reveals drop consisting of 2.8 % surfactant so/ulion.
a marked structure of deterministic nature The experiment takes about 10 hours, and
across the plane. The fact that the net 33800 drops constitutes the NMR signal,
which gives the 3D image after Fourier
transformation.
335

transverse velocity amounts to zero excludes the contribution of a net motion in


transverse direction of the drop itself. The magnitude of the motion of about ±15 mmls
and its spatial distribution supports the presence of measurable drop dynamics. This
could be either attributed to coherent fluctuations or oscillations of the drop itself, or to
the internal dynamics including circulations inside the drop, whereby the small
amplitude of the velocity variation excludes the presence of large scale vortices.

3. Conclusion

In conclusion we were able to measure position and velocity of and inside a free falling
water drop in all three spatial directions in a direct way and furthermore obtain
correlated information of position and velocity. This information tells us about the net
dynamic behavior of the falling and interior motion. The dripping process was shown to
be reproducible with respect to both frequency and shape of the drops. Therefore the
accumulation of information by a triggered cycle for one acquisition per drop was
feasible. For the fIrst time NMR imaging techniques -the only technique being able to
acquire in-situ dynamic information in a non-invasive and direct way- were applied to a
free falling drop and proved that reliable and reproducible information could be
extracted. Although the spherical drop-shape could be proved by two orthogonal 2D
spin-density images, the three-dimensional rendering of the falling drop by a 3D spin-
density imaging method is presented in Fig. 7. This visualizes the above mentioned
strength of the NMR imaging technique even more clearly. Considering that one 3D
experiment takes about 10 hours and a total of 33800 drops constitutes the NMR signal,
the quality of this 3D image obtained after Fourier transformation is excellent. This
work gives the basis for more comprehensive NMR studies with the ability to
reconstruct the 3D structure of the inner vortex-dynamics inside a drop and its outer
shape [18].

4. Acknowledgement

This work was funded by the Deutsche Forschungsgemeinschaft as part of the SFB 540.
Dr. Peter B1uemler prepared and wrote the proposal for the SFB 540 project and gave
the basic impetus for this work. We are grateful to Michael Adams, Stefan Aye and
Gunther Schroeder from the electronic workshop for setting up the photo-sensor and
trigger electronics.

5. References

[I) Cohen I., Brenner M. P., Eggers J. and Nagel S. R., (1990) Phys. Rev. Lett. 83, 1147.
[2] Brenner M. P., Eggers J., Joseph K., Nagel S. R., Shi X. D., (1997) Phys. Fluid 9, 1573.
(3) Donelly R. 1., Glaberson W., (1966) Proc. R. Soc. London Ser. A290,547.
(4) SFB540, 'Modellgestiitzte experimentelle Analyse kinetischer Phiinomene in mehrphasigenfluiden
Reaktionssytemen, 01.07.1999-30.06.2002.
(5) Hauser E. A., Edgerton H. E., Tucker W. B., (1936) J. Phys.Chem. 40,973.
(6) Edgerton H. E., Hauser E. A., Tucker W. B., (1941)J. Phys.Chem. 41, \017.
336

[7] Hunter J. C., Collins M. W., (1990) Int. J. Optoelectronics 5, 405.


[8] Dudderar T. D., Meynart R., Simpkins P. G., (1988) Optics Lasers Engng. 9, 211.
[9] Keane R. D., Adrian R. J., (1990) Meas. Sci. Technol. 1, 1202.
[10] Karger J., Pfeifer H., Heink W., (1988) Adv. Magn. Reson. 12,1.
[II] Callaghan P. T., (1991) Principles ofNMR Microscopy' Oxford: Clarendon Press.
[12] Bliimich B., (2000) NMR Imaging for Materials, Oxford Science Publications.
[13] lauterbur P.e., (1973) Nature 242,190.
[14] Mansfield P., Grannell P. K., (1975) Phys. Rev. B 12, 3618.
[15] Edelstein W. A., Hutchinson J. M. S., Johnson G., Redpath T., (1980) Phys. Med. BioI. 25,751.
[16] Han S., StapfS., BIOmich B., (2000) J. Magn. Reson. 146,169.
[17] Han S., BIOmich B., (2000) Appl. Magn. Reson 18, 101.
[18] Han S., StapfS., BIOmich B., (2001) Phys. Rev. Lett.' to be published.
Robust Characterisation of Flowing Emulsions using Regularisation
and Velocity-compensating Pulsed Field Gradient (PFG) Techniques

M.L. Johns, K. Hollingsworth, G.M. Davies and L.F. Gladden


Department of Chemical Engineering
University of Cambridge, u.K.

1. Abstract

The use of nuclear magnetic resonance (NMR) pulsed field gradient (PFG)
techniques to produce emulsion droplet-size distributions, is well established.
This is accomplished by employing a description of restricted self-diffusion in
spherical cavities to interpret the PFG data. We present two improvements of
this droplet sizing technique in order to increase its robustness. The droplet size
distribution is generally assumed to be a log-normal shape, which both stabilises
and simplifies the mathematical extraction of the droplet-size distribution from
the PFG data. An alternative technique is reported here, which employs
Tikonov regularisation theory to perform this extraction. Consequently no
assumptions are made with respect to the shape of the droplet-size distribution.
The technique is also developed to be able to characterise an emulsion under
flowing conditions. Conventional PFG is unable to distinguish between self-
diffusion and a distribution of velocities. We introduce a flow-compensated
PFG technique to enable self-diffusion alone to be detected, hence enabling
measurements of droplet-size distributions for flowing emulsions.

2. Introduction

Emulsions are frequently used in a wide range of industrial products including


detergents, cosmetics, pharmaceuticals and a wide variety of food products (e.g.
milk and butter). Essential to their functionality as industrial products are their
stability, texture, colour and viscosity. All of these emulsion characteristics are
strongly dependent upon the droplet-size distribution of the dispersed phase.
Hence an accurate and non-invasive description of the droplet-size distribution is
highly desirable, particularly if such measurements are required in actual
processing equipment or on the final packaged product. Such a non-invasive
description is available via the ability of PFG to measure the self-diffusion of
337
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 337-346.
© 2002 Kluwer Academic Publishers.
338

molecules. When these molecules are restrained to the interior of the emulsion
droplets, their resultant restricted self-diffusion behaviour can be interpreted so
as to determine the emulsion droplet-size distribution. Emulsion droplet sizing
using PFG techniques, was first employed by Packer and Rees [1] using the
description of signal attenuation in spherical cavities developed by Murday and
Cotts [2].

A variety of alternative techniques are available to measure droplet-size


distributions in emulsions. These include optical microscopy, electron
microscopy [3] and the use of a coulter counter. These techniques are however
invasive and not suitable for application to industrial processing equipment or
for packaged emulsions. Other alternative techniques, which can potentially be
applied in a non-invasive manner, include ultrasonic measurements [4], turbidity
meters, neutron scattering and light scattering [5]. However none of these
techniques provide the versatility of PFG, which is able to differentiate between
different chemical species, be used to characterise complex emulsion structures
(e.g. multiple emulsions [6]), be applied to emulsions containing air bubbles and
deal with both dilute and concentrated emulsions [7].

Sizing emulsion droplets using PFG however has a number of limitations, which
compromise its robustness. These limitations are:
(i) It assumes a droplet-size distribution shape (usually a log-nonnal) to
enable the mathematical extraction of the distribution from the raw PFG data.
(ii) It is unable to differentiate between a distribution of velocities and self-
diffusion and hence cannot be used to size flowing emulsion droplets. In
addition, for the same reason, convection currents can distort emulsion droplet-
size measurements in apparently stationary emulsions [8].
We address (i) by employing Tikonov regularisation [9], in which no shape is
assumed for the droplet-size distribution, to perfonn the mathematical inversion
of the PFG data to produce the droplet-size distribution. This is applied to both
theoretical droplet-size distributions as well as to experimental PFG data from
oil-in-water emulsions. With respect to (ii), we employ flow-compensating
PFG pulse sequences, which eliminate the effects of coherent motion (i.e. flow)
whilst measuring incoherent motion (i.e. self-diffusion). The stability of this
flow-compensating technique, as superficial velocity is increased, is
investigated.

3. Theory

3.1 EMULSION DROPLET-SIZING

When using PFG techniques to measure self-diffusion, the signal intensity, S,


detected can be related to the acquisition parameters using the following well-
known Stejskal-Tanner equation [10]:
339

~ = exp(- DO(ygO)2(~ - 013)) (1)


So

where g is the strength of the applied magnetic field gradient, So is the signal
detected when g is 0, Do is the self-diffusion coefficient, y is the gyro-magnetic
ratio, 0 is the time period over which g is applied and ~ is the time period over
which self-diffusion is monitored. However for the case of self-diffusion
restricted by a barrier, Eq. 1 is no longer applicable. For the case of self-
diffusion in spherical cavities, the following equation has been developed to
describe S [2]:

H(g,R)=~=_2y2g2 I: 2a~ X[20-+-] (2a)


So Do m=lamR -1 amDo

where
G = 2+ exp(-a~Do(~ -0))- 2exp(a~Doo)- 2exp(a~Do~)
(2b)
+ exp(-a~Do(~ + 0))

and am is the mth root of the following equation:

_1-J 3 / 2(aR) = J 5/2 (aR) . (2c)


aR
The radius of the spherical cavity, R, is the only free parameter in Eq. 2, hence
by fitting Eq. 2 to experimental PFG data for the dispersed phase of an emulsion,
it is possible to determine R. Most emulsions however contain a distribution of
droplet sizes and hence the following equation has to be used in place ofEq. 2:
~= I H(g,R).F(R)8R , (3)
So
where F(R) is the volumetric probability distribution of droplets with a radius, R.
Eq. 3 is a classic Fredholm Equation of the first kind. Extracting F(R) from Eq.
3 is notoriously difficult and is considered an ill-conditioned unstable problem
[11]. The assumption that F(R) is a log-normal shape simplifies this extraction
process considerably. Ambrosone et al. [12] relaxed this assumption but still
used generating functions to describe F(R) when inverting Eq. 3. In employing
Tikonov regularisation we make no assumptions with respect to the shape of
F(R), as described in the next section.

3.2 REGULARIZATION

Eq. 3 can be expressed in matrix notation as:


S=H.F (4)
340

where S is the vector of experimental data points as a function of increasing g, H


is a 2D matrix representing Eq. 2 as a function of g and R respectively and F is
a vector representing the droplet size distribution. A matrix inversion of Eq. 4 is
generally not possible as H is ill-conditioned and S contains a small amount of
noise. Hence F is determined which minimises the following function, E:
n m m T
E =~«Si - ~ Hi,j.F)(Si - L Hi,j.F) ) (5)
1=1 J=I J=I

where n is the number of experimental data points and m is the number of


discrete points defining the droplet size distribution. However in Eq. 5 small
errors in S will manifest as large fluctuations in F. Consequently the shape ofF
is dominated by numerous large spikes, which are physically unrealistic.
Tikonov regularisation deals with this problem by adding an additional term to
E, which penalises the 'spikiness' ofF in the following manner:
n m m T
E1 = ~«Si - ~ Hi,j.F)(Si - ~ Hi,j·Fj) ) + AD (6)
1=1 J=I J=I

where D is a function characterising the smoothness of F and A is a variable


determining the degree of smoothness of F. D can represent the area, 1st
derivative or 2nd derivative of F. In our implementation, we have selected the
2nd derivative ofF. Following the method of Wilson [13], this is achieved in the
following manner:
n m m T T , ,T
E
1
=;L«Si - ~ Hi,j"F)(Si - L Hi,j.F j ) ) + A(F D D F), (7)
1=1 J=I J=I

where D' is a 2D matrix representing the i nd derivative. F is hence selected


which minimises the function shown in Eq. 7. Selection of A proceeds by
repeating the minimisation of Eq. 7 for various values of A. Two options are
available to select A at this point, the L-curve method and General Cross-
Validation (GCV), both of which are described in Wilson [13]. The L-curve
method plots E as a function of an increasing A; a distinct inflection occurs, at
which point E increases dramatically. This point of inflection is chosen to select
A and hence F. For the GCV method a score is calculated for each value of
A which quantifies the ability of F to calculate the experimental data using only
m-l points to define F. The value of A, corresponding to a minimum in this
score, is selected.

3.3 FLOW-COMPENSATION

A conventional PFG pulse sequence is shown in figure lea). When a


distribution of velocities are present, the effect is to attenuate the NMR signal, S,
341

in a similar way to self-diffusion. Hence the effect of self-diffusion cannot be


isolated and flowing emulsion droplets cannot be sized using the conventional
PFG technique. An alternative pulse sequence to detect self-diffusion is shown
in figure 1(b). This is a stimulated-echo variant of that presented in Callaghan
and Xia [14]. Provided displacement during the first!::J2 time period is equal to
displacement during the second !::J2 time period, no attenuation of S will occur.
This is clearly the case for flow in a tube provided that self-diffusion across
streamlines is negligible. Self-diffusion is incoherent and hence displacement
during the first !::J2 time period is independent of displacement during the second
!::J2 time period. Thus flowing emulsion droplets can be sized using the pulse
sequence shown in figure I(b).
(a)
Echo
90° 90° 90° Acquisition
rJ.
A
pulses
nl I~! S ."
'U V

~ ~
gradient g g
pulses


Spoil Time
~
.i
:. 4

(b)
90° 90° 90° 90° 900 E~h?
AcqwsltIOn
r.f.
..
A
pulses
~ I~ 1S

gradient ~g Ig ~g ,g 'V V

..
pulses


Spoil Spoil Time
~ ~
:. AI2 : , AI2 ~

Figure I: (a) The conventional PFG pulse sequence. (b) The flow-compensating
PFG pulse sequence, which can be used to size flowing emulsions.
342

4. Experimental

The oil-in-water emulsion investigated consisted of xylene(or toluene)-in-water


with the following composition: 37 wt % xylene (or toluene), 6 wt % tween 40
and 57 wt % water. Tween 40 is a non-ionic surfactant in which the poly oxy-
ethylene chains contain an average of 40 units. The emulsion was prepared by
dissolving the surfactant in the water, with oil then being emulsified in the
resultant solution using a magnetic stirrer. The emulsion was then transferred to
a gently agitated storage vessel from where it was pumped through a 10 mm
inner diameter glass tube of length 2m; located in the r.f. coil of the NMR
equipment with signal detected from a 10 mm long section, 1m from the liquid
entrance.

The NMR equipment used was a 7 T vertical bore magnet equipped with a 15
mm diameter birdcage r.f. coil, tuned to 300.13 MHz to detect protons.
Chemical shift differences were used to selectively detect the xylene (or toluene)
signal. Magnetic field gradients with a maximum strength of 97 G/cm were
available to detect self-diffusion in three orthogonal directions. The self-
diffusion of oil molecules in the non-flowing emulsion droplets was dete~ted
using both the PFG pulse sequence (figure l(a)) and the flow-compensating PFG
pulse sequence (figure I (b)). The flow compensating pulse sequence was then
used to measure self-diffusion when the emulsions were flowing at superficial
velocities of 0.01 ms· l , 0.05 ms· 1 and 0.1 ms· 1 respectively. For verification
purposes, a sample of each emulsion was viewed using a Zeiss LSM510
confocal microscope, and a droplet-size distribution extracted.

5. Results and Discussion

5.1 VERIFICATION OF REGULARISATION TECHNIQUES

To test the regularisation techniques, a log-nonnal and a dual log-nonnal


distribution of droplet sizes, F, were used to produce signal attenuation data sets,
H(g,R), using equation 2, to which noise was added. The original droplet-size
distributions are shown in figures 2(a) and 2(b). The regularisation technique
was then applied to H(g,R) using the L-curve method, and the droplet size
distributions recreated. The required minimisation of Eq. 7 was achieved using
quadratic programming [IS], with non-negativity constraints applied to the
elements of F, the droplet-size distribution. The droplet-size distributions
predicted by regularisation appear in figures 2(a) and 2(b); agreement with the
original functions is very good.
343

I
j 0.02 17\~"--;::::::~:::::::~:::::::==~ 0.025 I-.----;::::::::::::=:::::===~

0.0 IS
-Original Function
••••• Regularisation Predlctio 1 0.02
-Original Function
.•••• Regularisation Prediction

.~
1c: 0.Ql5
]
·e 0.01
.'o"
::I

is 'E.. om
is
.8§
Z
0.005
f
Z
0.005

oL---~--~~~~~~~
o 5 10 IS 20 25 10 IS 20 25
Droplet Radius (Microns) Droplet Radius (Microns)

Figure 2: (a) Prediction of the regularisation method of the droplet-size distribution for a log-
normal distribution. (b) Prediction of the regularisation method of a dual log-normal distribution.
Agreement between the predictions and both original functions is very good.

5.2 VERIFICATION OF PFG AND FLOW -COMPENSATED PFG

In figure 3, the droplet-size distributions are shown for the non-flowing xylene-
in-water emulsion, as produced by confocal microscopy, the conventional PFG
technique (figure I (a» and flow-compensating PFG (figure I (b» .
Regularisation was used in producing the PFG droplet size distributions.
Agreement between all three methods is excellent. This serves as verification
that the PFG method is accurately predicting the droplet-size distribution and
that the flow-compensated PFG method is not introducing any additional errors.

0.025 r----r-.....,-,..--r---r-.....,-,..--..

ii 0.02
--Conventional PFG
••••• Flow-compensating PFG
EI Confocal Microscopy

!

:;
0.015

.D
.~ 0.01
is
]
§ 0.005
Z

o~~uoooom~~~~~~
o 5 10 15 20 25 30 35 40
Droplet Diameter (Microns)

Figure 3: Prediction ofaxylene-in-water emulsion droplet-size distribution by conventional PFG,


flow-compensating PFG and confocal microscopy. Agreement between the three methods is
excellent.
344

5.3 EFFECT OF AN INCREASING FLOW

The droplet-size distributions produced using the flow-compensated PFG


technique are shown in figure 4 for the xylene-in-water emulsion, as a function
of superficial velocity. Note that the data produced, when using the
conventional PFG technique applied to a flowing emulsion, could not be used to
determine a sensible droplet-size distribution. With respect to figure 4, as the
superficial velocity is increased from 0 ms'· to 0.05 ms'·, the droplet-size
distribution is seen to narrow with a distinct reduction in size of the larger
droplets due to increasing velocity shear. This is encouraging as if velocity was
distorting the measurement of self-diffusion, the droplet-size distributions would
be expected to increase with an increasing superficial velocity; clearly this is not
the case.

0.025

]'
:;j 0.02
--Oms'!
E
~r:: 0.015 - - - - - 0.0 1 ms'!
.9 ··· .... ··0 .05 ms'!
] . -0.1 ms'!
.~ 0.01
is
]
§ 0.005
Z

2 4 6 8 10 12 14
Droplet Radius (Microns)

Figure 4: Droplet-size distributions for a xylene-in-water emulsion as a function of an increasing


superficial velocity. The effect of an increasing velocity shear in reducing the size of the larger
droplets is clearly evident.

However as the superficial velocity is increased from 0.05 ms'· to 0.1 ms'·, the
droplet-size distribution is unaffected. This could potentially be due to a small
distortion of the self-diffusion measurements at the higher superficial velocity,
however we believe that the increasing velocity shear is no longer able to cause
further break-up of the droplets at a superficial velocity of 0.1 ms'·. In this
respect it is worth noting that the toluene-in-water emulsion produced droplet-
size distributions, which were independent of superficial velocity over the same
range. This suggests that droplet break-up in response to an increasing velocity
345

shear has a threshold beyond which further break-up is not possible, and that the
magnitude of that threshold is a complex function of composition. The
robustness of the flow-compensated PFG technique at higher superficial
velocities is the subject of on-going work.

6. Conclusions

Two improvements to the PPG technique of sizing emulsion droplets are


presented with the objective of rendering the technique more robust. The
extraction of the droplet-size distribution from the raw PFG data has been made
more general in that a shape for the droplet-size distribution no longer needs to
be assumed. The second improvement is the use of a flow-co~pensated PFG
pulse sequence, which enables emulsions to be sized under flowing conditions.
The motivation for these improvements is the development of PFG emulsion
droplet sizing for application to industrial processing equipment. In this
capacity it will be used as both a research tool to improve our understanding of
in-situ emulsion formation or as an on-line control device.

7. Nomenclature

Do Self-diffusion coefficient
D,D' 2nd derivative matrices
E,E' Error functions
F(R)(F) Probability Distribution of Droplet Radii (Volumetric)(Vector of F(R) as a
function of R)
g Strength of the applied magnetic field gradient
H(g,R)(H) Function describing S in a spherical cavity of radius, R (Eq. 2) (Vector of
H(g,R) as a function of g and R)
In nih order Bessel function
m Number of points defining F
n Number of experimental data points
PFG Pulsed field gradient
R Radius of droplets
S(8) Signal Intensity (Vector of S as a function of g)

o Time period over which g is applied


fl Time period over which self-diffusion is monitored
y Gyro-magnetic ratio
A Regularisation parameter
346

8. References

1. Packer, K.l and Rees, C. (1972) Pulsed NMR studies of restricted diffusion. 1. Droplet size
distributions in emulsions, J. Colloid Interface Sci., 40, 206.
2. Murday, 1.S. and Cotts, R.M. (1968) Self-diffusion coefficient of liquid lithium, J. Chem.
Phys. 48,4938.
3. Donald, A.M., Chaobin, H., Royall, P., Sferrazza, M., Stelmashenko, N.A. and Thiel, B.L.
(2000) Applications of environmental scanning electron microscopy to colloidal aggregation
and film formation, Colloids and Surfaces A, 174,37.
4. McClements, DJ. and Coupland, IN. (1996) Theory of droplet size distribution
measurements in emulsions using ultrasonic spectroscopy, Colloids and Surfaces A, 117,
161.
5. Min, K.Y. and Goldburg, W.1. (1993) Dependence of nucleation on shear-induced initial
conditions, Phys. Rev. Lett., 71, 569.
6. Lonnqvist, I., Hakansson, B., Balinov, B. and Soderman, O. (1997) NMR self-diffusion
studies of the water and the oil components in a W/orw emulsion, J. Colloid Interface Sci.,
192,66.
7. Balinov, B. Linse, P. and Soderman, O. (1996) Diffusion of the dispersed phase in a highly
concentrated emulsion: Emulsion structure and film permeation, J. Colloid Interface Sci.,
182,539.
8. Manz, B., Seymour, lD. and Callaghan, P.T. (1997) PGSE NMR measurements of
convection in a capillary, J. Magn. Reson., 125, 153.
9. Merz, P.H. (1980)J. Comput. Phys., 38, 64.
10. Stejskal, E.O. and Tanner, lE. (1965) Spin diffusion measurements: spin echoes in the
presence of a time-dependant field gradient, J. Chem. Phys., 42, 288.
II. Press, W.H., Teukolsky, S.A., Vetterling, W.T. and Flannery, B.P. (1992) Numerical Recipes
in Fortran, Cambridge University Press, Cambridge.
12. Ambrosone, L., Ceglie, A., Colafemmina, G. and Palazzo, G. (1999) General methods for
determining the droplet size distribution in emulsion systems, J. Chem. Phys., 110, 797.
13. Wilson, lD. (1992) Statistical approach to the solution of 1'st kind integral-equations arising
in the study of materials and their properties, J. Mater. Sci., 27, 3911.
14. Callaghan, P.T. and Xia, Y. (1991) Velocity and diffusion imaging in dynamic NMR
spectroscopy, J. Magn. Reson., 91, 326.
15. GAMS Development Corporation, 1217 Potomac Street, NW, Washington, DC, 20007
(1996).
PHOSPHOLIPIDS' SERA AND MONONUCLEAR CELLS
IN ACUTE LEUKEMIA, MALIGNANT LYMPHOMA
AND MULTIPLE MYELOMA-EVALUATION BY 3Ip MRS IN VITRO

*M. KULISZKIEWICZ-JANUS, **B. BACZYNSKI


*Department of Hematology, Wroclaw University of Medicine,
Wyb. Pasteura 4, 50-367 Wroclaw, Poland,
e-mail: mkj@hemat.am.wroc.pl
** Department of Chemistry, University of Wroclaw,
ul. Joliot-Curie 14, 50-383 Wroclaw, Poland
e-mail: stb@wchuwr.chem.uni.wroc.pl

1. Introduction

Phospholipids from tissue extracts have been used to differentiation between malig-
nant and normal tissue and to identify tissues undergoing malignant transformations
[1-10]. The investigations of31p spectra of sera of patients with hematological malig-
nancies (acute leukemia, malignant lymphoma, multiple myeloma) and other can-
cers were clinical trials over the introduction of MRS (magnetic resonance spec-
troscopy) to monitor the therapy [11-14]. Moreover, changes of concentrations of
phospholipids from extracts of sera and of mononuclear cells may explain the mech-
anism of their transport through cell membranes. Sphingomyelin (SM), as well as
phosphatidylinositol (PI) and phosphatidylcholin (PC), is essential element of cellu-
lar signal transduction. Products of its degradation, i.e. ceramide and sphingosine
are spontaneously transported through membranes and function as second messen-
ger directed information to the cells (I 5-18]. Study aimed in comparison of 31 P spec-
tra acquired from extracts of phospholipid blast cells of patients with acute leukemia
with these gained from lymphocyte extracts of healthy persons, with remarkable
attention paid to sphingomyelin behavior.

2. Material and methods

Studies were carried out on AMX 300 Bruker spectrometer 7.05 T. 31p spectra from
sera, methanol-chloroform extracts of sera, and extracts of cells from patients with
hematological cancers and healthy volunteers were performed. Blood samples were
collected by venous puncture after an overnight fast.
We acquired all together 645 spectra from sera. The details of methods used in inves-
tigation are given elsewhere [11,12,14].
347
J. Fraissard and O. Lapina (eds.J, Magnetic Resonance in Colloid and Interface Science, 347-354.
© 2002 Kluwer Academic Publishers.
348

In order to attain extracts the serum was obtained by centrifugation of the sample
(3000 rpm for 20 min), next 4 ml of sample was mixed with the methanol-chloroform
mixture (1:1:1, v/v/v) at 1000 rpm for 30 min. at room temperature. After that the mix-
ture was centrifuged at 3700 rpm for 16 min. The lower organic phase (liquid) was
separated and evaporated to dryness in a rotary evaporator. The dried lipids were
dissolved in 0.4 ml of chloroform and 0.2 ml of 200 mM aqueans EDTA Na 2 (pH
6.0) diluted by methanol in ratio I :4. Extractions of cellular lipids were performed
from 60.0 min mononuclear cells obtained by Ficoll buffy coat centrifugation. All
together 113 spectra from extracts of sera and 101 from extracts of mononuclear
cells were acquired. The method used to obtained the 31p spectra were described pre-
viously [14].

3. Results

In preliminary studies we observed that 31p MRS spectra of normal serum consist of
three peaks including a downfield peak due to Pi (inorganic phosphate) and two
additional upfield field peaks from phospholipids PE + SM (phosphatidyl-
ethanolamine + sphingomyelin) and PC (Figure IA). 31p spectra were performed in
healthy volunteers, patients with malignant lymphomas, multiple myeloma and
acute leukemia at the time of diagnosis (Figure IB, C, D) and repeated up to 13
times during chemotherapy [11-13].

A 8

PC

ppm 3 1 D ., ., .$ D -, -I -J

c D

ppm $ 1 D -, -I -$ ppm 3 1 D ., -I -3

Figure 1. 31 P spectra of sera of healthy volunteer (A), and of patients with malignant lymphomas (B),
multiple myeloma (C), acute leukemia (D). All acquired at the time of diagnosis.
349
PC

PE+SM

... _.,..~
8....____A... __
.~ ..... --
...
i i
fA
i
,.1 ,
i

"".,.
i
.f.Z
i
.#.4 ....
i

Figure 2. 31p spectra of sera after addition of sodium salt of cholic acid
A-<>fhealthy volunteer; B-<>fpatient with acute leukemia (14]

The sodium salt of cholic acid added to serum caused separation of three phospho-
lipid peaks located upfieJd from inorganic phosphate. Contrary to the earlier stud-
ies, peaks from PE+SM and PC, and also a peak from LPC (lysophosphatidyl-
choline) were observed (Figure 2A, B). The above mentioned separation method had
also been applied in investigations of patients with digestive tract tumors and with
renal cell carcinoma. Changes in phospholipids in the 31p MRS spectra observed in
these patients were primarily dependent on the advance of the disease[14].

A 51
-IntPC --_. In' lIE + SII

Int[%l

211
____________________ ~......
..........-
~I •

...,,,,'"
-----~-----------------------------------------------
111 111
1111 II a a
Ii
a ~
== = !!.......t --=+I
WHk.

-M-M-M -..-..
8 C

- , ,. 0'...., .,
"

Figure 3. Patient responded to therapy


- I
, ., .z .,

A-the diagram of percentages of PE+SM and PC peak intensity; B_31p spectrum before chemotherapy;
C_31P spectrum in complete remission [13]
350

Long-term follow-up studies showed a good correlation between this 31p MRS eval-
uation of sera and the response of the disease to the therapy. At the time of diagno-
sis spectra showed strongly reduced peak areas and intensities from phospholipids
(PC, LPC and PE + SM).
During chemotherapy important changes in spectra were observed [11-14]. In
responding patients the spectral profile changed to resemble that of normal serum
with increased peak intensities (Figure 3). In non-responding individuals peak inten-
sities were reduced (Figure 4).
III
A --,nfPC ---. Int , . + SII

Int l"J "


t:.ItrII
11
. ------------~---
-. • ----------= 11

~ = ~ Ii
CHEMOTHERAPY' .. .1 "
8 c

Figure 4. Patient non-responded to therapy


A-the diagram of percentages ofPE+SM and PC peak intensity; B_31p spectrum before chemotherapy;
C_31P spectrum in progress of disease [13]

Spectra of patients suffering from acute leukemia or Hodgkin's disease, who have
achieved complete remission for 4- I 2 years did not differ from spectra of healthy
volunteers (Figure 5A, B).

A B

Figure 5. 31p spectra of sera of patients with acute leukemia (A) and Hodgkin's disease (B), who achived
complete remission 12 and 9 years ago, respectively
351

Int[%] 50 r--------------------------------,
4~--------------------------------------------~~~
-·~
-e-lntPC ..... ""'lntPE+SM /'
40 +-----_ _ _ _ _------+-l----j

35 }--------------------------------------------------I-I-----1
...............-.....................:?.~~~.~.-......................................__........../ ................
30 ~~-----------------------------------------------i~------~

r\r--""
25 "' / #.:! ......._

20 \ ~. ##

15 .......... ~ ~ I'::'
, ...........~..................................................~....................................j.......................
10 ji) ~ .... - \. Control PE + SM ...........~ •.1
5 I~ :.1...... "...l............................:.3... ''.1....

, ~ III H'I
I--='-J III~L~::~ -
Weeks
11-/1--11-

- ,,-~frl
----tot Comp .... R.mI ... on

Figure 6. PE + SM and PC intensities during 400 weeks following the diagnosis of acute leukemia
At the time of complete remission PE + SM intensity was below control value, PC intensity was far
below normal, too. Not before after autologic bone marrow transplantation both parameters achieved
normal values and even exceeded

A C
••• ~~ ...................................................................~!!:!:!.'!£

•• • ••
!lop l1li.-
J • • " II II
•• • • • 11
• II II
• •
II-
S II-
D
,.........._._.......
tt-
_-_ .............
11-
................................................ _..........................................
Control file + .",

1.-
12-
............
....,

.-
I-

.-
..
2- 2-

• ., •
l1li.- l1li.-
I
• J • • t• II II
•• • • •
I
• II II

Figure 7. The diagrams of percentages of PC (part A, C) and PE + SM (part B, D) peak intensities
A & B-patients responding to the first course of chemotherapy; C & D-patients non-responding
352

31p spectra can prove the presence of the residual leukemia in patients when the
number ofleukemic cells equals to 109 and is not detectable in laboratory tests of the
blood and bone marrow. 31p spectra can depict the function of the transplanted
(Figure 6).
It is possible to estimate the efficiency of cytostatics: e.g. DHAD & ARA-C (mitox-
antrone and cytarabine) as a first course treatment in patients with acute leukemia
(Figure 7A, B, C, D).

PC

A 8
PC

... SII

LPC CPUS

., .
LI'C CPUS

-
ItS

u U I.' U U

- ... U U I., U ... I.' ... ... I.f U U I.,

Figure 8. 31p spectra from extracts of sera of healthy volunteer (A), and of patients
with acute leukemia (8)
... , .." ...1 ....

31 P MRS spectra of normal extract of serum consist of seven peaks due to phospho-
lipids. Beside previously identified, i.e. PC, LPC, PE and SM, some new ones were
observed: PI (phosphatidylinositol), PS (phosphatidylserin), CPLAS (phosphatidyl-
choline-plasmalogen; ~-acetyl-'Y-O-alkyl L-a-phosphatidylcholine) (Figure SA). In
patients with hematological cancers the values of peaks areas of PC, CPLAS, LPC,
SM decreased. In some patients peaks from PS, PI, PE were not observed (Figure
SB). In responding patients the spectral profile changed to resemble that of normal
extract.
31p MRS spectra of extract of mononuclear cells consist of peaks due to phospho-
lipids: PC, CPLAS, LPC, SM, PE+PI, PS. The peak ofLPC that has the most prog-
nostic value in sera, in extract of mononuclear cells was observed only in healthy vol-
unteers. The peak of PS is also characteristic for healthy volunteers (Figure 9A).
Important meaning for prognosis of disease course has got a presence of CPLAS .
The values, considerably lower than those observed in healthy people have decided-
ly wrong prognostic value (Figure 9B). However, patients with values considerably
exceeding those existing in healthy volunteers quickly achieved remission of disease.
These patients also have higher SM values than those non-responding to treatment,
but lower than healthy volunteers.
353

PC
PC
A 8

H,.,
... ...,.,
CPLAS

- -
Figure 9. 31p spectra from extracts ofmononuc1ear cells of healthy volunteer (A), and of patients
with acute leukemia (B)

4. Conclusion

The experiences already achieved pointed out that 31p spectra at present didn't allow
for diagnosis of hematological disorders, but were of great importance in monitor-
ing of therapy of the diseases under consideration [11-14]. Moreover, the changes of
concentrations of phospholipids in sera and cells in hematological cancers are prob-
ably due to the increased uptake of phospholipid metabolites in proliferating blast
cells, and their disturbed transport through cell membranes. Mechanism of the
changes of concentrations of the phospholipids in sera and cells in acute leukemia
and disturbed transport of phospholipids through cell membranes remains
unknown. It was suggested, that the pool involved in the SM cycle of signal trans-
duction is localized to the intracellular portion of the human leukemia cells, most
likely to the inner leaflet of the plasma membrane [15-18]. It is very likely that
observed reduction of the level of SM in blast cells' extraction is due to activation of
sphingomyelinase (SMnase) by tumor necrosis factor (TNF). SMnase cleaves SM,
generating choline phosphate and ceramide, that-among other things-mediate pro-
grammed cell death (apoptosis), induce differentiation and inhibit the growth of
leukemia cells. As SM is an important participant of signal transduction pathway,
the reason of reduced level of SM in the extracts of leukemia cells and changes con-
cerning the CPLAS area observed by us needs further investigation.

5. References

I. Bental, M. and Deutsch, C. (1993) Metabolic changes in activated T cells: an NMR study of human
peripheral blood lymphocytes, Magn. Reson . Med. 29,317-326.
2. Gillies, R.J. (1994) NMR in physiology and biomedicine, Academic Press, San Diego.
3. Leray, G. and De Certaines, J.D. (1994) Proton NMR spectroscopy of plasma lipoproteins: a marker
of the immune function in cancer disease? Anticancer Research 14, 1839-1852.
4. Merchant, T.E., Kasimos, J.N., de Graaf, P.W., Minsky, B.D., Gierke, L.W. and Glonek, T. (1991)
354

Phospholipid profiles of human colon cancer using 31p magnetic resonance spectroscopy, Int. J.
Colorect. Dis. 6, 121-126.
5. Merchant, T .E., Meneses, P., Gierke, L.W., Den Otter, W. and Glonek, T. (1991) Magnetic resonance
phospholipid profiles of neoplastic human breast tissues, Br. J. Cancer 63,693-698.
6. Merchant, T.E., Minsky, B.D., Lauwers, G .Y., Diamantis, P .M ., Haida, T. and Glonek, T. (1999)
Esophalleal Cancer Phospholipids Correlated with Histopathologic Findings: a 31 P NMR Study,
NMR Biomed. 12, 184-188.
7. Negendank, W. (1992) Studies of Human Tumors by MRS: a Review, NMR Biomed. 5, 303-324.
8. Podo, F. (1999) Tumor phospholipid metabolism, NMR Biomed. 12,413--439.
9. Ruiz-Cabello, J. and Cohen, S. (1992) Phospholipid metabolites as indicators of cancer cell function,
NMR Biomed. 5,226-233.
10. Szwergold, B.S., Kappler, F. , Boldes, M., Shaller, C. and Brown, T.R . (1994) Characterization of a
phosphonium analog of choline as a probe in 31p NMR studies of phospholipid metabolism, NMR
Biomed. 7,121-127.
11. Kuliszkiewicz-Janus, M. and Baczynski, S. (1995) Chemotherapy-associated changes in 31p MRS
spectra of sera from patients with multiple myeloma, NMR Biomed. 8, 127-132.
12. Kuliszkiewicz-Janus, M. and Baczynski, S. (1996) Application ofP NMR spectroscopy to monitor of
chemotherapy-associated changes of serum phospholipids in patients with malignant lymphomas,
Magn. Reson. Med. 35,449--456.
13. Kuliszkiewicz-Janus, M. and Baczynski, S. (1997) Treatment-induced changes in 31 P-MRS (magnet-
ic resonance spectroscopy) spectra of sera from patients with acute leukemia, BBA 1360,71-83.
14. Kuliszkiewicz-Janus, M., Janus, W. and Baczynski, S. (1996) Application of P NMR spectroscopy in
clinical analysis of changes of serum phospholipids in leukemia, lymphoma and some other non-
-hematological cancers, Anticancer Research 16, 1587-1594.
15. Bettaieb, A., Record, M., Come, M.G., Bras, A.c., Chap, H., Laurent, G., and Jafrezou, J.P. (1996)
Opposite effects of tumor necrosis factor alpha on the sphingomyelin-ceramide pathway in two
myeloid leukemia cell lines: role of transverse sphingomyelin distribution in the plasma membrane,
Blood 88,1465-72.
16. Bruno, A.P., Laurent, G., Averbeck, D., Demur, C ., Bonnet, J., Bettaieb, A., Levade, T ., and
Jaffrezou, J.P. (1998) Lack of cerami de generation in TF -I human myeloid leukemic cells resistant to
ionizing radiation, Cell Death Differ. 5, 172-82.
17. Jarvis, W.O., Fornari, F.A., Traylor, R .S., Martin, H.A., Kramer, L.B, Erukulla, R.K, Bittman, R.,
and Grant, S. (1996) Induction of apoptosis and potentiation of ceramide-mediated cytotoxicity by
sphingoid bases in human myeloid leukemia cells, J. BioI. Chern. 27, 8275-84.
18. Szala, S. (2000) Swoista indukcja apoptozy w kom6rkach nowotworowych (in Polish), Nowotwory
50, 1\1-121.

6. Abbreviations

ARA-C-cytarabine, CPLAS-phosphatidylcholine-plasmalogen; ~-acetyl-y-O-alkyl


L-a-phosphatidylcholine, DHAD-mitoxantrone, LPC-Iysophosphatidylcholine,
MRS-magnetic resonance spectroscopy, PC-phosphatidylcholine, PE-phos-
phatidylethanolamine, Pi-inorganic phosphate, PI-phosphatidylinositol, PS-phos-
phatidylserine, SM-sphingomyelin.7
VANADIUM-51 3QMAS NMR AND ITS APPLICATION FOR THE
STUDIES OF VANADIA BASED CATALYSTS.

O.B. LAPINA\ P.R. BODART2, J-P. AMOUREUX2


I Boreskov Institute of Catalysis, pr.Lavrent 'eva 5, 630090
Novosibirsk, Russia, olga@catalysis.nsk.su
2Universite de Lille-I, 59655, Villeneuve d'Ascq, France

1. Introduction

Yanadia based catalysts are widely used in industry for many purposes such as,
selective oxidation of hydrocarbons, reduction of nitrogen oxides with ammonium
(SCR), and cleaning of flue gases. Despite they have been extensively investigated
by modem physico-chemical techniques, complexity of the catalytic systems and
technical limitations have prevented a clear view of the active species structure.
During the last few years, the Multi-Quantum Magic Angle Spinning
(MQMAS) was largely used for precise characterization of local structure of
quadrupolar nuclei with half integer spin [1-3]. Simple in realization and very
effective in resolution this method has became very popular, and spectacular results
have been obtained with a number of nuclei (aluminum, sodium, rubidium, boron)
[3-7]. However, it is commonly admitted that small value of quadrupolar moment
and large value of chemical shielding anisotropy are a limitation to MQMAS and,
in particular, prevent its application to vanadium nucleus [8].
In this communication, we demonstrate that MQMAS technique can be
successfully applied to vanadium nucleus and used for the study of supported
vanadia catalysts.

2. Experimental

The 51y 3QMAS spectra were recorded on a BRUKER DSX-400 spectrometer


operating at 105.2 MHz for Sly. A 2.5-mm MAS probe was used at spinning
frequencies in the range 30 to 35 kHz. The three-pulse Z-filter sequence was
employed for the recording of pure absorption 3QMAS spectra with great
sensitivity [9-11]. The lengths of the two first pulses were respectively 5 and 2.25
IlS with a RF field strength of few tens of kHz for vanadium resonance. The third
pulse (of duration lOllS) was adjusted to be a selective 90° on the central transition.
The delay between the two last pulses was set to 1 ms. The experimental time
355
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 355-363.
© 2002 Kluwer Academic Publishers.
356

varies between 2 and 10 hours but for all the standard samples, a resolved spectrum
could be obtain in less than 4 hours. A 2s recyCling delay was used, except for
NH4 V0 3 which was recorded with recycling time of 10 s; 288 transients were
recorded with 64 increments in the indirect dimension (Fl) and the spectral width
in Fl was set equal to the spinning frequency. The spectrum of 15%V20s1Zr0 2 was
recorded with a 4-mm MAS probe spinning at 15 kHz.

3. Results and Discussion

Due to the small value of vanadium quadrupolar moment, the SAtellite TRAnsition
Spectroscopy method (SATRAS) has proven to be a convenient NMR technique
for vanadium characterization [12-13]. Automatic analysis of the intensity of well-
resolved satellite spinning sidebands allows determining the complete set of
quadrupolar and Chemical Shielding Anisotropy (CSA) tensor parameters as well
as their relative orientations. Thanks to SATRAS, precise NMR data have been
obtained for most of the individual vanadium-oxide compounds of the system
V20s.MxO y (M = mono-, di-, three- and tetra- valent metals) [l3-16]. In real
systems (e.g. catalysts), defects and distortions of the structures induce
distributions of quadrupolar and CSA parameters. These distributions broaden
both, static and MAS spectra, and completely prevent any SATRAS measurement.
Moreover, it has been shown for VO/fi02 catalyst, that large quadrupolar
coupling constant values could be observed in real systems [17]. It is therefore
clear that the analysis of these materials require a technique, other than SATRAS
since spectra with resolved satellite spinning sidebands cannot be produced.
In this work, we have performed a methodological study for applying
MQMAS technique to the characterization of different types of vanadium sites in
oxide systems.
Data processing and quantification of the sheared spectra was performed
according to the following approach. In a sheared MQMAS spectrum, isotropic
chemical shift (v iso) and composite quadrupolar coupling constant (f..), both in
Hz units, can be extracted from the position of the center of gravity of the line
(V~\l2' V~id)' according to [11]:
G ')...2
v ±112 =viso - Vo AI,1/2
(1 )

with

(2)
357

C _ 34m(4m 2 -1) (3)


1m -
, 9(-3+41+41 2 )

(4)

(5)

v 0 is the Larmor frequency.


Inversion of the previous system (1) gives
G G
Vid AI,1I2 +v±1I2D1,m
viso =
AI,1I2 C I,m +DI,m (6)

and the application of the classical error propagation law gives

L1Viso =Ie A 1 +D
I,m 1,112 I,m
I~AI'1I21L1Vid +IDI,mlL1V~/J=I~~(~~~~2)IL1Vid + ~~L1V~/2 (7)

In the case ofI=7/2 and 3Q experiment m=3/2, we obtain


5 G 10 G
L1 Viso = '6 L1 Vid + 27 L1 V±1/2 (9)

v 0 490 ( G 34 G ) (10)
L1A.=T3 L1V id + 4S L1V ±1I2

Observed lines are very sharp and well resolved, for all the sites in all the samples
studied therefore the spectra have been processed in power mode, to avoid errors
coming from any misseting of the phases. The centers of gravity of the lines have
been calculated by integration of the summations on the axes of columns (or rows)
that were associated with the peaks. L1V~ and L1V~1I2 have been estimated has
1110 of the linewidth at half height of the summation on both axes (Figure 1). All
358

the lines were very similar and we took dV~d=150Hz


1
and Av G 12=50
±l
Hz for all the
experiments. ~ v iso is independent of v iso and A, and is constant for all the lines:
1.4 ppm, whilst !!J.. is inversely proportional to A.

Figure J Summation onto the isotropic axes of peak 3


of AIVO. the Iinewidth is estimated to be
1500Hz and the dV~ estimated as 150Hz is
represented as an horizontal line, situated at
the maximum of the peak.

-65 kHz -60


5ly 3QMAS spectra were obtained for of all types of vanadium sites and spectra
are presented here:

REGULAR TETRAHEDRAL OXYGEN ENVIRONMENT OF QO TYPE


According to SATRAS data [14, 18-20] for vanadium in regular isolated
tetrahedral oxygen environment (QO type), almost spherically symmetric chemical
shielding tensor with small value of anisotropy is typical. The quadrupolar
coupling constant can vary from 1 to 6 MHz and the chemical shielding asymmetry
parameter can take values between 0 and 1.
Lanthanum vanadate (LaY04 ) is a convenient sample to compare MQMAS
and SATRAS since it has a very simple structure, there is a single vanadium site
forming an isolated regular tetrahedron characterized by a large quadrupolar
coupling constant. The 5ly 3QMAS spectrum of LaY04 (Figure 2) shows an
intense signal, and the parameters deduced from the frequencies of the center of
gravity of the line (according to Equations 6, 9 and 10) agree with the measurement
performed by SATRAS. Indeed, there is a good agreement for the quadrupolar

r ·700
·850

·800

~. ~ Il '[ ·750
S.

_ ·500
l .700

·650
~ _ _ ~_-,-I -400

-400 ·500 ·600 ·700 -800 ·600


~,,(ppm)

Figure 2. Slv 3QMAS spectrum of LaV04 Figure 3. SIV 3QMAS spectrum of AIVO.
359

parameter and in particular the incertitude tJ.').. is small since ').. relatively large.
However, a small difference is observed for the isotropic chemical shift values
(Table I .). This may be attributed to a temperature effect. Indeed, during the
rotation, the friction of the rotor increases the temperature of the sample. At a
spinning speed of35 kHz the sample temperature can increased by 20° and induced
a line shift. In the case of vanadium containing compound this shift may be
indirectly induced by the strong temperature-dependence of paramagnetic centers
(y4+) that may be present in sample as impurities. Nevertheless, a shift of the line
only affects the calculated isotropic chemical shift value and does not corrupt the
calculated').. value.
Figure 3 shows the 51 y 3QMAS spectrum of AlY04. The crystalline
structure is composed of three nonequivalent vanadium sites of type QO. The
overlapping spinning sidebands observed in MAS, preclude SATRAS application,
while the high resolution obtained in the 3QMAS spectrum allow a precise
quantification of the isotropic chemical shift and ').. parameter. This case clearly
illustrates the superiority of MQMAS over SATRAS, when several sites produce
overlapping spinning sidebands.

SLIGHTLY DISTORTED TETRAHEDRAL SITES OF Ql TYPE


For vanadium atoms in slightly distorted tetrahedral sites, sharing one oxygen atom
with one adjacent tetrahedron (Ql type), asymmetric chemical shielding tensor is
typical, but with larger value of anisotropy than in QO environment. Quadrupolar
coupling constant varies from 2.0 to 10.0 MHz and chemical shielding asymmetry
parameter changes from 0.1 to 0.9. [16, 18-20]
Figure 4 shows 51 y 3QMAS spectrum of BaN207. In this sample,
vanadium atoms are distributed between three nonequivalent QI sites. As in the
case of AIY04, the overlapping spinning sidebands observed in MAS, limit
SATRAS application, while a fully resolved 3QMAS spectrum can be obtained and

~~ -700,- - - - - - - -A

-600
E
0..
E -600
Q.
0..
~
::5 -550 J
-500 QIS
-5001'---~--~----'
-500 -550 -600 -650
-500 -600 -700
Ih/(ppm)
o2l'(ppm)
Figure 4. Sly 3QMAS spectrum of BaNIO, Figure 5. Sl y 3QMAS spectrum ofNH,vOl
360

allows a characterization of the material, (extracted NMR parameters with those


obtained from SATRAS are given in Table 1.)

STRONGLY DISTORTED TETRAHEDRAL SITES OF Q2 TYPE


For vanadium atoms in strongly distorted tetrahedral sites sharing two oxygen
atoms with two adjacent tetrahedral units (Q2 type), an asymmetric chemical
shielding tensor with a large anisotropy is expected, the chemical shielding
asymmetry parameter is in the range 0.6 to 0.8 and the quadrupolar coupling
constants varies typically from 3.0 to 6.0 MHz [13-15, 18-20].
Figure 5 shows the SIV 3QMAS spectrum of NH4V0 3 • Despite of a large
chemical shielding anisotropy in that compound (245 ppm) [14] a very good
spectrum is obtained in 3QMAS experiment (see Figure 5).

PENTA- AND OCTA- COORDINATED VANADIUM


Penta-coordinated vanadium atoms have a nearly-axial large shielding anisotropy.
In this environment, the asymmetry parameter varies from 0.3 to 0.6 and the
isotropic shift is in the range -510 to -520 ppm. The quadrupolar coupling constant
has moderate value between 2.0 and 4.0 MHz [15, 20].
Vanadium atoms in distorted octahedral environment have also a nearly-
axial shielding anisotropy with a large value of anisotropy, but the asymmetry
parameter is close to zero and the quadrupolar coupling constant does not exceed
2 MHz
Figures 6 and 7 show the 3QMAS spectra of two compounds (CS 2V40 11

1\
~J\",j
-600.,-_ _ __ __ --" ~
-600

~
-550 '------
5.c. -550 i ,-~
:=i c. V )

-500 j -500~-----::=------1 ~
-500 -550 -600
li2"(ppm)
Figure 6. Sl y 3QMAS spectrum or cs.YIO " Figure 7. Sl y 3QMAS spectrum of Rb /Y 60 ,6
361

and Rb 2V60 16) containing nonequivalent vanadium atoms with five or six
coordinations.
In CS2V40 l1 there are three types of vanadium sites: two in distorted
tetrahedral pyramids and the third one with octahedral coordination. Because of
the overlapping sidebands in MAS, it is impossible to measure all NMR parameters
for each vanadium sites by SATRAS. While, the well resolved 3QMAS spectrum
observed for this compound allows to determine the isotropic chemical shift and
the composite quadrupolar coupling constant.
In Rb 2V6016 there are two types of vanadium sites, both in octahedral
coordination. A well resolved 3QMAS spectrum is observed for this compound,
and although both sites are observed, the quadrupolar coupling constant of the
weak line is not easily measured from the spectrum, only a maximum value can be
estimated.

TABLE 1. Composite quadrupolar coupling constant (A) and isotropic chemical shift (Oiso)
measured by 3QMAS and SATRAS techniques.
A / (MHz)
3QMAS ID 3QMAS ID
4.9 (0.7) 2.48* -659.2 (1.4) -662
3.5 (0.9) 2.48* -741.4 (1.4) -743
3.8 (0.8) 2.48* -773.2 (1.4) -776

LaV04 6.0 (0.5) 5.6 -604.8 (1.4) -609

NH4V0 3 3.1 (1.0) 2.95 -564.7 (1.4) -563.7

Rb 2V60 16 2.47 (1.3) 3.12 -514.3 (1.4) -515


<1.8 (1.8) 2.37 -547.6 (1.4) -547

CS2V4O l1 1.9 (1.7) 1.29 -511.9 (1.4) -513


2.6 (1.3) no -569.8 (1.4) no
2.4 (1.4) 1.81 -577.7 (1.4) -575

Ba2V207 2.27 (1.4) 2.25* -579.6 (1.4) -579


3.5 (0.9) 2.25* -587.8 (1.4) -589
1.1 (2.9) 2.25* -599.6 (1.4) -600
For the 3QMAS experiments, errors are reported in brackets.
no: not observed
* the Avalue could not be accurately determined by SATRAS and was estimated.
NMR parameters extracted from 3QMAS spectra and obtained by SATRAS
experiments are collected in Table 1. There an overall is very good agreement
between 3QMAS and SATRAS data. For complex spectra with overlapping lines,
MQMAS has significant privilege. However, the high value of the !1f.. incertitude
362

when 'J. . is small has to be deplored and for compounds with very small quadrupolar
coupling constant SATRAS may be more accurate than MQMAS. However, a
formal prove would require an error analysis on the 'J. . value given by the SATRAS
technique. In fact, it is likely that in SATRAS, 1l'J... increases when 'J. . increases,
similarly to what is occurring in MQMAS experiments, because the extraction of
the quadrupolar parameter, in both experiments, relies on the measurement of the
second order quadrupolar shift of the satellite transitions.
Table 1 shows that the incertitude Il A on the 'J. . value measured in
MQMAS experiments is large when 'J. . is weak. In fact, the accuracy could be
improved by increasing the quantum order m and decreasing the static magnetic
field. As can be deduced from the development of Equation (8) given below, in
Equation (11):

/)'A=~{80(1 - 2Il12 /).yG +27201\21-1) /).yG } (11)


2A 9m(4m2 -I) ±m 81(3 + 21) ±1I2

However, at low field .1V~\l2may increase since the quadrupolar lines are then
broader, and furthermore the accuracy of Yiso may decrease, depending on how
~v~ evolves with the quantum order (i.e. how the linewidth in the Fl dimension
evolves with the quantum order).
The results obtained show that combination of ultra high rotation
frequency (30-35 kHz) with low power radio-frequency excitation allows the
application of MQMAS to vanadium nucleus at all vanadium coordination.
However, when the quadrupolar interaction is weak the incertitude on its measure
is important but this does not affect the accuracy on the isotropic chemical shift.

~
·700
u' ~
/ ~
/
1·600

,'~,
,e.
~

Figure 8. Sheared. Sly 3QMAS spectrum of

:-1-'-"-"-"_"..,..'_--..__..--1
-400 ·500 ·600 ·700
t 15%Y20s/Zr02 recorded at a spinning frequency of IS
kHz.

2'(ppm)
These results are very promising and we have tried to apply the technique
to real catalyst systems, spectra of VOx/zr0 2 system have been recorded and
Figure 8 shows the spectrum of 15%V20sIZr02• This spectrum shows a signal that
is aligned with the isotropic chemical shift direction indicating that (i) the
quadrupolar coupling constants of vanadium atoms in the material are weak, (ii)
363

the chemical shift interaction is largely distributed.

Acknowledgments

This work was supported by RFBR 01-03-32364, INTAS 97-0059, PAl 04522WK
grants. One of the authors (0. Lapina) thanks Universite de Lille for the visiting
professor grant.

4. References
I. Frydman L., Harwood J.S., (1995) Isotropic Spectra of Half-Interger Quadrupolar Spins from Bidimensional
Magic-Angle Spinning NMR, J.Am.Chem.Soc. 117,5367-5368
2. Medek A., Harwood J.S., Frydman L, (1995) Multiple-quantum magic angle spinning NMR: A new method
for the study of quadrupolar nuclei in solids, J.Am. Chem.Soc. 117, 12779-12787
3. Fernandez c., Amoureux l-P., (1996), Triple-quantum MAS NMR of quadrupolar nuclei, Solid State NMR,
5,315-321
4. Amoureux, J.-P., Fernandez, c., (1998) Triple, quintuple and higher order multiple quantum MAS NMR of
quadrupole nuclei, Solid State Nllci. Magn.Reson. 10,211-223.
5. Fernandez c., Amoureux l-P., (1995) 20 multiquantum MAS-NMR spectroscopy of 27AI in
aluminophosphate molecular sieves, Chem.Phys.Lett., 242,449-454
6. Peeters M.P.1., Kentgens A.P.M., (1997) A 27AI MAS, MQMAS and off-resonance nutation NMR study of
aluminium containing silica-based sol-gel materials, Solid State Nuc/.Magn.Resos., 9, 203-217
7. Vosegaard T., Florian P., Grandinetti P., Massiot D., (2000) Pure Absorption-Mode Spectra Using a
Modulated RF Mixing Period in MQMAS Experiments, Jour..Magn.Reson., 143,217-222.
8. Medek, A., Frydman, L. (1999) Quadrupolar and Chemical Shift Tensors Characterized by 20 Multiple-
Quantum NMR Spectroscopy. J.Magn. Reson., 138(2),298-307
9. Brown S.P., Heyen S.1., Wimperis S. (1996) Two-Dimensional MAS Multiple-7Quantum NMR of
Quadrupolar Nuclei. Removal of Inhomogeneous Second-Order Broadening, Jour. Magn.Reson., A, 119,
280-284
10. Amoureux, J.-P., Fernandez, C., Steuenagel S., (1996), Z-filtering in MQMAS NMR, Jour. Magn.Reson.A
123,116.
I J. Bodart, P.R. (1998) Distributions of the Quadrupolar and Isotropic Chemical Shift Interactions in Two-
Dimensional Multiple-Quantum MAS NMR Spectra, J. Magnt. Reson. 133,207-209.
12. Skibsted, l, Nielsen, N.C., Bilds0e H., Jakobsen, H.1. (1992) SIV MAS NMR spectroscopy: determination of
quadrupole and anisotropic shielding tensors, including the relative orientation of their principal-axis
systems, Chem.Phys.Lett. 188,405-412.
13. Skibsted, J., Nielsen, N.C., Bildsee H., Jakobsen, H.1. (1993) Magnitudes and Relative Orientation 0 SIV
Quadrupole Coupling and Anisotropic Shielding Tensors in Metavanadates and KV 30 s from SIV MAS NMR
Spectra., J.Am. Chem.Soc. 115, 7351-7362.
14. Skibsted, l, Jakobsen, H.J.H., Jakobsen, H.1., (1998) SIV Chemical Shielding and Quadrupole Coupling in
Ortho- and Metavanadates from SIV MAS NMR Spectroscopy, lnorg.Chem. 37,3083-3092.
15. Nielsen, U.G., Jakobsen, H.J., and Skibsted 1 (2000) Characterization of Divalent Metal Metavanadates by
SIV Magic-Angle Spinning NMR Spectroscopy of the Central and Satellite Transitions, Inorg. Chem. 39,
2135-2145.
16. Nielsen, U.G., Jakobsen, H.J., and Skibsted J. (2001) SIV MAS NMR Investigation of SIV Quadrupole
Coupling and Chemical Shift Anisotropy in Divalent Metal Pyrovanadates, J. Phys. Chem. 8, lOS, 420-429
17. Shubin A.A., Lapina O.B., Bondareva V.M., (1999) Characterization of strongly bonded V(V) species in
VO/Ti0 2 catalyst by static and MAS solid-state SIV NMR spectroscopy, Chern. Phys. Lett., 302,341
18. Lapina O.B, Mastikhin V.M., Shubin A.A., Krasilnikov V.N., Zamaraev K.I., (1992) SIV Solid State NMR
Studies of Vanadia Based Catalysts, Progress in NMR spectroscopy, 24, .457-525.
19. Mastikhin V.M., Lapina O.B., (1996) Vanadium Catalysts: Solid State NMR, in Encyclopedia of NMR
(Nuclear Magnetic Resonance); Grant D.M., Harris R.K., Eds.; John Wiley&Sons Ltd.; Chichester, UK, .8,
4892-4904.
20. Khabibulin D.F., Shubin A.A., Lapina O.B., (2001) Correlation of SIV NMR parameters with Local
Environment of Vanadia Sites, NATO ARW, Magnetic Resonance in Colloid and Interface Science,
St.Petersburg, 108
SIMULTANEOUS EPR AND TPR STUDY OF THE V-Ce-O
CATALYSTS REDOX PROPERTIES

J. MATTA, E. ABI-AAD*, D. COURCOT, A. ABOUKAIS


Laboratoire de Catalyse et Environnement E.A. 2598, Universite du
Littoral - Cote d'Opale, MREID, 145 avenue Maurice Schumann, 59140
Dunkerque, France. * abiaad@univ-littoralfr

Abstract
Vanadium cerium oxides, with different VICe atomic ratios, were prepared using the
impregnation method and calcined under air at 500°C. Physico-chemical studies have
shown that at low vanadium content, polymeric V-0-V chains are stabilised on the ceria
surface. Increasing the vanadium content tends to favour the formation of the CeV04
and V20 5 phases. The redox properties of these oxides have been simultaneously
investigated by TPR and EPR techniques. V-0-V chains and V20 5 species are more
easily reducible than the CeV0 4 phase. The reduction of V20 S to V20 3 proceeds in
several steps, the intermediate species being V6 0 13 , V0 2 and VS0 9 . The reduction of
V20 5 species interacting with ceria support leads to VO oxide. EPR measurements
performed at T=-269°C have permitted to observe progressively different signals of V4+
in addition to vanadium ions in V2+ (3d3) paramagnetic configuration. This attribution is
based on an EPR signal at g=3.956 with eight well resolved hyperfine lines
(A=96Gauss), which may be attributed to the perpendicular components of one of the
fine transitions corresponding to the V2+ spectrum. At high reduction temperature,
CeV0 4 phase leads in one step to CeV03 and a continuous and partial reduction of
Ce02 into CeZ03 is observed.

1. Introduction

The Ce02 and V20 5 oxides are commonly used as catalysts in the depollution reactions
such as the oxidation of diesel soot particles. Vanadia (V 205) based catalysts are proved
to be quite effective for Selective Catalytic Reduction (SCR) of nitric oxide NO and
also in different oxidation processes [1-2]. In parallel, ceria (Ce02) has traditionally
been used as a promoting component in the three way catalysts for automotive exhaust
treatment. It exhibits a large deviation from its stoichiometric composition (Ce02_x)
making ceria a promising material for use either as a support or active catalyst in
oxidation - reduction reactions [3].
Recently, different V-Ce-O oxide catalysts have been prepared in order to be used in the
catalytic oxidation of diesel soot particles [4]. The combination of different physico-
chemical techniques has allowed us to define the nature of different vanadium oxide
species present in the V -Ce-O compounds calcined at different temperatures from 400
to 800°C. In the temperature range of 400-550°C, polymeric V-0-V chains highly
365
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Inteiface Science, 365-374.
© 2002 Kluwer Academic Publishers.
366

interacting with ceria surface are identified for low vanadium content. With the increase
of the vanadium content vanadium tetrahedral surface species and then the V20 5 phase
are evidenced in temperature range of 400 to 4S0°C. When the solids are heated at high
temperatures C2:S00°C), the formation of the CeV0 4 phase occurs from the reaction
between V20 S and Ce02 oxides. Consequently, a single electron is trapped in an oxygen
vacancy and can be considered as a probe to the CeV04 phase presence [S] .
The determination of the nature of vanadium species can provide useful information
leading to a better understanding of the elementary steps of heterogeneously catalysed
reactions. In fact, the basic steps of catalytic oxidation processes at the surface of metal
oxides consist in redox interactions involving the solid and one or more reactants. In
this order, the redox properties of the identified vanadium species are simultaneously
investigated by Temperature Programmed Reduction (TPR) and Electron Paramagnetic
Resonance (EPR).

2. Experimental

2.1. SOLID PREPARATION

Ceria (Ce02) was prepared by precipitation of cerium hydroxide from Ce(N03)3,6H 20


with a NaOH solution as described in [6]. The solid is calcined at SOO°C for four hours
under a flow of dried air. Subsequently, a solution of vanadyl oxalate VOC20 4 is
impregnated on ceria to prepare xVIOCe samples with different VICe atomic ratios
(x=1.08, 1.67, 2.73 and S.06), deduced from chemical analysis [S]. After drying at
100°C, the solids freshly prepared are calcined between 400°C and 800°C under a flow
of dried air for 4 hours.

2.2. TEMPERATURE PROGRAMMED REDUCTION

Temperature Programmed Reduction (TPR) measurements were performed on a


NETZSCH STA 409 apparatus. About 100 mg of calcined samples at SOO°C
(xVlOCeSOO) or reference compounds (Ce02, V20 S and CeV0 4) were considered and
placed in a Ah03 crucible. All solids were first treated under dried air at 300°C for one
hour and then cooled to 20°C in order to eliminate the physisorbed water due to ambient
humidity. The TPR results were corrected with respect to the weight loss corresponding
to water departure. The temperature of the samples was raised at a rate of SoC min-l
from SO to 9S0°C under a flow (7SmL.min- l) of 10 vol. % H2 in N 2. The exploitation of
the weight losses has permitted to evaluate the reducibility properties of such samples in
these conditions.

2.3. ELECTRON PARAMAGNETIC RESONANCE

The Electron Paramagnetic Resonance (EPR) measurements were performed at -269°C


on a EMX BRUKER spectrometer with a cavity operating at a frequency of -9.5 GHz
(X band). The magnetic field was modulated at 100 kHz. The g values were determined
from precise frequency and magnetic field values. For EPR measurements, the samples
were placed in a micro flow reactor and treated under a flow (7SmL.min- l) of 10 vol. %
367

H2 + 90 vol. %N 2 in the same conditions as above. The microflow reactor was


assembled with an EPR quartz tube into which the catalyst could be transferred under
the same atmosphere.

3. Results and Discussion


3.1. TEMPERATURE PROGRAMMED REDUCTION

The redox properties of the V-Ce-O catalysts, associating the oxidation states Ce 4+/Ce 3+
and V5+/V4+, have been then investigated by TPR. The TPR profiles of the xV I OCe500
samples (x = 1.08; 1.67 and 2.73) are shown in "Figure 1.". Corresponding
thermogravimetric (TG) measurement values are in TABLE I.

100 300 500 700 900


Temperature (OC)
"Figure I." TG and DTG curves obtained over the TPR o/the xVIOCe500 samples

For all the samples, different weight losses were reported: the first one started slightly
above 400°C and is clearly observed in the range of 490-530°C, the second, evidenced
at higher temperature (600°C<T<730°C) depends on the vanadium content in the
samples. Indeed, a weight loss is observed at 665°C for 1.67VlOCe and occurs at 657°C
for 2.73VI0Ce. Finally, a continuous weight loss, clearly observed for the 1.08VIOCe
sample in the range of 770°C, is evidenced on the TG curves, for all the xVlOCe
samples, until 950°C. In order to have a reliable attribution ·of these different
phenomena, Ce02, V 20S and Ce V0 4 oxides have been investigated by TPR "Figure 2. ".
368

TABLE I. Experimental and theoretical weight losses (%) obtained from the TPR measurements

4000C<T<6000C 600°C<T<730oC T>700oC


Sample V20s--.2VO CeV0 4 --. CeV03 2Ce02 --. Ce203
total loss theoretical el(perimental theoretical el(perimental theoretical el(perimental
I.08VIOCe 3.040 1.425 1.420 4.397 1.620
I.67VIOCe 3.510 1.523 1.520 0.414 0.414 4.070 1.576
2.73VIOCe 4.380 1.773 1.770 1.047 1.047 3.420 1.563

The presence of these latter oxides in the xVIOCe500 samples have been highlighted in
a previous work where each xVIOCe500 solid could be considered as a combination of
these three oxides in different proportions [5].

-I 1
~ -2
!
(,!) -3
....
-4

-5

-6
100 200 300 400 500 600 700 800 900
Temperature (0C)

"Figure 2. " TG curves obtained over the TPR a/CeO]. V1o., CeV04 and 5. 06 VI OCe500

The Ce02 oxide calcined at 500°C and used as support for the xVIOCe samples is not
affected by the reducing conditions of the TPR treatment "Figure 2.". However,
literature data [7] clearly show surface and bulk reduction of ceria at -500°C and
-850°C respectively. The behaviour observed in our case can be related to the residual
alkaline cations resulting from the preparation method of the Ce02 support. In fact, the
presence of these cations has been revealed by the chemical analysis (320 ppm of
sodium) while the calcination treatment induces their migration to the solid surface [8].
This alkaline layer would protect the ceria surface from reduction by H2.
The reduction of V20 5 to V203 is evidenced between 600°C and 790°C and proceeds in
several steps, the intermediate species being V60 13 , V0 2 and V50 9 • Experimental and
theoretical weight losses associated to different reduction steps are reported in TABLE II.
These data are in agreement with those observed in the literature showing the reduction
of V20 5 into V 203 in this range of temperature [8]. Nevertheless, the impurities content,
the preparation method and the partial pressure of hydrogen can have major influence
on the reduction level of vanadium oxide. Generally, the reduction of V 203 into VO is
observed over a temperature of 1000°C.
369

TABLE II. Experimental and theoretical weight losses (%) obtained from the progressive reduction ofV 20 5

Reaction Tem~erature ex~erimentalloss theoretical loss


V20 S -+ V6 0 13 625°C 5.84 5.86
V60 13 -+ V0 2 640°C 2.93 2.93
V0 2 -+ VS09 710°C 3.53 3.51
VS09 -+ V20 3 775°C 5.23 5.28

Finally, CeV0 4 phase is reduced in one step at 705°C. The mass loss of 6% obtained
during this reduction can be due to the formation of CeV0 3 in agreement with
Yokokawa et al [9,10].
For the first weight loss observed in the range of 400-600°C, the theoretical values,
deduced from the composition of the sample [5] and the reduction of V20S into V203,
are always smaller than the experimental losses. Thus, the possibility of a more
pronounced reduction of V20S into VO is considered and the resulting theoretical losses
are reported in TABLE 1. These values are in perfect agreement with the experimental
losses measured from the TG curves which confirms the reduction ofV 20 s into YO. In
addition, the TPR results clearly show that the polymeric V-O-V chains (1.08VI0Ce)
are more easily reducible than V20S phase. However, it is already demonstrated that the
increase of the vanadium content leads to a decrease of the" vanadium species - ceria
support" interaction [5] . Thus, similar to palladium and copper supported on ceria
[11 ,12], the V20 S - Ce02 interaction facilitates the reduction of the vanadium species.
For higher vanadium loading, 1.67VI0Ce, 2.73VlOCe and 5.06VI0Ce, the V20 S - Ce02
are more easily reducible than pure V20S oxide" Figures 1. and 2. ".
The second mass loss (TABLE I) increases with the increasing of the vanadium content
and can be attributed to the reduction of CeV0 4 phase, present in the solids, into
CeV0 3. In fact, this CeV0 4 phase is missing in the 1.08VI0Ce500 sample [5] which
explains the absence of the weight loss in the temperature range of 600-730°C in this
case "Figure 1.". In addition, the reduction temperatures are shifted towards smaller
values with the increase of the vanadium content "Figures 1. and 2.". The presence of a
trapped electron nearby the CeV04 phase [5], the VO and Ce02 oxides would facilitate
the reduction ofCeV0 4 into CeV0 3 .
In all the cases and on the contrary of pure oxides "Figures 1. and 2.", a continuous
weight loss is observed until 950°C. In addition, the total experimental weight loss is
considerably higher than the theoretical values corresponding to the reduction of V20 S
and Ce V0 4 . Thus, a partial reduction of the Ce02 support is evidenced (TABLE I) for
all the xVI0Ce500 samples treated in these conditions.

3.2. ELECTRON PARAMAGNETIC RESONANCE

The general redox properties of a metal oxide can be investigated by means of the
electron paramagnetic resonance technique which is capable of detecting the
paramagnetic centres often formed upon one electron redox processes. In fact, EPR
offers a powerful and sensitive technique for investigating the oxidation states, surface
and bulk co-ordination and the physical form of a transition metal oxide. EPR technique
has been extensively used to characterise supported vanadia systems in order to
370

elucidate the nature of the paramagnetic species, their local structure and the effect of
the active phase - support interaction [13].
EPR spectra of ions with S>Y2 and 1:;t0 may be analysed using an axial symmetry spin-
Hamiltonian:
H= g//pHzS z +g.lP(HxS x +HySy)+A//lzSJ +A.l(IxSx +lrS y )+ 1
DlS~ -tS(S+I) +E(S~ _S~)eq.
where D and E are the crystal-field parameters, S and I denote the total electron and
nuclear spins, g is the g-factor and A is the Hyper Fine Structure (HFS) constant. In the
case of S=Y2, D=E=O.
"Figure 3. " shows the EPR spectra, recorded at -269°C, obtained before and after the
reduction of the 1.08VI0Ce500 sample. For the calcined sample, the EPR spectrum is
the superimposition of two signals with g<2 denoted by A and B and described
elsewhere [5, 6]. The A and B signals are characterised by gxx(A)=I.9654, gYY(A)=1.9626,
gzz(A)=1.9398 and gxx(B)=1.9692, gyy(B)=1.9670, gzz(B)=1.9465 and attributed to the
presence of two different Ce 3+ (fl ions; ge>g.L>g//) sites in the solid or more precisely to
an interaction between conduction electrons and 4f orbitals of Ce4+ ions in the Ce02
matrix. These signals are also present on the EPR spectrum of the ceria support before
impregnation. The A signal has been attributed to Ce3+ species with easily removable
ligands and the B signal to Ce 3+ ions stabilised by some lattice defects Ce02_x [6].
Finally, it is important to notice that no signal relative to V(IV) species is observed
indicating that the totality of the vanadium species are in V(V) oxidation state.
The EPR signal at g=4.27 "Figure 3. " is the well-known EPR transition for Fe3+ ions;
electronic configuration (3d\ This spectrum is characteristic of high spin Fe3+ ions
(ground state 6S 512 ), located in octahedral or tetrahedral [14] co-ordination sites having a
strong rhombic distortion of the crystal field. Indeed, if in (eq.l) where S=5/2 and 1=0
(since the natural abundance of the 57 Fe isotope (I=Y2), is only of 2.15%) the terms of
the crystalline field have the same values as the Zeeman term or are larger (D,E~gPH),
the spectral features are strongly affected by the relative magnitudes of the coefficients
D and E. In the case of a" purely rhombic" ligand field, IEIDI=1I3, an isotropic signal
at g=4.27 is observed. The line with g=4.27 is due to transitions within the middle
Kramers doublet [15-19]. The lower and upper doublets exhibit strongly anisotropic
values of effective g-factors which give rise to very low intensity of corresponding EPR
transitions in poly-crystalline solids. Fe3+ ions should be considered as impurities in the
solid. When reduced at 400°C the EPR spectrum of the I.08VIOCe500 sample exhibits
different EPR signals. It consists of several independent EPR spectra due to different
paramagnetic species. The most intense part of the spectrum has been attributed earlier
[20] to two sites ofV4+ (3d l) characterised by EPR parameters in eq. (1) with S=1I2 and
1=7/2 (99.76% of 51 V natural abundance with nuclear spin 1=7/2). Site{l) gll=1.923,
A I =174 G ; g.L=2.000, A.L=76 G, giso = 1.974, Aiso = 109 G and site(2) gll=1.890, A I =205
G; g.L=2.000, A.L=76 G, giso = 1.963, Aiso = 119 G. In the case of isolated V (IV)
species, information about their environment can be obtained from g and A values.
These could correspond to V0 2+ ions in octahedral symmetry axially distorted since
1.955< giso<I.980 and 80 G<A iso < 120 G [21]. Generally, lower giso and Aiso values are
expected in the cases of V4 + species in octahedral symmetry 1.920<giso<1.950 and
60 G<Aiso< 90 G [22-24].
371

1.08VIOCe 500 Gain

A+B
signals

Calcined

(g=4.27) 10

V""/
Reduced at 400°C

IIIII111

Reduced at 500°C

500 1000 1500 2000 2500 3000 3500 4000 4500 5000
H (Gauss)

"Figure 3." EPR spectra recorded at -269°Cfor the I.08 VIOCe500 sample at different temperatures of
reduction tr~atment

The second EPR signal is centred at g=3.956 and exhibits eight well-resolved HFS lines
with A=96 G. This EPR signal is also due to vanadium ions and could be attributed to
the perpendicular components of one of the fine transitions corresponding to the y2+
(3d3) spectrum. The theoretical estimation of the EPR parameters from eq. (1) with
S=3/2 and 1=7/2 within the limits of strong axial crystal field (g~H«D) gives rise to
axial symmetry spectrum with gll",,2 and gl.",,4 [25]. This prediction is in good agreement
with the experimental results found in the literature [26]. The gll",,2.0 component cannot
be seen in our case because of the large y4+ spectra intensities situated just in the range
of g=2.0. Moreover, the intensity of the parallel lines in an anisotropic spectrum is
generally weaker than that of perpendicular ones.
372

2.73VIOCeSOO Gain

Calcined
12.5
FeJ+

Reduced at 400'C

Reduced at SSO'C

2.5

AH9i52 G
1"'3.956
Reduced at 600'C

6
Reduced at 800'C

500 1000 1500 2000 2500 3000 3500 4000 4500 5000
H (Gauss)

Figure 4. EPR spectra recorded at - 269°C for the 2. 73 VIOCe500 sample at different temperatures of
reduction treatment

Theoretically, the EPR signal at g=3.956 could also be due to two other types of
vanadium paramagnetic centres. It could be interpreted:
(i) as a forbidden (LlMs=±2) transition of a pair y4+_y4+ (S=1 ; I=II+Iz=7) [23];
however, in this case, 15 (and not 8) HFS lines (2I+ 1), with the binomial intensity
distribution, should be observed
(ii) as the same type of the forbidden transition of y3+ (S=1 ; 1=7/2), but, regarding this
possibility, literature data confirm that it is unusual to observe y3+ ions because
such species are not stable. Their EPR spectra can be mainly recorded at
-269°C and at high frequency [26]. Moreover, in earlier works, this forbidden
transition (LlMs=±2) of y 3+ ions has never been observed at g=3.956 because of the
strong zero field splitting [26].
373

It should be important to notice that in our case, the 8 lines centred on g=3.956 have
been observed either at -269°C, -196°C or at room temperature [20].
For a reduction temperature of 500°C, the intensity of the EPR signal centred at
g=3.956 increases whereas, that of the signal corresponding to y4+ species decreases.
These results are in perfect agreement with those reported in the TPR part of this work.
Indeed, at this temperature of reduction (500°C) for the 1.08YIOCe500 sample, the
major part of the vanadium species is present as YO oxide "Figure 1. ".
The EPR spectrum of the calcined 2.73YI0Ce500 sample "Figure 4." shows, in
addition to the Fe3+ ions (g=4.27), two EPR signals in the region of g<2. The first is an
isotropic signal with g=I.964 and a line width of i1H=100G superimposed with a second
one presenting a very bad resolution of a HFS making impossible the determination of
the EPR parameters. Nevertheless, these signals can be attributed to y4+ species or
trapped electrons nearby vanadium atoms during the calcination of the solid [21] .
When reduced at 400°C the EPR spectrum of the 2.73YIOCe500 sample exhibits the
same EPR signals, in the range of g",,2, observed for the 1.08YlOCe500 sample and
assigned to Y02+ ions in octahedral symmetry axially distorted. On the contrary, the y2+
EPR signal (g=3.956) is not observed in this case confirming our TPR results. In fact,
with the increase of the vanadium content in the samples, the reduction of the vanadium
species proceeds at higher temperature. In this order, the 2.73YIOCe500 sample is
reduced at 550°C "Figure 4. " with respect to the TPR data showing the reduction of the
vanadium species at a temperature of 527°C "Figure 1.". A signal with g=3.956 and
i1H=652G is observed and can be attributed to the y2+ ions. However, in this case, the
vanadium content is considerably higher leading to the formation of agglomerates
which increases the dipolar interaction and consequently the broadness of the EPR lines.
An other signal centred at g=1.977 (i1H=95G) can be related to the presence of small
agglomerates ofy 4+ in the solid. With the increase of the reduction temperature (600°C)
the width (i1H=72G) and the intensity of this signal decrease "Fi!;.ure 4." indicating the
reduction of the y4+ ions. In parallel, the EPR signal relative to Y + ions is not observed
which could be explained by high dipolar interactions leading to a large line width of
the signal, mixed up with the base line of the spectrum. For a reduction temperature of
800 D C, the EPR spectrum is the result of the superimposition of different EPR signal
relative to y2+, y 3+ and Ce 3+ which makes the assignment of the EPR lines very
hazardous. Indeed, at this temperature the TPR results show the simultaneous presence
of the YO, CeY03 and Ce203 oxides.

4. Conclusion

The redox properties of Y -Ce-O oxides have been simultaneously investigated by TPR
and EPR techniques. Main information are deduced about the redox behaviour of the
different vanadium species in the samples. Thus, polymeric Y-O-Y chains and Y 20 S
species are more easily reducible than the Ce Y0 4 phase. The reduction of pure Y 20S to
Y 20 3 proceeds in several steps, the intermediate species being Y60 I3 , Y0 2 and YS0 9 .
The reduction of Y 20S species interacting with ceria support leads to YO oxide. EPR
measurements performed at T=-269°C have permitted to observe progressively different
signals of y4+ in addition to vanadium ions in y2+ (3d3) paramagnetic configuration.
374

This attribution is based on an EPR signal at g=3.956 with eight well resolved hyper
fine lines (A=96Gauss), which may be attributed to the perpendicular components of
one of the fine transitions corresponding to the V2+ spectrum. At high reduction
temperature, CeV0 4 phase leads in one step to CeV03 and a continuous and partial
reduction of Ce02 into Ce203 is observed.

5. Acknowledgments

The authors would like to thank the "Conseil General du Nord", the "Region Nord - Pas
de Calais" and the European Community (European Regional Development Fund) for
financial supports in the EPR and Thermal Analysis apparatus purchase.

6. References

J. AhlstrOm, A F., and Odenbrand A., (1990), Appl. Catal. 60,157.


2. Lietti L., and Forzatti P., (1994), J. Catal 147,241 .
3. Komer R., Ricken N., and Riess I., (1989), J. Solid State Chem. 78, 136.
4. Matta J., Abi-Aad E., Courcot D., and Aboukals A, (to be published)
5. Matta J., Courcot D., Abi-Aad E., and AboukaYs A, (in press) J. Solid State Chem.
6. Abi-Aad E., Bechara R., Grimblot J., and Aboukals A, (1993),1. Chem. Soc. Faraday Trans. 5,793.
7. Fomasiero P., Kaspar J., and Graziani M., (1999), Appl. Catal. B : Environmental 22, L11 and references
therein. .
8. Roozeboom F., Hazeleger, M.e., Moulijn, J.A., Medema J., De Beer, V.H.1., and Gellings, P.1., (1980),
1. Phys. Chem. 84, 2783.
9. Yokokawa H., Sakai N., Kawada T., and Dokiya M., (1990),1. Am. Ceram. Soc. 73,649.
10. Cousin R., Courcot D., Abi-Aad E., Capelle S., Amoureux J-P ., Dourdin M., Guelton M., and
AboukaYs A., (1999), Colloids and Surfaces, A 158,43.
1 J. Binet C., Jadi A., Lavalley, J.C., Boutonnet-Kizling M., (1992),1. Chem. Soc. Faraday Trans. 88(14),
2079.
12. Kundakovic L., Flytzani-Stephanopoulos A, (1998), Appl. Catal. A : General 171, 329.
13. Giamello, E., (1998), Catalysis Today 41,239.
14. Brodbeck, C.M., and Bukrey, R.R., (1981) Phys. Rev. 24, 2334.
15. Griscom, D.L., (1980),1. Non Cryst. Solids 40,211.
16. Schreurs, J.M ., (1978),1. Chem. Phys. 69,2151.
17. Dowsing, R.D., and Gibson, J.F., (1969) 1. Chem. Phys. 50, 1,294.
18. Zhilinskaya, E.A. , and Lazukin, V.N., (1982),1. Non-Cryst. Solids 50,163.
19. Lazukin, V.N., Chepeleva, LV., Zhilinskaya, E.A. and Chemov, A.P., (1975), Phys. Stat. Sol. B 69, 399.
20. Abi-Aad E., Bennani A., Bonnelle, J-P., and AboukaYs A, (1995), J. Chem. Soc. Faraday Trans. 914, 99
21. Abi-Aad E., Courcot D., Aboukals A, Baems M., Bruckner A., Gueiton M., and Vedrine, J.e., (2000),
Catal. Today 56, 37J.
22. Davidson A., and Che M., (1992), J. Phys. Chem. 96,9906.
23. Bogomolova, L.D., Khabarova, A.N., Klimashina, E.V., Krasilnikova, N.A., and Jackin, V.A., (1988),1.
Non Cryst. Solids 103, 319.
24. Centi G., Parathoner S., Trifiro F., AboukaYs A, Aissi C.F., and Guelton M., (1992),1. Phys. Chem. 96,
2617.
25. Zhidomirov, G.M. , (1975) "Interpretation of complex EPR spectra ", Nauca, Moscow.
26. Abi-Aad E., Zhiliskaya, E.A., and Aboukals, A, (1999), J. Chim. Phys. 96, 1519 and references therein.
STATIC AND DYNAMIC NMR STUDIES ON COSMETIC EMULSIONS

1. PLASS, D. EMEIS
Analytical Research Dept. BeiersdorfA G
Unnastrasse 48, 20245 Hamburg, Germany

Two cosmetic oil-in-water emulsions with a stearyl phosphate emulsifier are examined
by means of static and dynamic IH and 3lp NMR techniques. The interfacially bound
emulsifier can be detected by high resolution (HR) NMR spectroscopy, whereas the
excess emulsifier exists in one case as a solid lipid phase not detectable by this
technique. This emulsion and the emulsifier raw material, consisting of monostearyl
phosphate as well as distearyl phosphate, are examined by solid state cross polarisation
magic angle spinning (CPMAS) NMR spectroscopy to prove the existence of solid
emulsifier phases. In contrast, in the second emulsion the excess emulsifier together
with a co-emulsifier forms an emulsifier gel network detectable by long-time high
resolution 3lp NMR spectroscopy and water self-diffusion measurements.

1. Introduction

Emulsifiers are amphiphilic molecules which, due to their hydrophilic and lipophilic
molecular segments, can .be solvated at the interface of immiscible liquids, thus
reducing the interfacial tension. The development of emulsion-based cosmetic products
requires efficient emulsifier systems which ensure the stability of the dispersion even
after 3 years of storage. Although the composition of the formulation ingredients is
determined mainly from a cosmetic viewpoint, the emulsifier must also stabilise the
emulsion independently over a wide range of variables such as the pH, characteristics of
the oil phase, and the salt concentration (1).
In the case of oil-in-water emulsions (o/w emulsions) frequently anionic emulsifiers
of the form R-(CH2kCH3 (n = 12-20, R = carboxylate, sulphate, phosphate) are used.
The ability of many ionic emulsifiers to form aggregates in the water phase (also
denoted as an emulsifier gel network) is a further important stabilisation factor for o/w
emulsions due to a steric hindrance of droplet coalescence.
NMR spectroscopy methods statistically average over sample volumes of approx.
1 cm3 and are frequently applied to complex surfactant systems like colloids,
microemulsions and emulsions because non-invasive measurements make in situ
examinations possible (2-4). The spectra of the molecules reflect the chemical
environment of the nuclei, which enables a detection of the different states (interface,
aggregates ... ) of the emulsifier in the emulsion. Dynamic NMR measurements like
determination of self-diffusion coefficients provide information about molecular
dynamics and spatial dimensions of aggregates and cavities (e.g. droplet sizes) (5). The
375
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Inteiface Science, 375-382.
© 2002 Kluwer Academic Publishers.
376

resolution of proton and carbon spectra of emulsions is frequently reduced by


inhomogeneous line broadening of the signals arising from high sample viscosity.
Detection of the emulsifier is also complicated by overwhelming signals of other highly
concentrated constituents like water and oil compounds in the spectra.
By 31p NMR spectroscopy the hydrophilic phosphate group of the stearyl phosphate
emulsifier used in this study can be detected selectively (6). Some 31p NMR studies
examining the head group motion of the amphiphiles have already been performed on
phospholipid membranes or lecithin-stabilised dispersions (7-9). The study presented
here has been designed to characterise the molecular properties of the stearyl phosphate
emulsifier in two cosmetic o/w emulsions with slightly different compositions.
31p NMR high resolution spectra are measured and proton pulsed gradient spin echo
(pGSE) experiments for water self-diffusion measurements carried out. A comparison
of static high resolution and solid state CPMAS spectra provides information on solid
compounds.

2. Experimental Procedures

2.1. PRODUCTION OF O/W EMULSIONS

A representative composition of the examined cosmetic emulsions is listed in Table 1;


all ingredients are of technical purity. The oil phase is prepared with the emulsifier at
75°C so that the latter is completely dissolved. The water phase is prepared at the same
temperature and slowly added to the oil phase while gently stirring with a kitchen aid
mixer (40 Rlmin). The dispersion formed is then stirred for 5 min (60 Rlmin) before
homogenisation in a rotor-stator homogeniser (Zehner AG, Switzerland) at 2870 Rlmin.
Afterwards the emulsion is stirred again at 30-40 Rlmin until room temperature is
reached. Then the perfume is added and a second homogenisation step performed at RT.

TABLE 1: Representative composition of the examined cosmetic emulsions.

ingredients weight % description

stearyl phosphate 2.5 emulsifier


lipid mixture(alkyltriglycerides, -ethers) 9 oil
~l p~bene mixture 0.5 jlreservative
fragrance q.S. perfume
carbomer (polyacrylic acid) 0.1 stabiliser
sodium hydroxide (10 %) 1.3 neutraliser
glycerol (86 %) 3 moisturiser
water 83 .5 continuous phase

Emulsion A is stabilised by the emulsifier Hostaphat™ CS 120 alone, emulsion B by


HostaphatTM CC 100 (both from Clariant, Frankfurt, Germany) together with a cetyl
alcohol acting as a co-emulsifier. The concentrations of the emulsifiers can vary
slightly.
377

2.2. lH AND 31p NMR SPECTROSCOPY

All 31p high resolution NMR measurements are performed on Broker DPX 250 and
DRX 500 spectrometers operating at 31 p frequencies of 101.254 MHz and
202.456 MHz. The temperature is kept constant at 300 K for all NMR measurements
and varies less than ±O.2 K. 10 mm multinuclear probes with proton decoupling coils
are used without sample spinning. Proton multi pulse decoupling is applied with a
Waltz16 sequence. For spectral reference an external standard of phosphoric acid
(H3P04, 85 %) is set at 0 ppm.
lH NMR self-diffusion measurements are carried out according to the pulsed gradient
spin echo experiment ofE. O. Stejskal and J. E. Tanner (10). For the experiments, a lH
5 mm rnicroimaging probe with a shielded z-gradient is used. Sinusoidal gradients of
5 ms duration 8 with strengths varying from g = 0 to 1050 G/cm are applied during the
stimulated echo sequence in order to obtain 100 to 95 % signal decay. The gradient
strength is calibrated by the self-diffusion coefficients of water (2.3 x 10-9 m2 /s at 298 K)
and glycerol (2.15xlO-12 m2 /s at 298 K). 320 to 512 transients at about 60 different
gradient strengths are recorded to provide fits to the data with a mean uncertainty of
about 20 %. The diffusion time !'l. is set to 100 ms. The data are analysed by the theory
of gaussian diffusion leading to Equation 1 for the diffusion dependence of the signal
attenuation.

( 1)

The attenuation of the NMR signal in pulsed field gradient experiments induced by
molecular diffusion is described by the gyromagnetic ratio y, the gradient pulse width 8,
the variable gradient strength g, the diffusion time !'l. and the diffusion coefficient D.
Fits of Equation (1) to the experimental data are performed in a logarithmic-quadratic
plot with a home built C-script running on a SGI-UNIX workstation. In most cases of
water self-diffusion in emulsions the data need to be described by two or more diffusion
coefficients representing different water pools in the heterogeneous system. The
intensities 10 at g = 0 of the different diffusion terms allow a quantitative estimation of
free and bound water in olw emulsions.
The raw emulsifier and the emulsions have been studied by solid state NMR methods
to focus on crystalline parts of the emulsifier in the emulsions not detectable by high
resolution spectroscopy. Cross polarisation (CP) in combination with magic angle
sample spinning (MAS) enhances the 31 p sensitivity for solid samples due to
polarisation transfer from the protons during the contact time and sharpens the lines by
reducing dipole-dipole interactions. Measurements are carried out on a Broker DSX 500
spectrometer, operating at a 31p frequency of 202.5906 MHz, with a 2.5 mm
multinuclear MAS probe. Sample spinning rates are about 20 kHz for the emulsifier and
10 kHz for the emulsions. Contact times for cross polarisation transfer (CP) are between
5 and 10 ms (12). Undoubtedly, during the rotation emulsions are destroyed by
centrifugal forces even in the small rotor used here. Nevertheless, the method allows
detection of crystalline components in the emulsions by cross polarisation transfer.
378
3. Results & Discussion

Figure 1 shows the 31p spectrum of the emulsifier HostaphatTM CS 120 dissolved in fully
deuterated tetrahydrofuran (dg-THF). The characteristic signal splitting is due to the J-
coupling of the phosphorus nucleus with the OCH2-group of one or two fatty alcohols.
The phosphate content is due to the manufacturing process; no signal of a triester is
detectable.

phosphate

~
2 .6 2 .2 1 .8 1.4 1.0 0 .6 0 .2 -0 .2 -0 .6 ·1 .0 -1 .4 -1 .B -2 . 2
(ppm)
Figure 1. lip NMR spectrum of the stearyl phosphate emulsifier in ds-THF without
proton decoupling.

From these NMR spectra in combination with mass spectroscopy the following raw
material composition in weight% for the two different emulsifiers is obtained (Table 2):

TABLE 2: Composition of the two emulsifiers.

monoester diester phosphate fatty alcohol


(%) (%) (%) (%)
Hostaphat™ CS 120 33.8 38.3 .7 3.4
HostaphatTM CC 100 60.4 14.1 6.2 8.1

Noticeable is the higher concentration of the diester in HostaphatTM CS 120 compared


with HostaphatTMCC 100.
Figure 2 compares spectra of the emulsion A and the emulsifier in a MAS rotor. The
upper spectrum is taken with the rotor at rest. The signal at 0.8 ppm belongs to
inorganic phosphate from the raw material as well as from hydrolysis of the emulsifier
during the preparation process of the emulsion. The signal at 0 ppm is due to the
monoester. The spectrum in the middle is a CPMAS spectrum and hence emphasises
solid, not dissolved ingredients. Due to centrifugal forces the emulsion is destroyed
leading to changes of the chemical shifts, but clearly a third signal is detected. The
comparison with the CPMAS spectrum of the emulsifier as a raw material (bottom)
allows the conclusion that this signal is due to the diester. Obviously, emulsion A
contains the diester in a solid phase only.
379

static
emulsion with 1.5 %
Hostaphat'" CS 120

CPMAS emulsion with 1.5 %


Hostaphat'" CS 120

CPMAS Hostaphat'" CS 120


raw material

/0 8 -2 -4 -6 -8 -/0 -12 -14 -16 -18


wpm)

Figure 2. IIp spectrum of emulsion A in a MAS rotor at rest (top),


31p CPMAS spectrum of emulsion A (middle), 31p CPMAS spectrum of the emulsifier raw material (bottom).

Recrystailisation experiments done with the raw material show significant differences in
the solubility of mono- and diester.
The small angle X-ray scattering diagram of emulsion A shows a sharp peak due to a
characteristic repeating distance of 4.4 nm. This peak is comparable to that of the raw
material HostaphatTM CS 120 and points out that there is a highly ordered phase in the
emulsion with a characteristic extension of 4.4 nm (13).
Due to its much lower concentration the CPMAS spectrum of emulsion B does not
show any diester signal nor do small-angle X-ray data show signals of crystalline
components. Instead of this, the long-time registered 31 P high resolution spectrum
shows besides the signal of inorganic phosphate a broad, asymmetric background
(Figure 3, upper trace).
Because this effect is unchanged in the emulsifier gel, that is the emulsifier swollen in
water, (Figure 3, lower trace) it is due to the emulsifier molecules not located at the oil-
water interface. These molecules form aggregates leading to a preferred orientation of
their axes to the magnetic field. Due to molecular rotations about this preferred
direction (mainly about the C-O-P bonds of the alkyl phosphate), the shielding tensor is
no longer averaged along this axis and the chemical shift is orientation dependent. This
may be quite similar to phospholipid membranes and vesicles as found earlier by other
authors (7-9).
380

(1

emulsien

, emulsifier gel

(ppm)

Figure 3. Long-time registered 31p NMR spectrum of emulsion A (top) and the emulsifier gel without the oil
phase (bottom),

Figure 4 compares the results of the water self-diffusion measurements on the two
emulsions.

Emulsion A: Emulsion B:
Hostaphat'" CS120 Hosta hat'MCC100
100 1.00E·08 ,.---....:...---r----"""T 100

90 90
~ 1 .00 1,OOE·09
80 80
~
.s 70 oS 70
~ 100E-1 'E 1.00E·10

.
Q)
'(j (j
60 "0
;;:: 60 "0
~o ID
n Q)

u 1,ooE-11
c
50 ~ S 1,OOE.11 50 ~
o OJ c ::J
,Q Ci
'iii
:E
40 '"
ID III 40 <g
~ 1.ooE-1 ~ 1,OOE·12
30
30
Gi
III
~III
20 20
1.00E·13
10 10

o 1. 00 E· 14 o

Figure 4. Comparison of the results of water se lf-diffusion measurements on the two emulsions. Left scal es:
self-diffusion coefficients, right scales: percentage oft h.: component .

In both emulsions the signal decay must be fitted by two exponential functions resulting
in two diffusion coefficients. The larger coefficient is ascribed to the relatively mobile
381

molecules of free water (self-diffusion coefficient of pure water 2.3xlO-9 m2/s), the
other one being about four magnitudes smaller to the immobilised water. The two
emulsions differ in the relative percentage of the two components: In emulsion A there
is about 100 % free water, whereas in emulsion B immobilised and free water are of
almost equal quantity.
In emulsion A, the monoester of phosphoric acid is mainly concentrated at the oil-
water interface, the diester is insoluble and forms crystals, maybe together with the
excess monoester. No emulsifier gel network is formed, the self-diffusion of the water
molecules is only slightly hindered.
In emulsion B, the co-emulsifier induces the development of an emulsifier gel
network in the water phase. Compared with emulsion A, the diester is much lower
concentrated and detectable neither by high resolution nor by CPMAS methods. The
water molecules are partly immobilised in the gel network, partly form pools with a
self-diffusion coefficient comparable to that of pure water. The molecular motion of the
emulsifier molecules in the network is fast enough for high resolution detection, but
non-isotropic leading to a broadline resonance.
The aim of this work is to show that it is possible to get structural insight by NMR
methods into systems like emulsions in between the solid and the fluid state by making
the most use of the main advantages of NMR, that are only minor preparation
procedures, statistical averaging over macroscopical sample volumes and a low risk of
artefacts in most cases due to non-invasive measurements.

4. Acknowledgement

The authors are grateful to Dr. P. Arlt for introduction to the diffusion measurement
procedure and to Dr. M. Bertmer for introduction to the DSX 500 at the MARC, RWTH
Aachen. We also would like to acknowledge the co-operation of Mr. A. Bleckmann and
Mr. R. Kropke in emulsion production.

5. References

1. Lagaly, G., Schulz, O. and Zimehl, R. (1997) Dispersionen und Emulsionen, Steinkopff, Dannstadt,
p.233 .
2. Delpuech, J-J. (1994) Dynamics ofSolutions and Fluid Mixtures by NMR, Wiley, New York, p.345.
3. Griffiths, L., Horton, R., Parker, I., and Rowe, R.C. (1992) NMR Pulsed Field Gradient Characterization
of the Behavior of Water in Gels and Emulsions containing Cetostearyl Alcohol and Cetrimide, J.
Colloid Interface Sci. 154, 238.
4. Goetz, RJ., Khan, A, and EI-Aasser, M.S. (1990) IT PGSE NMR Investigation of the Supramolecular
Structure of a Dilute Gel Phase, J. Colloid Interface Sci. 137,395.
5. Soderman, 0 ., Lonnqvist, I., and Balinov, B. (1992) Emulsions - A Fundamental and Practical
Approach: NMR Self-DiffUSion Studies of Emulsion Systems, Droplet Sizes and Microstructure of the
Continuous Phase, Kluver Academic Publishers, Netherlands, p.239.
6. QuiD, L.D., and Verkade, 1.G. (1994) Phosphorus-3) NMR Spectral Properties in Compound
Characterization and Structural Analysis, VCH-Verlag, New York, p.283.
7. Yoshinori, M., Kazuhiro, C., and Masahiro, T. (1992) Effect of the Oxidation of Free Fatty Acids on the
Interfacial Adsorptivity of Lysophosphatidylcholine/Free Fatty Acid/Ovalbumin Complexes, Biosci.
Biotech. Biochem. 11,1814.
382
8. Arnold, K., Gawrisch, K., and Volke, F. (1979) 31p NMR Investigations of Phospholipids: l. Dipolar
Interactions and the 31p NMR Lineshape of oriented Phospholipid/Water Dispersions, Stud. Biophys. 75,
189.
9. Kohler, S.1., and Klein, M.P. (1976) 31p Nuclear Magnetic Resonance Chemical Shielding Tensors of
Phosphoryl ethanolamine, Lecithin, and Related Compounds: Application to Head-Group Motion in
Model Membranes, Biochemistry 5, 967.
10. Stejskal, E.O., and Tanner, 1.E. (1965) Spin Diffusion Measurements: Spin Echoes in the Presence of a
Time-Dependent Field Gradient, J. Chern. Phys. 42, 288.
II. Karger, 1., Pfeifer, H., and Heink, W. (1988) Principles and Application of Self-Diffusion Measurements
by Nuclear Magnetic Resonance, Adv. Magn. Res. 12, I.
12. Pines, A., Gibby, M.G., and Waugh, 1.S. (1972) Proton-Enhanced Nuclear Induction Spectroscopy. A
Method for High Resolution NMR of Dilute Spins in Solids, J. Chern. Phys. 56, 1776.
13. Bartos, S. (2001) Thesis, Universitiit Hamburg
WATER MAGNETIC RELAXATION

IN SUPERPARAMAGNETIC COLLOID SUSPENSIONS:

THE EFFECT OF AGGLOMERATION

A. ROCH#, F. MOINY*, R.N. MULLER#, P. GILLIS*

# Department of Organic Chemistry and NMR Laboratory,


* Biological Physics Department,
University of Mons-Hainaut, B 7000 Mons Belgium

Superparamagnetic (SPM) nanoparticles in aqueous suspensions are known to shorten


the nuclear magnetic relaxation of water protons. For transverse relaxation, this effect is
enhanced when agglomeration of elementary SPM cores occurs. This feature is
explained by considering the agglomerate as a large magnetized sphere, which at high
field contributes strongly to the secular part of the transverse relaxivity. Longitudinal
relaxation is modeled by analyzing diffusion through the permeable coating of the
agglomerate as a virtual exchange between normal decaying modes and the bulk water.
This model is used to interpret the flattening of the Nuclear Magn~tic Relaxation
Dispersion profiles during a chemically induced agglomeration of ferrite particles.

1. Introduction

Superparamagnetic colloids improve the contrast in MRI (Magnetic Resonance


Imaging). The particles currently used for this purpose are composed of very small
individual ferrite crystals about 10 nm in diameter, stabilized by a polysaccharide or
polyelectrolyte coating [1, 2]. These particles are mainly used to shorten the transversal
relaxation time. Their performances as contrast agent depend on parameters like size,
clustering, ... These parameters are generally not easily controlled from the synthesis of
the particles; a theoretical knowledge of the relaxation properties is thus in tum useful in
order to specify morphological and physical parameters of the particles, especially
through the interpretation of so-called Nuclear Magnetic Relaxation Dispersion profiles
(NMRD), i.e. the dependence of the water magnetic relaxation rate on the static
magnetic field. Results are sometimes expressed in terms ofrelaxivity, i.e. the water
proton relaxation rate enhancement per millimole of iron per liter of solution.
383
1. Fraissard and o. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 383-392.
© 2002 Kluwer Academic Publishers.
384

Superparamagnetic contrast agents are classified following their size: particles


containing only one ferrite crystal are called USPIO (Ultra Small Particles of Iron
Oxide), while relatively larger particles containing several magnetic crystals within the
same flake of permeable coating are called SPIO (Small Particles ofIron Oxide) [3].
For USPIO, a theoretical approach assuming a uniform distribution of the
magnetic crystals within the solvent is rather realistic and allows the calculation of the
nuclear magnetic relaxation rate [4-6]. This assumption is no longer valid for SPIO,
which contain several ferrite crystals per particle. The present work is aimed at
accounting for the agglomerated structure of SPIO, starting from the theory valid for
USPIO.

2. Theory

Effects arising from the aggregation of magnetic grains are twofold: on the one
hand, those related to the global structure of the cluster and to the magnetic field
distribution around it, and on the other hand, those limited to the inner part of the
aggregate. While the former ones essentially affect TJ and TJ * [5, 7], the latter ones
govern T,.

2.1. GLOBAL EFFECT OF THE AGGLOMERATED STRUCTURE

The cluster itself may be considered as a large magnetized sphere whose total
magnetic moment increases according to the Langevin's law. The global magnetization
of the agglomerate is always aligned on the external field. Its effect on relaxation is
characterized by a long correlation time, because of its large size, so that its contribution
mainly affects the secular term of the relaxation rate. This contribution is given by the
outer sphere diffusion theory, provided the motional averaging condition is fulfilled:
LIm 'Da < 1, where LIm is the difference in angular frequency between the local field
experienced by a proton at the cluster surface and in the bulk; 'D" is the translational
diffusion time around the cluster ('D" = R/ID, R" being the cluster radius and D the
water diffusion coefficient) [8]. This secular contribution explains the increase of liTe at
high field [5, 7].
On the contrary, the longitudinal relaxation rate, which does not contain any
'0 'e
secular term, goes to zero for frequencies larger than 1I where is the correlation
time of the fluctuation responsible for relaxation. Diffusion around the agglomerated
particle is characterized by a correlation time too long to contribute significantly to the
longitudinal magnetization recovery at usual imaging fields. Furthermore, at low field,
i.e. for frequencies smaller than lire, the global magnetic moment of the agglomerate is
too weak to allow for a significant contribution to the longitudinal relaxation rate. For
small agglomerates satisfying the motional averaging condition (for magnetite, the
diameter must be smaller than 200 nm), the transverse and longitudinal relaxation rates
arising from diffusion around the agglomerate are thus approximated by :
385

l/h,- 0 (1)

liT]" - (16n/135000) [n f.JS,JVgL(X)]2 NAC" /(R"D), (2)

where f.Jsp is the magnetic moment of a crystal, Ng is the number of crystals in an


n
agglomerated particle, is the gyromagnetic ratio of proton, NA is Avogadro's number,
C" is the agglomerate concentration, and x = f.Js,JVg BJ(kT), Bo being the static field, k
the Boltzmann constant and T the temperature. The Langevin function is defined as
follows:

L(x) = coth(x) - llx (3)

2.2. INSIDE THE CLUSTER

The assumption of a uniform distribution of the crystals within the whole


suspension is obviously abandoned, and is replaced by the assumption of a uniform
distribution of the elementary crystals within a permeable coating flake, which is the
agglomerated particle.
We will rest on a formal analogy between our diffusion problem and a chemical
exchange situation. In a system where water fastly exchanges between a bulk fraction,
containing the majority of the solvent molecules, and an ensemble of bound fractions,
each characterized by a chemical shift and a fast intrinsic relaxation, the probability that
a molecule attached to site B, one of the binding sites, at t = 0, will still be attached to
the same site at time t, decays exponentially with t. The longitudinal relaxation rate
calculated in such a situation is [9] :

liTI = lITh+"
I ~
In n' (4)
n Tn +~
where T/ is the relaxation time in the bulk fraction,/" is the water fraction occupying
site n (f" « 1), Tn is the residence time on site n, and T/ is the relaxation time on site n.

If W( r, t) is the density of probability for a water proton to be at point r at time t,


the evolution of this probability is governed by the diffusion equation:

8W( r , t)/8t = D f1W(r , t). (5)

r = 0 corresponds to the center of the spherical agglomerate. At t = 0, the water


molecule is supposed to be somewhere within the agglomerate, so that:

W(r, 0) =0 if I r I > R"


W( r ,0) = liV if I r I 5. K, (6)

where V is the volume of the aggregate, and R" its radius.


386

The space where the molecules are allowed to diffuse is much larger than the
agglomerate itself. If we consider this space as a larger sphere of radius Re , in order to
avoid breaking the spherical symmetry of the problem (Re is then half the mean distance
between two agglomerates), the solution ofEq.(5), starting from the initial conditions
given in Eq.(6), may be written as:

3B. ,
W(r,t) =~
~ 2 n 2 sm(nnr / Re)e- n Il
'DIIR1
e (7)
n=1 2n7r Ra r
where Bn = sinan/an - cosan , with an = ntrR,/Re . Eq.(7) may be read as if at time t = 0,
each decaying mode appearing in the solution of the diffusion equation, with a decay
time Tn = R//(n2~D), is a virtual water fraction with a relative population (excluding the
water outside the agglomerate) given by :

P =~ = (B n / n )2 (8)
n <Xl <Xl

LPI L(B1 / /)2


1=1 1=1

leaving site n at a rate equal to 11 Tn' The factor Pn in Eq.(8) was obtained by integrating
W(r,O) over a sphere of radius Ra .
In order to transfer these results into Eq.(4), we need noting thatJ", the relative
amount of each "fraction", is equal to tPm t,
where is the volume fraction occupied by
the agglomerates. T/ , the intrinsic relaxation time in fraction n, is obtained from the
theory valid for USPIO (5), and is expressed assuming a relaxivity r, (IIT/ = Cg r/,
where Cg is the iron molar concentration within the aggregate). Given that C, the total
iron concentration, is equal to t,Cg, we finally get:

11 ~ =11 ~b + CI n=1
Pn
11 rIg + Cg Tn
, (9)

where T/ is the relaxation time in the bulk fraction.


Fig. 1 displays the population fractions in each decaying mode, i.e. our virtual
"binding sites", as functions of the decay time of the same mode. Each point on the
graph corresponds to one value of n.
Fig. 2 is the theoretical prediction from Eq.(9). The USPIO relaxivity r/ depends
on several parameters characterizing the elementary grains embedded in the coating,
namely their radius, their saturation magnetization (Ms), and the Neel relaxation time,
typical of superparamagnetism (TN)' These parameters were adjusted in order to fit the
experimental curve corresponding to an agglomerate of 50 run in radius shown in Figure
5. The best values were 8.2 run for the grain radius, TN = 1.8 ns, Ms = 49 ueml(g Fe), and
a volume fraction occupied by the ferrite crystals inside the agglomerate of 0.02.
387

0.14
___ R. = 100 nm
0.12
c...
c 0.10
- - Ra =400nm
r'\
0
+=0 + \
co 0.08 I ~
<.)

J=
c 0.06 r \

,
0
+=0 ~
co 0.04
"5 \
a.
0 0.02
c... "'.
0.00

0.001 0.01 0.1 1 10 100 1000


Decay time t (f.!s)

Figure 1. Display of the relative fractions associated to the diffusion modes (Eq.(8)), expressed as a function
of the decay time Tn , for D = 3.5 X 10.9 m2/s and R"IRe = 0.1

60 - , - - -- - - - - - - - - - -- - - - - ,
.....
~
Ra = 50 nm
~ 50 Ra = 100 nm
.....E Ra = 150 nm
:e 40 Ra = 250 nm
Z' _ .. Ra = 350 nm
'>
.~ 30
/--"
'" '" "-
-
~
/
~ 20
./ / ......
'6
.a ---- ."
.... .... . . . ..
.' "-
\.'
'g> 10
."

o ~: ~~:::. : = : .-=~~~.~
..J

o ~-----_,-------,-------,-----~
0.01 0.1 1 10 100
Larmor frequency (MHz)
Figure 2. Theoretical longitudinal NMRD profiles from Eq .(9). The parameters characterizing the
elementary grains are given in the text.
388
3. Material and methods

Transverse relaxation rates were measured on suspensions of dextran coated


magnetite particles called AMI25, delivered by Advanced Magnetics (Cambridge, MA,
USA), by means of three low field spectrometers working respectively at Larmor proton
frequencies of 10,20 and 40 MHz (Minispec Bruker PCIO, PC20 and PC40).
The effect of agglomeration on longitudinal relaxation times was studied on
suspensions of manganese zinc ferrites coated with anionic polyacrylate, synthesized by
Dr K. Korman, BASF. A controlled flocculation of these coated manganese zinc ferrites
was induced by adding Ca++ counterions to the suspension. The agglomeration kinetics
was followed by measurement of the size by photocorrelation spectroscopy (PCS, light
scattering) (Brookhaven Instrument Corporation BI-160, Holtsville, New-York, U.S.A).
Water proton longitudinal relaxation rates were measured at different stages of
agglomeration using an IBM research field cycling relaxometer (IBM Research,
Yorktown Heights, New York, USA). This relaxometer is designed to provide NMRD
profiles, i.e. measurements of the solvent proton longitudinal relaxation rate as a
function of the static magnetic field.

4. Results and discussion

4.1 . TRANSVERSE RELAXATION

Figure 3 clearly shows that the measured relaxivity is much larger than expected
from the sole contribution arising from the inner part of the aggregate (dotted line).
Adding a contribution due to diffusion around the whole aggregate allows reproducing
the steep increase of relaxivity at high field, although the saturation of this increase
predicted by the theory is not confirmed by the experimental measurements. The volume
ratio of ferrite in the agglomerate was adjusted to fit the experimental results.

4.2: LONGITUDINAL RELAXATION

Because of the polyacrylate coating, the particles are negatively charged, which
prevents them to approach one another. The Ca ++ divalent cations neutralize the
carboxyl anions, and the negative electric charge on the coating decreases down to the
vanishing of the repulsive forces responsible for the stability of the solution. Figure 4
shows the agglomeration kinetics of the zinc ferrite coated with polyacrylate induced by
addition of Ca++ counterions to the suspension.
389

200
......,.-.
~
....,E 150
!!!-
~ ... _----.J
'5
'x
III 100
~
Q)
....
If)
Q) 50
>
If)
c::
~
~
0
0.01 0.1 1 10 100
Larmor frequency (MHz)

Figure 3. Transverse relaxivity of AMI25 samples. The triangles are experimental data. The dotted line is the
prediction from USPIO theory (without accounting for agglomeration), with the grain parameters obtained
from the fitting to Eq.(9) of the longitudinal relaxivity for RI/ = 120 nm (from PCS measurements) in Fig. 5
(grain radius: 6.4 nm, specific magnetization: 41 emu/(g Fe) and Neel relaxation time: 1.8 ns). The solid
line is obtained by adding the contribution given in Eq.(2), with a volumic ratio of ferrite in the agglomerate
of 0.032.

700

600


500

E
.s 400
'"
;:;J
'0
<1l 300
0::

200


100

0
0.0 0.5 1.0 1.5 2.0
Time (hours)

Figure 4. Evolution of the hydrodynamic size, as measured by PCS, during the Ca++ induced agglomeration.
390

NMRD profiles also evolve with clustering (see Figure 5): the profiles become
flatter on increasing agglomeration. Comparison of the clustering-dependent
experimental curves with the theoretical ones (Figure 2) shows the adequacy of our
modeling with the physical and morphological properties of clustered
superparamagnetic colloids. However, one can remark that the flattening of the
longitudinal relaxivity profile is more pronounced than predicted: for instance, going
from R" = 50 nm to R" = 100 nm reduces the amplitude of the relaxivity peak by a factor
of2 on Fig. 5 (experimental results), but only by a factor of 1.6 on Fig. 2 (theoretical
predictions from Eq.(9». This discrepancy is likely to be due to a questionable implicit
theoretical assumption - Cg , the concentration of the crystals within the agglomerates,
was supposed to be invariant during the agglomeration process, and the same value was
used to draw all the curves presented in Fig. 2. However, the effect of the Ca ++ ions, i.e.
decreasing the repulsive forces between the crystals, not only increases the size of the
agglomerates, but also allows the crystals to get closer to each other, which in tum
increases the concentration in the agglomerate. Now, Eq.(9) clearly shows that
increasing Cg , all other variables being unchanged, will decrease the relaxation rate.
The success of our modeling is somehow surprising because the assumption of a
uniform distribution of ferrite crystals is far from realistic: the structures known to
minimize the magnetic energy of associated ferrite crystals are shaped as rings or open
loops [10]. The good agreement between the predictions of our simple model and the
experimental results is therefore an indication that the distribution of the elementary
grains within the aggregates does not playa critical role regarding relaxation.

60
..-I
~

• Ra = 50 nm

• • •
::lE 50 Ra =80nm
E
..-I Ra= 120 nm

A
.e40
"
Ra =210nm

Z:-
.s: • Ra = 420 nm
.~ 30
• •
~
• • • •
~ 20 , . • A
"
A A
A
u I
! "
•" • • " ••
A

~10
0
-l

T

0
0.01 0.1 1 10
Larmor frequency (MHz)

Figure 5. Evolution of the experimental NMRD profiles during agglomeration (T = 37°C). The sizes
indicated on the figure are determined according to the mean measurement time, using the data of Fig. 4.
391

5. Conclusion

This work unambiguously demonstrates that accounting for agglomeration is


mandatory for explaining the relaxation properties of SPM colloids. Published
theoretical models [4-6] do not incorporate such a feature : they only apply to non-
agglomerated suspensions, the so-called USPIO, where each flake does not generally
contain more than one ferrite crystal. Agglomeration modifies longitudinal as well as
transverse relaxation, but in very different ways.
Considering the entire agglomerate as a single magnetized entity accounts for the
observed increase of the transverse relaxation rate at high field. However, the classical
outer-sphere theory only provides a very crude and semi quantitative approximation of
its value. A more accurate calculation of this contribution probably requires further
recourse to numerical methods described elsewhere [7, 11].
Longitudinal relaxation undergoes quite different modifications under
agglomeration: NMRD profiles become flatter, because of an important decrease of the
relaxation rates. This effect is interpreted as the result of a virtual exchange between the
water inside the agglomerates, rapidly relaxed, and the waters within the bulk, relaxing
much slower. The time spent by the water molecules within the agglomerate is not
simply given by the usual diffusion time TDll = R,/ID, it is instead deduced from the exact
solution of the diffusion equation in a finite spherical space, producing an ensemble of
decaying modes each characterized by its own decay time. These modes play the same
role as binding sites with a short relaxation time, and with a residence time
corresponding to the decay time of the diffusion modes.
This model is shown to reproduce the decrease of the longitudinal relaxivity
observed during a chemically induced agglomeration process. For small size aggregates
(Ril < 50 om) in pure water, this phenomenon is often masked behind an underestimation
of the saturation magnetization, fitted to a value lower than the magnetometric direct
measurement.

6. References
Magin, R.L., Bacic, G., Niesman, M.R., Alameda, J.c. , Wright, S.M., and
Swartz, H.M. (1991) Dextran Magnetite as a Liver Contrast Agent, Magn. Reson. Med. 20, 1-16.
2 Mail', G., and Steck., W. U.S Patent Application NO 4,810,401 (7/03/1989).
3 Muller, R.N., Roch, A., Colet, J.M., Ouakssim, A., and Gillis, P. Particulate Magnetic Contrast Agents,
in A.E. Merbach and E. Toth (eds). The Chemistry o/Contrast Agents in Medical Magnetic
Resonance Imaging, John Wiley & Sons, Ltd, Chichester, pp. 417 -435.
4 Roch, A. , and Muller, R.N. (1992) Longitudinal Relaxation of Water Protons in Colloidal Suspensions
of Superparamagnetic Crystals, Proceedings 0/ the 11th Annual Meeting of the Society of Magnetic
Resonance in Medicine, Works in Progress, p. 1447.
5 Roch, A., Muller, R.N., and Gillis, P. (1999) Theory of Proton Relaxation Induced by
Superparamagnetic Particles, J. Chem. Phys. 110,5403-5411.
6 Roch, A., Muller, R.N., and Gillis, P. (2001) Water Relaxation by SPM Particles: Neglecting the
Magnetic Anisotropy? A Caveat, J. Magn. Reson. Imaging 14, 94-96.
7 Moiny, F., Gillis, P., Roch, A., and Muller, R.N. (1992) Transverse Relaxation ofSuperparamagnetic
Contrast Agents: A Numerical Analysis, Proceedings of the 11th Annual Meeting of the Society of
Magnetic Resonance in Medicine, Works in Progress, p. 143\.
392

8 Ayant, Y., Belorizky, E., Alizon, J., and Gallice, 1. (1975) Calcul des densites spectrales resultant d'un
mouvement ah:atoire de translation en relaxation par interaction dipolaire magnetique dans les liquides,
J. Phys. 36.991-1004.
9 Zimmerman, J.R.. and Brittin, W.E. (1957) Nuclear Magnetic Resonance Studies in Multiple Phase
Systems: Lifetime of a Water Molecule in an Adsorbing Phase on Silica Gel, J. Chern. Phys. 61, 1328-
1333.
10 Chantrell. R.W.. Bradbury, A.• Popplewell. J., and Charles, S.W. (1982) Agglomerate Formation in a
Magnetic Fluid. J. Appl. Phys. 53.2742-2744.
II Weisskoff. R.M., Zuo, C.S., Boxerman, J.L.. and Rosen, B.R. (1994) Microscopic Susceptibility
Variation and Transverse Relaxation: Theory and Experiment, Magn. Reson. Med. 31,601-610.
A Stray Field Imaging Study of the Drying Process of Precasting Materials used
in a Steel Making Converter

Koji Saito I)', Yoshitoshi Saito I), Peter I.McDonald 2) , John Godward 2)

1) Nippon Steel Corporation, Advanced Technology Research Lab.


20-1 Shintomi Futtsu City, Chiba, 293-8511, JAPAN
2) Department of Physics, University of Surrey

ABSTRACT
Optimization of the drying process conditions in a steel-making converter in a steel
work is very critical since the process is off-line and time-consuming. However, it is
important to optimize these conditions (temperature, surfactants, etc.), as steam
explosion can readily occur with insufficient drying time. To help understand this
process, we have demonstrated that we can monitor the drying of real refractory mortar
using Stray Field Imaging. We chose this method because of the possibility of detecting
short Tz components. This paper shows the effect on the drying rate of varying water
content in different materials. In particular, we find that the free water loss rate is
relatively independent of water content. However the bound water loss rate is more
affected. It is clear that imaging gives useful information for optimizing drying
conditions. Using this data, we can adjust and optimize the drying process and time in
steel works.

I.INTRODUCTION
NMR imaging is a powerful tool that is widely and successfully used in medical
applications. Its potential for solving materials science problems has been recognized
only recently. NMR imaging of solids is still a challenge despite significant
achievements in this field [I]. NMR microscopy allows one to study macroscopic
transport of liquids in materials by monitoring transformation of spatial distribution of

393
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 39~02.
© 2002 Kluwer Academic Publishers.
394

liquid content in time [2]. Another important process that involves mass transport in
materials is drying [3-13]. The conventional approach uses evaluation of the total
amount of liquid lost versus time (drying curves), e.g., by sample weighing, while its
distribution within the porous object is determined using a gravimetric method based on
sample sectioning. Determination of liquid distribution in porous solids nondestIUctively
and in real time by means of various imaging techniques has been introduced only
recently. Nevertheless, the drying of a variety of materials has been studied by NMR
imaging, including that of wood. [13]
It is well known that the presence of water and water transport play key roles in
determining the pore structure and long term durability of cementitious mortar materials
in the case of drying. In real life, some problems, ex. hydrogen-explosion, can arise
because of the poor drying conditions. Up to now, however, a severely limiting factor

has been that resonance line widths in cements are very large. Consequently, only the

most mobile water in the largest capillaries, cracks, and voids has been observed. Only
more recently have STRAFI imaging techniques [14] been examined. Surface relaxivity
has been exploited by McDonald et al [15] using the complementary technique of
magnetic resonance relaxometry. These authors show that the initial IH spin-spin
relaxation time (T2) of water in a dried 0.5 water to cement (w/c) ratio mortar paste
depends on the water content, indicating that the rapid exchange model of relaxation can
be applied to determine the pore surface area to volume ratio. This paper presents the
results of a STRAFI imaging study of the hydration and drying of mortar pastes and of
subsequent water transport in dried material. This process may also be performed in
underwater environments, and such cases were also studied. The objective is to correlate
the bulk relaxation analysis of samples sealed during cure, which are assumed to be
spatially uniform, with the stIUcture of the paste based on traditional understanding.
Quantitative STRAFI imaging is then used to observe deviations from this model for

samples dried in the open or under water where spatial heterogeneity is expected. The

knowledge of the mechanism of the drying of these refractory mortars will allow to
optimize drying in real steel making process.

2.EXPERIMENTAL
STRAFI profiles were obtained by mov"ing the sample vertically through the sensitive
layer that defined by a combination of applied RF frequency and position within a field
gradient. In this work a 9.4 T superconducting solenoid magnet is used, producing a field
395

gradient of 58 T m· 1 in the fringe field. The proton resonant frequency at the point of
optimum gradient is around 235 MHz. The field gradient produces a change in resonant
frequency of around 2.3 MHz mm· l . A multiple quadrature echo train was recorded.
Each profile consists of 200 separate NMR echo decay trains recorded 60 microns apart
along the sample length. The total profile (sample) length is thus 12.0 mm. Each train
consists of 32 echoes with a pulse gap, echo time (2",) of 35 microseconds. T2
information was extracted from the echo decay trains by fitting the equation

M = Mo exp(-2itIT2)+c (c = constant, i = echo number) (1)

to the decays as a function of time and position.


Typical refractory mortar samples are prepared as Sample I(Sample3+surfactants) and
Sample 3 (AI20): 92%, MgO: 7%, CaO: 1%). For most of the studies, mortar samples
were cast in glass pots and NMR tubes from real industrial materials (Kurosaki Yougyou
Ltd.) using a food processor to mix the water and mortar powder and a vibration table to
remove incorporated air. The tubes and pots were sealed. In the case of samples cured
under water, additional water was carefully added to the pots after vibration and before
closure. For the measurements made during curing, the samples were examined in-situ
(50 0c). The sample weights were monitored at all stages of the experiment.

3.RESULTS AND DISCUSSION


F i~ur.: I shows typical STRAFI profile of refractory mortar Sample I.
I.

.
--....:...---....:...--------,- - - - -..
F.... ~!)ltIII1QII

Figure 1. Representative STRAFI echo train profiles recorded at 50 ·C, every 40 min, during the drying of
reti'actory mOl1ar.

The STRAFI experiments produced a large number of water content profiles. The
details of the front shape are distorted by the finite time required to record the profile.
These are better determined in a time of fljght experiment in which the signal intensity is
recorded at a fixed location as the front passes. Because data acquisition is considerably
more rapid, there is less distortion. The analysis discussed below uses a combination of
396

time of flight and profile data. Figure 2 shows typical results of STRAFI experiments for
the capillary water (free water) of Sample 1. The drying of refractory mortar proceeds
from the surface, consequently close to the surface the free water ratio decreases with
proximity to the surface. However, behind this initial region the ratio is maintained
constant throughout the remaining sample as water diffuses more rapidly from the lower
regions. In order to test whether or not the observed STRAFI profiles are due to Fickian
diffusion, we plot the diffusing front position as a function of 1'12, where t is the time. In
the case of Fickian diffusion, such a plot will be linear so long as the front does not
approach the far end of the sample. We define the front position as the position where
the front reaches half the maximum concentration. The front position is measured
relative to the original sample surface whose position is established from a profile
measured before the drying.
100
0-1 mill
..
4>-

c~ 90 -0- l -2 ml1l
0'
·z
- 2-3mm
.: 80 ........ 3 -4mm
0:: . ...... 4-5mm

....:
I-
~ - 5-6 mm
70 - 6-7 mm
~ - 7-80101
~ 60 -- 8-901m
~ -b- 9- 10 mm
t..
- 10 - llm01
50
-+- 11-12m01

40
0 100 200 300 400 500 600
Drying time (min.)

Figure 2 .The typical results of STRAFI experiments for the capillary water (free wa ter) of Sample I

Figure 3 and 4 show the 12 positions from surface froot as a function of tt /2 for a
selection of typical drying experiments results for Sample 3 and Sample I, respectively.
The data in the case of capillary water are all well approximated with linear fits,
indicating that Fickian diffusion is taking place. In addition the slope of all positions are
almost identical for Samplel, indicating that the drying is homogeneous. On the other
hand, the slope at the bottom position in the case of Sample 3 is different from the
others. We can obtain a "characteristic"diffusion coefficient D, from these plots since we
know that the characteristic diffusion length is given by (Dav t)112 which can be equated
to the front position. Day is simply the slope of the front position versus the ttl2 plot. This
397

measure will be used frequently in the subsequent discussion since it is a robust and
straightforward way of parametrizing the profile position data and is related to the
magnitude of the mutual diffusion coefficient.

.... 0· 111101
..,. 1· 20101
~
0
80 ..... 2·3mm
..... 3-4mm
~
-- 4 -50101
c2 60 -- S-Gm01
<'I
I: ..... 6-701111
.2..0 .. 7-8mm
[/}
I:
40
-- 8-9m 01
.....
0

e .... 9-1 0 01111


..... 10-1101111
0..
20
~ 11 - 12111111
Av cr'~g c

0
0 5 10 IS 20 25
Drying time 112 (min.)1 /2
Flgole J The 12 iXl:;itlons from sur face froll! as a funclion of rOOI I for a SClcclion of tYPlc,,1 cloyOl1&

e.xpcromenls III the case of Sample 1

120

--- 100
~
- O- II1Hll
"'-'
CJ
- 1-211101
-; 80 - 2·3 mm
~ .... 3-4111111
-; ..... 4-50101
I:
b.ll
60 ..... 5·6 11101
Ci5 -- 6-7111111
c::
0 40 - 7-8 mm

...
(5 - 8-9mm
.... 9·I Oml1l
0..
20 - 10-1 111101
-0- 11 . 12111 111
- ,\ ve l ag e
0
0 5 10 15 20 25 30 35
(Drying timc)l/2 (min 1/2.)
(· .gurc 4 The 12 pOsilions Irolll su rface fro nl as a function of rOOl I for iI «Iccilon of lypical clrYll1g

experiments in the case of Sample 3

The average diffusion coefficient of free water for the drying of Sample 3 at 50°C was
found to be 8.7xIO-s cm 2 /s. For instance, moisture diffusivity in brick and clay is in the
398

range 10-5_10- 3 cm2/s for various moisture contents in the materials, while the self-
diffusion coefficient for water is only 2.1x 10-5 cm2/s . In the case of Sample 1, the
average diffusion coefficient for drying at 50°C was found to be 9.7xlO-5 cm 2 Is.
Although total drying time was quite different between Sample 3 (I 080min.) and Sample
1(480min.), it is surprising that the average diffusion coefficient of free water is almost
the same between both samples (see Figure 5).

100 i
100
---.
~
80 ........,
0

2:!
~

o 0::
_ sample J
60 en
en
0 __ sample 3
....J
80 __ sample 1
~

40 t::
~ samp le 3
.':F
(/)
t:
70
20 .....0
0

l..
Q..,

60 o
o 5 10 15 20 25 30 35
(Drying Time )112 (min.)112

Figure 5. Typical drying pattern for both samples

On the other hand, what about bound water behavior? Stray-field imaging is a
very powerful tool for monitoring all capillary water (long T2, component) and bound
water (short T 2, component). It is found that the recorded quadrature echoes, which in
the case of combined water reflect a degree of line narrowing and in the case of mobile
water are strongly diffusion attenuated, can be well represented by two-component
exponential decays, with relative amplitudes of the components corresponding to the
proportions of capillary water and chemically combined water. Figure 6 and 7 show
profiles of bound water for Sample 1 and 3, respectively: Water transport in mortar
pastes with a fine pore structure is generally considered to be dominated by capillarity.
Following earlier work on unsaturated flow theory in soils, Hall has described how water
n·ansport in cementitious materials can be modelled as an effective Fickian diffusion
process in which the concentration-dependent hydraulic diffusion coefficient is defined
399

as the product of the hydraulic conductivity, K, and the derivative of the capillary
potential, 'I', with respect to water concentration, c. Hence:

Dc/dt=d(D( c)dc/dx)/dx, D(c)=K(c)d'I'/dc (2)

where x and t are position and time variables, respectively. A feature of generalized
Fickian diffusion is that the water front should advance as tin. We note that this is the
case in the experiments reported here .
......
..!..- 12
.c
'(i; <> O- lmm

-
c 10
c:
Q) - 1· 201m
-0- 2-3mm
c; 8 - 3-4111111
c ..... 4· 5mm
~
i.i3 - 5·60101
6
c:
Q)
... 6-701'"
c:: - 7-8111111
8. 4 ..... 8-9ml1l
E - 9- 10 111(1)
0
U -<>- 10-IIIIIRI
2
t: - 11 - l2mlll
0
..c::
(f) 0
0 5 10 15 20 25 30 35
Drying time 112 (min.) 1/2
Figure 6. The 12 profi les of bound water fo r Sample 3

2 12
.c 90- 1111111
.~ 10 ... ) . . 2mm

-
!c ~ 2-3mlll

8 - 3-4R101
..... 4-5mm
,- 5-6 mm
c:c:.> ... 6-7 mm
c:: - 7·811101
8. 4 .... 8·9 01111
E . . . - - - -....... - 9- 1011101
o
U 2 - 10·11011l1
1: _-_ _II
__- Ilmm
_ _ _ ...J

o
..c::
(f) o -------------------------------
o 5 10 15 20 25
Drying time 112 (min.)112
- --
FIgure 7 The 12 pro files of bo und water for Sample I (invol ving Surface active agent)

While il is poss ible that th is is due t o the additional drying damagi ng th e del icale ge l
400
structure of the mortar, we believe that the result is significant and reflects the different
uptake characteristics of refractory mortar to capillary and to bound water. Based on the
tits, the diffusion coefficient for bound water of Sample 1 is calculated to be D(c)=(3.4-
6.2)* 10-5 cm 2/s. On the other hand, the diffusion coefficient for combined capillary water
of Sample 3 is difficult to detennine as Fickian diffusion because of bad linearity of
Figure 6. Thus it appears that the Case 2 type diffusion or Case 3 type diffusion is
occurring in the case ofibound water for Sample 3.

0 ·- -- - - - - ------- -_.- ---,,- 0


,.-.
,..-...
, -0.5 - -0.1
I
'-'
'-' .....
<lI

.....
<lI C'l
l.. - sample ) free
C'l
l.. -0.2 l.. waler
l.. ~
-1.5 In .... sample3 fr ee
~ c
In
C
-0.3 C'l water
-2 "
l..
!-< , -=- sample)
~ l.. bound water
l.. -0.4 .....<lI .... samplc3
.....
Q)
~
C'l bou nd water
~ ~
Q)
-0.5 '0
Q) C
l.. ::J
~ 0
-3.5 -0.6 p:)

-4
0 2 4 6 8 10 12
Distance from Surface (mm)

Figure 8. T he total behavior of dry ing for Sample I and 3 at each posilion

Figure 8 shows the total drying for Sample 1 and 3 at each position. As these results
depend on the amount of water at a given position in the sample versus time, the axis of
Figure 8 means the dynamics of water transport in the sample. In general, free water is
more easily transferred from the bottom to the surface than bound water for drying and
homogeneous drying isideal at each position. In the case of Sample 3, it is clear that the
profile of both free water and bound water are almost the same and the large difference
between water transfer rates between the near surface and near bottom indicates
inhomogeneous drying. Moreover, the profiles of the two water transfer rates is almost
parallel. This means that bound water has a strong chemical affinity for refractory mortar
which causes the long drying time (1080 min.). On the other hand, it is clear that in the
case of Sample I the profiles of both free water and bound water are quite different. As
401

the distance from surface increases, both water transfers slow down, but the difference in
water transfer rates between the near surface and near bottom is small. Moreover, the
bound water transfer rate is almost constant because of the surfactants. This means that
in the Sample 1 homogeneous drying is maintained. The whole drying process occurs in
very favorable conditions because the surfactant weakens the relationship between the
refractOlY mortar material and bound water [16-17]. These results show the surfactant is
very effective and helpful to optimize the drying condition for refractory mortar in real
process.

4.CONCLUSION
It has been shown that the amplitudes of a simple two-component T, analysis of
mortar pastes correlate well with the expected fractional volumes of chemically
combined and capillary water. It has further been shown that stray-field imaging can
quantify all water. The techniques allow detailed quantitative study of the effects of
drying regimes that result in non-uniform drying and the effects of surfactants in the
water. Differences between the drying process of capillary water and that of combined
capillary water have been observed and quantified. These results are very useful to
optimize the dlying conditions in real steel-making process.

REFERENCES
I. Bluemich,B. and Kuhn,W, Eds.(1992), In 'Magnetic Resonance Microscopy', VCR, Weinheim

2. Bluemich,B., Bluemler.P .• and Saito.K .• (\998) in 'Solid State NMR of Polymers'. Elseiver. Amsterdam.

123-161

3. Gummerson. R.J .; Hall, C.; HofT. W.D.; Hawkes. R.; Holland. G.N.; Moore. W.S. (1979) Unsaturated

water now within porous materials observed by NMR imaging. Nature 281. 56-57;

4. Papavassiliou. G.; Milia. F.; Fardis, M.; Rumm. R.; Laganas. E. (1993) 'H nuclear magnetic resonance

imaging of water diffusion .in hardened cement pastes. J. Am. Ceram. Soc. 76. 2109-2111

5. Kaufmann, J.; Studer. W. (1995) One-dimensional water transport in covercrete-application or non·

destructive methods.Mat. Struct. 28.115-123

6. Fordham. E.J., Roberts, T.P.L.; Carpenter. T.A.; Hall.L.D.; Maitland. G.c.; \-Iall,C. (1991) Nuclear

magnetic resonance imaging of simulated voids in cement slurries. AIChE J.37,1895-1899

7. Nunes. 1'.1 Randall. E.W.; Samoilenk Feio. G. (1996). The hardening of Portland cement studied by 'H

NMR stray field imaging. J. Phys. D App\. Phys. 29.805-808


402
8. Bogdan. M.; Balcom. B.).; Bremner. T.W.; Armstrong.R.L. (1995) Single-point imaging of partially dried,

hydrated white Portland cement. J. Magn. Reson . A 116. 266-269

9. Black. S.; Lane. D.M.; McDonald. P.1.; Mulheron. M.;Hunter. G.; Jones. M.R. (1995) The visualisation

of the ingress of polymer treatment coatings into porous building materials by stray tield magnetic

resonance imaging. J. Mat. Sci.Lett. 14.1175-1177

10. Bohris. A.1.; Newling. B.; McDonald. I'.J. (1998) A broad line NMR study of water absorption and

transport in fibrous cement rooting tiles. J. Mat. Sci. 33. 859-867

II. Halperin. W.P.; Jehng. J.-Y.; Song. Y.-Q. (1994) Application of spin-spin relaxation to measurement or
surface area and pore size distributions in a hydrating cement paste. Magn.Reson. Imaging 12.169-173

12 . Jehng, J.- Y.; Sprague, D.T.; Halperin. W.P.(I 995) Pore structure of hydrating cement paste by magnetic

resonance relaxation analysis and freezing. Magn. Reson. Imaging 14.785-792

13. Papavassilioll. G.; Fardis. M.; Laganas. E.; Leventis. A.;Hassanien. A.; Milia. F.; Papageorgiou. A.,

Chaniotakis. E. (1997) Role of surface morphology in cement gel growth dynamics: a combined nuclear

magnetic resonance and atomic force microscopy study. J. App!. Phys. 82,449-452

14. Samoilenko. A.A.; Artemov. D.Y.; SibeIdina. L.A. (1988) Formation of a sensitive layer in experiments

on NMR subsurface imaging of solids. JETI' Lett. 47. 417-419

15. McDonald. P.1.; Perry. K.P.; Roberts. S.P. (1993) A repetitive pulse variant of broad line gradient echo

magnetic resonance imaging. Meas. Sci. Techno!. 4. 896-898

16. Hall, C. (J 989) Water sorptivity of mortars and concretes: a review. Magazine Concrete Res. 41. 51-61

17. Crank, J. (1975) The Mathematics of Diffusion. chapt. 7. Oxford,UK: Oxford University Press ;
THE ADSORPTION OF POLYELECTROLYTES TO COLLOIDAL PARTICLES
MONITORED BY IH RELAXATION OF THE SOLVENT

B. SCHWARZ, M. SCHONHOFF

Max-Planck-Institute for Colloids and Interfaces,


D-14424 PotsdamlGolm, Germany

1. Introduction

In this work we investigate the hydration water in polyelectrolyte monolayers of Poly-


(allylamine hydrochloride) (PAH) and Poly-(diallyldimethyl ammonium chloride)
(PDADMAC). The polyelectrolytes are adsorbed to colloidal particles in aqueous or salt
solution. The zeta-potential and the I H NMR transverse relaxation rate of the water
protons are determined in dependence of polymer surface coverage. The relaxation of the
solvent in colloidal dispersions is increased by solvent immobilisation in surface layers,
and thus monitors the average amount and motional dynamics of solvent molecules at the
surface.

2. Methods and Basics

In order to obtain comparable relaxation results concerning surface immobilisation of


water, samples are prepared at a fixed (4% vol. in final samples) particle concentration.
Polyelectrolyte solutions are added at varying concentrations. The electrophoretic mobility
(zeta-potential) is determined after further dilution. For determination of the relaxation
time T2 the Carr-Purcell-Meiboom-Gill (CPMG) sequence, described as dl-900 x,-(d2-
180° x'-)n, is applied. The 90° pulse length is 14.2 Ils, the delay d2 is 20 ms, and the sample
temperature is set to 25.0° ± 0,3 0c.
In the regime of 'fast exchange' the relaxation rates of molecules in two sites are averaged
quantities, weighted by the fraction ./A,B of molecules in the respective site A or B [1]:
R2 = fAR2A + fBR2B (1)
In agreement with fast exchange, for all samples a monoexponential decay from the
integral of the echo is observed. To facilitate the evaluation of relaxation rates in
composite systems, a specific relaxation rate R2sp , normalised on R/ of free water, has
been introduced [2], and is defined as:
R2sp = (R2 - R~)/ R~ (2)
R2sp monitors the relative change of the relaxation rate due to the presence of water binding
sites and is an additive parameter, In adsorption samples, the quantity L1R 2sp is calculated
403
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 403--408.
© 2002 Kluwer Academic Publishers,
404

from the experimentally determined R2sp by


M 2sp = R2sp - (Ri; + R:~) (3)
Here, the water immobilisation of a noninteracting system, given by the sfcecific relaxation
rates of the particle solution, RJspCP, and of the polymer solution, R!sp 01, is subtracted.
Throughout this paper, we consider R!sp as a measure of water immobilisation, as reflected
by the quantities h ·R JX in Equ. (I).

3. Results and Discussion

In order to investigate the immobilisation of water in adsorption layers, the effect of water
binding to either particles or polymers in solution has to be considered. As reference data
relaxation measurements of binary systems of particles in water (R!spCP), and polymer
solutions (R!s/o/ ) are investigated and published elsewehere [3].
The adsorption of polyelectrolyte to polystyrene latex particles is monitored by
electrophoretic measurements and relaxation rates in samples of fixed particle
concentration (4 % wt.) and varying polymer concentration. The zeta-potential (see Fig. 1)
is negative for uncoated particles, due to the negative stabilisation charges. With the
addition of positively charged polyelectrolyte, each zeta-potential is increasing mono-
tonously with the polymer concentration. After a charge reversal the increase is finally
resulting a plateau when the surface is saturated with polymer. Hence the zeta-potential
monitors the adsorption process of each polyelectrolyte in a qualitative way. Comparing
the zeta-potential for both polyelectrolytes in water the same value is reached, though the
saturation occurs at larger adsorbed weight for PDADMAC. This, however, corresponds to
the same number of monomers per surface area. In the presence of salt, saturation is
achieved at higher concentration, indicating a larger adsorbed amount.

~~.-~~~.--.--~.--.--~,-~~

50 ! ..
. .'
A o· ~
A
40

>"
30

20 c-
.." ;>

.§. 10 o

:sc
8... -10
0 o
~.

o
V -20
Fig. 1: Zeta-potential of particles coated with -30
~ : PDADMAC in H20. 6: PAH in H20 and
0 : PAH in a 0.25 M NaCI solution
-50 L....-L-........L_..........- 1 -............_L.......o......L-............_L.......o......L-............:-""':"
-0.02 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22
polyme.r I~mll

Relaxation data are displayed in Fig. 2a-c, where the horizontal line L!R!sp = 0 corresponds
to a noninteracting system, with R!sp equal to the sum of the values for the particle solution
and the polymer solution at the respective concentrations. Deviations from this line arise
from differences in the water immobilisation due to polymer-particle interactions. In all
systems investigated, a decrease of R!sp is observed with increasing polymer concentration,
405

followed by a maximum and a further decrease at higher polyelectrolyte concentrations.


Finally ,1Rlsp forms a plateau in which further added polymer does not interact with the
particles and therefore does not cause further deviations from zero.
In the salt free solution for the polyelectrolytes PDADMAC (Fig. 2a) and PAH (Fig. 2b),
the plateau values are at negative ,1Rlsp, indicating a net increase of water mobility and
therefore less bound water in the layer compared to the bare particles. This increase can be
attributed to a release of particle bound water due to replacement by adsorbing polymer
segments.

a.) 0.2 PDADMAC


Fig. 2: IJR},p af the adsorption of a.)
PDADMAC in H20, b.) PAH in H20 and 0.1
c) PAH in a 0.25 m NaCI solution at o
varying polymer concentrations. The M 0.0 J.e-~o...;.....t'---------------l
particle concentration is 4 wt.%. The lines lsp 0

are guides to the eye.


·0.1

o
·0.2 o

The adsorption of PAH from a


0.25 M NaCI solution (Fig 1)
b.) ILl
causes no significant difference PAH
in the shape of the zeta potential
0.1
curve. Similarly, the relaxation
rate curve shows the same
M 1sp 0.0 J..\--:--->f'f~-------------l
structure at low polymer
concentrations as compared to -0.1
adsorption from water solution.
At high polymer concentrations, ·0.2

a plateau is formed as well,


however, in salt solution the
plateau value is at positive ,1R],p, c.)
0.2 PAH INaCI
indicating a net decrease of water
mobility.
The structure of the ,1R2Sp curves
in dependence of polymer
0.1

: ~"
, ,. ......
M1spo.0 ~--~...,.---'.P-'-~~---------1
: ... 0'.. ...
" "

"
, '.
concentration can be interpreted
taking into account the . 0 .1
~ ..,
'.'" :"
,;.
"
..•
contributions of water in
different sites. The water -0 .2
"
fractions which contribute to R2sp
in samples containing polymers 0.04 0 .06 0 .08 0.10 0 .12 0 .14 0 .16 0.18

and particles, are the following: polymer concentration [mg/mll


jA : water bound to the particle
surface (hydration shell of the
stabilisation charges),fB: hydration water in an adsorbed polymer layer,fc: water bound to
free excess polymer in solution and !D: free water. Starting from the zero value at zero
polymer concentration, ,1R2sp is first decreasing. At this low polymer amount the dominant
406

effect which the interaction of polymer and particles has on the water mobility is the
replacement of surface bound water molecules by polymer segments. This leads to a
decrease of fA and an increase of ID. Since it is RJA > RJD , the result is a net decrease of
R!sp, reflecting a net 'mobilisation' of water. Simultaneously, however, an immobilisation
of water is expected to take place, since the motional dynamics of adsorbing polymer
chains are changing. This causes an increase of the relaxation rate of the hydration water
of the polymer as it is adsorbed, or, in other terms, an increase oflB whilefc is decreasing.
In a first approximation the immobilisation should scale linearly with the adsorbed amount
of polymer, assuming that the hydration of adsorbed and free chains is identical. This
immobilisation is starting to compensate the decrease offA at a coverage where additional
polymer segments adsorbing do not release further water from site A. In the case of
adsorption from water solution (Fig. 2a, b), the net effect at high coverage is an increase of
water mobility, while in the presence of salt (Fig. 2c) the contribution from fB
overcompensates that of water release from the particle surface. This result is consistent
with the formation of a flat layer of low thickness in the presence of no salt and with the
formation of a coiled layer of higher thickness in the presence of salt.
Additional contributions to LlRJsP can arise from an increase of the polymer density in the
layer with increasing coverage, leading to an increased net water mobility due to release of
a fraction of the water from the layer. Such effects might account for the decrease of LlR!sp
observed after the maximum at about 0.04 mg/ml. On the other hand, in this region of
rather low surface coverage a polymer chain might be attracted by different particles,
leading to bridging flocculation. This would cause water immobilisation due to a strong
motional restriction of water molecules bound to bridging polymer chains, and might
account for the formation of a maximum in LlR Jsp . Fig. 3 shows R Jsp determined in samples
of varying particle concentration. The data are rescaled to the same surface area.

0.8

0.7

0.6

cI 0.5

Fig. 3 Influence of bridging flocculation 0.4


II
at different particle concentrations at a t:. 6
t:.
fixed polymer/particle ratio. The ratio 0.3
corresponds to the maximun of [jR,'P in
the adsorption of PAH from H20 and
0.2
rescaled to R,.,p at 4 wt-%. 2 3 4 6
weight%-Iatex

An influence of coagulation occurs only at particle concentrations> 5% wt. The results are
equal for both polyelectrolytes and for PAH in salt solution. Thus coagulation does not
influence the data in Fig. 2.
It is interesting to note that all three plateau values of LlRJsP (Fig.2) are reached already in
the region of charge reversal (Fig. 1). In the case of adsorption without salt for both
electrolytes (Fig. 2a, b) the saturation of the relaxation rate (plateau) is reached at about S
407
= 0 (compare Fig. 1) and therefore at a lower coverage than saturation of zeta-potential.
Consequently the last adsorbing polymer chains do not immobilise further water. From
scaling theory [4], a picture has been put forward where the polymer is adsorbing in a flat
configuration, with different chains overlapping to form a 2D network (see Fig. 3). Within
such a model, our data imply that the addition of further chains does not change the
hydration water properties in the network layer, once charge reversal is achieved. The
addition of polymer after charge reversal probably leads to a closer packing within the
adsorption layer, such that the immobilisation of the hydration water of additional chains,
which would cause an increase of tJ.R2.,p can be compensated by a release of water from
the layer.
Contrary to these results in the presence of salt (Fig. 2c) the saturation of the relaxation
rate is reached at the same polymer concentration as the saturation of the zeta-potential
(see Fig. I). Therefore the adsorption of the last chains immobilises further water. Such a
picture would be expected for the adsorption of a monolayer of globular coils rather than a
network of overlapping rods, since the surface charge would scale with the fraction of the
surface covered by polymer coils (Fig. 3). It is reasonable that the polymer should adsorb
in a coil-like geometry from salt solution, since the hydrodynamic radius is smaller due to
electrostatic screening, and it has been argued before for planar layer systems, that with
increasing salt content a transition from adsorbed rods to coils occurs [5]. Even close to
saturation, additional adsorbing chains add immobilised water to the layer, since they
adsorb to previously uncovered parts of the surface. Fig. 4 gives a simple model of the
polymer configuration in adsorption layers in dependence of the salt content.

Fig, 4 Sketch of the structure of


adsorbed polyelectrolyte layers.
Flat chains adsorb from water
solution and coil-shaped chains
adsorb from salt solution.
no alt with salt

4. Conclusions

We have demonstrated here, that solvent relaxation is a suitable tool to resolve details of
the adsorption process in monolayer formation. The molecular picture suggested is a
network of overlapping rods in the case of adsorption from water solution. On adsorption
from salt solution, the results point at the formation of a layer of laterally organised coils
The details resolved here for monolayers on colloidal particles are consistent with previous
results on the structure of mono layers on planar support.

5. Acknowledgements

We would like to thank H. Zastrow for the preparation of the HZI-Iatex, A. Beer for
assistance with zeta-potential measurements. H. Mohwald is acknowledged for general
support. This work is funded by the German Science Foundation, DFG
Schwerpunktprojekt "Polyelektrolyte".
408

6. References

I, Hanus, F, and Gillis, p, (1984) Relaxation of water adsorbed on the surface of silica powder, J Magn. Reson.
59, 437-445,
2, van der Beek, G, p" Cohen Stuart, M, A, and Cosgrove, T. (1991) Polymer adsorption and desorption
studies via I H NMR relaxation of the solvent, Langmuir 7, 327-334.
3. Schwarz, B. and SchOn hoff, M. (2002) A I H NMR Relaxatil?n Study of Hydration Water in Polyelectrolyte
Mono- and Multilayers adsorbed to Colloidal Particles, Col/Old and SUlfaces A 198-100. 293-304,
4. Netz, R R, Joanny J. F. (1999) Adsorption of semi flexible polyelectrolytes on charged planar surfaces:
Charge compensation, charge reversal, and multilayer formation , Macromolecules 32, 9013-9025,
5, y, Lvov, G. Decher, H, Mohwald (1993) Assembly, structural characterisation, and thermal behaviour of
layer-by-Iayer deposited ultrathin films of poly (vinyl sulfate) and poly(allylamine), Langmuir 9, 481-486,
1-D AND 2-D DOUBLE HETERONUCLEAR MAGNETIC RESONANCE
STUDY OF THE LOCAL STRUCTURE OF TYPE B CARBONATE
FLUOROAPATITE

H. SFIHI 1 and C. REy2


lLaboratoire de Physique Quantique, CNRS FRE 2312, ESPCI, 10 rue
Vauque/in, 75005 Paris, France.
"CJRIMAT, CNRS UMR 5085, Institut National Poly technique, 38 rue
des 36 Ponts, 31400 Toulouse, France.

Abstract
The local structure of a type B carbonate fluoroapatite has been investigated by 1-D
and 2-D 13C{ IHrF and 31p{ IH/19F} MAS Nuclear Magnetic Resonance. The results
clearly show the existence of two type B carbonated sites and three fluorine sites.
One of the C0 32 sites is located near the apatite surface and in very close proximity
to strongly adsorbed water. The other type is close to two of the three fluorine sites.

1. Introduction
The association of C032- ions with apatites is particularly important for understanding
the mineral structure of bone. Indeed, carbonate is the third most abundant constituent
of bone and tooth mineral after calcium and phosphate, and plays an important role in
the maturation and the reactivity of the mineral phase of calcified tissues. Depending
on the substitution sites, one distinguishes type B carbonate apatite where some of
the P0 4 groups are replaced by CO/" and type A carbonate apatite where the CO/"
are located in apatitic channels generally occupied by hydroxide and/or fluoride ions.
Various methods such as molar composition determination [1], X-ray diffraction [1-
9], 'scanning electron microscopy [6], infrared spectroscopy [6-13], Raman
spectroscopy [14], electron spin resonance [9,16], IH and I3C solid state nuclear
magnetic resonance [17] have been used to investigate the properties, the morphology
and the local structure of synthetic and natural (dental enamel and bone) carbonate
apatites. The mUltiplicity of the techniques clearly reflects the complexity of the
material. In addition, variability may be observed depending on the synthesis method.
For synthetic carbonate apatite, all of the mentioned studies concern carbonate
hydroxyapatite (types A and B) in different forms (powders, crystals and films). In
order to understand better the structural modifications induced in bone mineral by
diseases such as fluorosis [18-21] or by fluoride salts used in the treatment of
osteoporosis [21-23], 1-D CH
and 19F MAS, I3C MAS, CP-MAS and DCP-MAS)
and 2-D I3 C (H}, 13C{19F }, 3Ip{IH} and 3Ip{19F} heteronuclear chemical shift
correlation (HetCor) MAS nuclear magnetic resonance were used in this work to
investigate the environments of CO/"and of F ions in a weakly hydroxylated type B
carbonate fluoroapatite. Carbonate fluoroapatite constitutes a model compound close
to highly fluoridated mineral bones and geological deposits.
409
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 409-422.
© 2002 Kluwer Academic Publishers.
410

2. Experimental

2.1 PREPARATION OF CARBONATE FLUOROAPATITE.


The carbonate fluoroapatite powder sample was prepared by double decomposition
between a calcium nitrate and ammonium phosphate solutions containing fluoride and
carbonate ions, according to a precipitation method proposed by Vignoles [24]. In
order to observe the BC NMR easily, the carbonate ions that we used was enriched
sodium bicarbonate (99 % BC). The phosphate-carbonate solution (0.43 g of
(NH 4hHP0 4 ; 0.30 g of ~F, 1.56 g of NaHC0 3 and 5 ml of a 20% ammonia
solution in 80 ml of deionized water) was added dropwise with a peristaltic pump into
the boiling calcium solution (0.640 g of Ca(N03)z,4H 20 and 5 ml of a 20 %
ammonium hydroxide solution in 200 ml of deionized water). The precipitate was
filtered, washed with deionized water and dried in an oven at 100°C.
The precipitate obtained was characterized by chemical analysis, X-ray diffraction and
FTIR spectroscopy.
Calcium was determined by complexometry with EDTA [25], phosphorus of
phosphate groups by spectrophotometry of phosphovanadomolybdic acid [25],
carbonate by coulometry of the carbon dioxide evolved during acid dissolution, and
the fluoride was measured with a ion-selective electrode after dissolution of the
sample. The chemical composition is reported in table I.

TABLE J. Chemical composition of the carbonate fluoroapatite

Ca (% weight) P (% weight) CO) (% weight) F (% weight)

38.25 14.90 7.43 4.79

The atomic ratio (Table II) is consistent with the formation of a very carbonate - rich
fluOfoapatite.
Table II. Atomic ratios of the carbonated fluoroapatite

CalP CaI(P+C) C/(P+C) F/(P+C)

1.98 1.58 0.205 0.42

The FTIR data (not shown) indicate the formation of a type B carbonate apatite, where
carbonate ions replace phosphate ions. Very faint OH· bands are observed at 714. and
691 cm-', corresponding to the existence of a very low small fraction of OR groups
hydrogen-bonded to fluoride ions. The presence of cation vacancies in this type of
apatite where carbonate ions replace phosphate is in agreement with by the Ca/(P+C)
ratio which indicates a deficiency of calcium ions. Similarly, the F/(P+C) ratio reveals
an excess of fluoride ions (the theoretical FIP value for pure fluoroapatite is 0.33)
which has been attributed to the incorporation of F ions in the oxygen vacancies
created by the replacement of phosphate by carbonate groups [24]. However, such a
substitution has been questioned [26], and has never been proved by direct
determinations.
2.2 NMR SPECTROSCOPY
'H and 19F magic angle spinning (MAS) NMR measurements were performed at
500.13 and 282.46 MHz with Bruker ASX 500 and ASX 300 spectrometers operating
in a static field of 11.7 and 7.1 T, respectively. The two spectrometers are equipped
411

with 4 mm ultra fast (0 - 18 kHz range) spinning probes. For both nuclei, the
spinning frequency was 15 kHz. The recycle delay was 5 s for IH and 30 s for I9F, in
accordance with their spin lattice relaxation times. The IH and I9F chemical shifts were
referenced to external tetramethylsilane (TMS) and hexafluorobenzene (C6F6),
respectively.
The l3C MAS l3CCH} CP-MAS and DCP-MAS spectra were obtained at 125.77
MHz (11.7 T) at different contact times in CP-MAS and different reverse times in
DCP-MAS. The recycle delay (D1) was 5 s for CP-MAS and 800 s for MAS. The
relatively high D 1 value in MAS is due to the long l3C T I which is estimated to be
200 s. The spinning frequency was 5 kHz.
Differential cross polarization (DCP) [27] which was also called cross polarization
with polarization inversion (CPPI) [28,29], cross polarization depolarization (CPD)
[30] and inversion recovery cross Polarization (IRCP) [31,32] is particularly efficient
for resolving overlapping spectra. This method, which is more sensitive than the
normal CP, can be used to separate two or more distinct groups (sites or chemical
groups) on the basis of the difference (which could very small) in the strengths of
their dipole coupling. From the experimental point of view, this is achieved by
reversing the phase of the IH rf field after a forward transfer or contact time, te, such
that the l3C polarization is transferred back to IH during the reverse time t" This
causes the l3C spectrum to successively reduce in amplitude, pass through a null, and
finally become inverted.
The l3C{I9F} CP-MAS spectra were obtained at 75.47 MHz (7.1 T). The spinning
frequency was 5 kHz.
The 2D l3C{ IH} and l3C{ 19F} heteronuclear chemical shift correlation (HetCor) MAS
NMR [33-38] measurements were made at 125.77 MHz (11.7 T) and 75.47 MHz (7.1
T). The spinning frequency was 12 kHz.
The 3Ip{ IH} and 31p{ I9F} HetCor MAS NMR measurements were made at 202.45
MHz (11.7 T) and 121.47 MHz (7.1 T). The spinning frequency was 12 kHz.
For all 2D HetCor measurements a contact time of 5 ms was used. This value gives
the maximum of polarization as deduced from CP-MAS measurements. Phase
sensitive detection was obtained using TPPI phase cycling.
For both l3C{ 'Hl I9 F} and 3Ip {IH j I9F} CP-MAS, the Hartmann-Hahn conditions were
achieved on the sample studied. In order to reduce the time of the experiment in
l3C{ I9F} and 3Ip {I9F} CP-MAS and HetCor, a 90° flip-back-pulse was used to bring
the I9F magnetization along the Zeeman field just after the cross polarization (CP)
period. The long I9F spin lattice relaxation times in the rotating frame (TIP) make the
use of the flip-back-pulse possible. By this means the recycle delay is reduced from 25
s to 4 s. Furthermore, in all CP and HetCor experiments, eight saturation pulses were
applied on polarized spins (i.e. l3C and 3Ip) before the 90° pulse on polarizing spins
(i.e. IH and I9F). This suppresses the residual signal which could arise from the direct
excitation (during the CP period) of the polarized spins. This signal could be strong
particularly at very short contact times (0 -500 Ils). The l3C and 3Ip chemical shifts
were referenced to external TMS and to 80% H 3P0 4, respectively. The other
experimental conditions are given in the figures.
In this study HetCor [33,34] is used to correlate specific IH (or I9F) and l3C (or 3Ip)
with peak positions in the two-dimensional (2-D) NMR spectrum. This method is
analogous to conventional CP, except that the IH (or I9F) magnetization is allowed to
412

evolve for a period t .. which corresponds to the first time domain (F,) of a 2-D NMR
experiment, before the magnetization is transferred to dipolar-coupled l3C (or 31p) in
close spatial proximity for the second time domain (F:!), during which the signal is
detected. This method has been used in various compounds such as silicas and zeolites
[35], silica gels [36], aluminosilicates [37], fluorinated y-alumina [38], in bone and
various calcillm pho.sphates [39].

3. Results and discussion


3. 1. IH AND 19 F MAS NMR
The IH and 19F MAS NMR spectra are shown in figures I-a and 2-c, respectively. For
comparison the IH MAS NMR spectrum of stoichiometric hydroxyapatite [HAp,
CalO(P04)60H2] is also given (Fig. I-d).

d)
iii i , i , , iii' , iii' iii i
15 10 5 0
(wn/TMS)

Figure 1 : IH 1-0 MAS NMR spectra Figure 2 : 19F 1-0 MAS NMR spectra
a)'H MAS. a) 19F projection of the 2-0 IlC {19F } HetCor MAS.
b)'H projection of the 2-0 1lC{ IH} HetCor MAS. b) I~ projection of the 2-0 3Ip{'9F} HetCor MAS.
c)IH projection of the 2-0 3Ip{'H} HetCor MAS. ) 19F M S
d)IH MAS of HAp. CA.
The IH MAS spectrum contains two main resonance lines: a broad one centered at 5
ppm and a very structured one at 2 ppm. Similar results were reported by Beshah et
al. [17] for carbonate hydroxyapatite. The former is assigned to H20 which could be in
an adsorbed andlor structural form, as previously mentioned (17]. The latter is
located in the region of OH groups (0-3 ppm). At least four resonance lines are
observed in this domain: two resolved lines at 0.9 and 1.2 ppm and two shoulders at
the same chemical shift (0.2 ppm) as that of the IH of HAp, and at - 2 ppm. As
previously reported for various fluorohydroxyaptites, these resonance lines could
reflect the different configurations of OH groups in the apatitic channels [40-42,45].
413

The poor resolution observed for these lines couid be due to the remaining dipole
coupling (mainly IH_IH, IH_19F, ) which is not completely averaged by fast MAS.
The 19F MAS NMR spectra shows clearly the presence of two fluoride ion sites
represented by two relatively broad resonance lines at 65 and 78 ppm with different
intensities. The line at 65 ppm is attributed to fluoride ions located in the apatitic
channel, because its chemical shift is very close to that observed in
fluorohydroxyapatites [42-47]. Its broadness results mainly from a continuous
chemical shift distribution, although 19FYF, 19F_IH, and to a lesser extent 19 F_13 C and
19 FJl p dipolar couplings could also contribute. This chemical distribution is due to
the presence of OH groups in the apatitic channel which leads to different
configurations of F' ions [40,41,45]. In addition, the presence of cot ions in the
lattice could also influence the chemical shift of fluorides located in the channel. In
particular, one distinguishes F- ions close to cot and those close to pot. This being
said, the principal result revealed by 19F MAS concerns the fluorine site at 78 ppm.
However, the existence of such a site is not surprising considering that the F/(P+C)
ratio reveals an excess of fluoride ion in the sample. Its position within the structure is
still to be specified and will be discussed in detail later. The peak of weak intensity at
43 ppm is attributed to teflon impurities.

3.2. 13 C CP AND DCP MAS NMR


The DC MAS NMR spectra obtained at two Larmor frequencies (vo) are shown in
figure 3-a.

(a)

t: = 10 ms

t: = 5 ms

t: = I ms

..........,............,__..., _r-,........,............,--. t: = 0.5 ms i 4 , • $ ,


174 172 170 168 166 164 174 172 170 168 166 164
(ppm I TMS) (ppm/TMS)

Figure 3 : DC {IH} (left) and DC tF} (right) CP-MAS NMR spectra obtained at
different contact times,t:.
(a) DC MAS spectra: (left) VII = 121.47 MHz; (right) vo = 75.47 MHz.
At the two frequencies the spectrum consists of a single asymmetric line with
maximum intensity at 170 ppm. Similar results were reported by Besha et al. [17] on
414

type B carbonate hydroxyapatite. Note that no peak (166 ppm) corresponding to type
A carbonate species [17] was detected.
The 13C{ IH} CP-MAS spectra (Fig. 3) show that the cot located at 170 ppm are less
cross polarized than that 168.5 ppm, whatever the contact time. The carbon cross
polarization time constant depends mainly upon the magnitude of the 13C_IH dipole-
dipole coupling, i.e. the carbon-proton internuclear distance and the number of
couRled spins. The I3 C MAS measurements (Fig. 3-a) show that the fraction of
cot located at 168.5 ppm is much lower than that of the carbonates at 170 ppm.
Therefore, the fact that the amount of carbon at 170 ppm is underestimated in the
cross polarization measurements means that most of them are more distant from the
proton, probably close to fluoride ions.
In contrast to the 13C{IH}, the 13C{19F} CP-MAS spectra at different contact times are
very similar to the MAS spectrum. Nevertheless, the MAS spectrum is still much
broader than the 13C{ 19F} CP-MAS one, whatever the contact time. This means that
the C03 groups at 168.5 ppm are weakly cross-polarized by fluoride ions, and that
these carbonates are more distant from fluoride ions than those at 170 ppm. More
precisely, considering the fact that most of fluoride ions are located in the apatitic
channel, this result signifies that the C03 groups at 168.5 ppm will not be located in
the bulk but near the surface. These groups correspond to the labile C03. In addition
and as discussed above, the C03 groups at 170 ppm are on average more strongly
coupled to the fluoride ions than to the protons.
The I3 C CP-MAS measurements raised
several Questions and did not clearly show
whether the C03 groups at 168.5 ppm or a
part of those at 170 ppm are more strongly
coupled to the protons. In order to elucidate
this Question, a differential cross polarization
(DCP) experiment was performed at different
reverse times, te (Fig. 4).
The choice of the contact time which gives
the reverse of the I3 C magnetization in DCP
measurements is empirical. The value used in
this work (tc = 300 us) is same as that used
by Wu et af. f271 to separate the overlapped
protonated (HP04) and unprotonated (P04)
phosphate groups in bone mineral.
The DCP shows clearly the existence of two
type B carbonate sites characterized by
slightly different cross polarization time
constants, confirming the Qualitative analysis
Figure 4 : 'lC{IH} DCP-MAS NMR of the I3 C CP-MAS spectra. The amplitude of
I3 C magnetization decreases with increasing
spectra at different reverse CP times tr,
te, passes through a null, and becomes
and at contact time,
negative. The fact that the peak at 170 ppm
Ie = 300 "s : INS = 8196), inverted first means that the cross
415

polarization time constant of the corresponding C03 groups is slightly shorter than that
of the C03 at 168.5 ppm. This means that, at the local level, the former are slightly
closer to the protons (more strongly coupled) than the latter. Therefore, the I3C CP-
MAS results cannot be explained by a difference in the distances between protons and
each of two types of the C03 group, but by a difference in the total number of protons
cross polarizing each of them. The total number of protons cross polarizing the labile
C0 3 (168.5 ppm) is greater than that of those cross-polarizing the regular type B C03
(170 ppm). This means that these protons are different and, as we will see in the
next section, they correspond to adsorbed water molecules and to the OH groups in
the apatitic channel, respectively.
3.3. 13C{ lH} AND I3C {19F} HetCor MAS NMR
The NMR measurements presented in the previous sections, indicated clearly the
existence of two main type B carbonate sites. On average, one carbonate (168.5 ppm)
is slightly more coupled to the protons than the other (170 ppm), although locally we
observed the opposite. However, this other carbonate is strongly coupled to the
fluoride ions which are located in two different sites. They also show that the
carbonates strongly coupled to the protons (168 .5 ppm) are hardly coupled to the
fluoride ions (weakly cross polarized by fluorine) .

~o

",'

60 ,,'
E
.2;
c
80 ~
~
,;;
IOOf-

I 112 170 168 166


He dimension (ppm fTMS) "e dimensio n (ppnvTMS)

Figure 5 : Contour plots and projections of 2-D Figure 6 : Contour plots and projections of the 2-D
13C{IH} HetCor MAS NMR . 13C{19F} HetCor MAS NMR

In order to obtain more detailed information on the environment of these different


sites (carbonates and fluorine ions), and particularly which carbonate ions are in close
proximity to which protons and to which fluoride ions, 13C{ lH} and I3C {19F} HetCor
MAS NMR measurements were performed on the sample.
Figures 5 and 6 show the contour plots of 2-D 13C{ lH} and 13C{ 19F} HetCor MAS
NMR and the projections in the F2 (13C) and FJ CH
or 19F) dimensions.
The 2-D l3C{ lH}HetCor MAS experiment shows that the water protons (5 ppm) are
closer to the C03 at 168.5 ppm than to that at 170 ppm. This is clearly illustrated in
Figure 7-a. The 1-D spectrum obtained by summation of the 2-D I3C row data clearly
indicates that the C03 groups at 168.5 are mainly cross-polarized by the protons in the
416

range 4-10 ppm (H 20). The OH protons (0-2 ppm) practically hardly participate at all
in the polarization of these carbonates. However, these protons interact strongly with
the C03 groups at 170 ppm (Fig. 7-b). This result indicates a heterogeneous
distribution of ions and can be interpreted in several ways.
It may be interpreted as follows: if the C03 at 168.5 ppm were in the bulk, they
should be cross-polarized by the OH protons OH (0-2 ppm) as well as by fluoride ions
located in apatitic channel. This is not the case. Therefore, we conclude that these
C03 groups are close to the apatite surface (labile C03) and that the water molecules
are mainly in the adsorbed form (strongly adsorbed), as mentioned in the previous
section. Concerning the OH protons, comparison of the projection in I H dimension
with the IH MAS spectrum (Fig. I-a and I-b) reveals that only the OH protons at 0.2
and 2 ppm and appearing as a shoulder in the MAS spectrum participate in the cross-
polarization of the C03 groups at 170 ppm. This means that the protons at 0.9 and 1.2
ppm are not or are very weakly coupled to both C03 sites. Two similar peaks have
been reported in octacalcium phosphate by Yesinowski et al. [41]. These peaks were
attributed to the protons of water molecules undergoing rapid reorientation on the
NMR time scale [41]. This attribution would be fully consistent with the fact that the
0.9 and 1.2 ppm protons were neither cross-polarized by the I3C of both C03 sites nor
by the 31 p of P0 4 (see below). Nevertheless, these water molecules are, according to
these authors, located in isolated sites in the structure of OCP. In our case the water
would be located at the surface (weakly adsorbed) rather than in the structure. We
assign the OH protons at 0.2 and 2 ppm to OH'OH (HAp) and to OHF
configurations in the apatitic channel, respectively, as previously reported for
fluorohydroxyapatiteli containing different amounts of fluoride ions [41]. This
assignment is corroborated by the relative intensities of the corresponding peaks,
which are in good agreement with the chemical data. Indeed, the amount of fluoride
ion is greater than that of OH- and therefore the OH ....F configurations should be
predominant. In addition, the downfield location of the corresponding peak (2 ppm
instead of 0.2 ppm) could be explained by strong hydrogen OH ....F bonds.
The most important result revealed by the 2-D 13C{ 19F} HetCor MAS experiment
concern the two fluoride ion sites and particularly the site located at 78 ppm. Both
types of fluoride ion are coupled to the carbonate species, located at 170 ppm (see also
Figures 7-c and 7-d), as expected. However, the relatively high intensity of the peak
corresponding to fluoride ions at 78 ppm, compared to that observed in the 19F MAS
spectrum (Fig. 2-c-right), means that these fluoride ions are more strongly coupled to
the carbonate ions than those located at 65 ppm (fluoride ions in the apatitic channel).
Indeed, they are better cross-polarized. At 500 Ils contact time (result shown), the
peak at 78 ppm is twice as intense as that at 65 ppm. These results indicate that the
fluoride ions corresponding to the former peak are closer to carbonate ions at 170 ppm
than those corresponding to the latter. These fluoride ions are probably located at the
oxygen vacancies created by the replacement of phosphate groups by carbonate
groups (see Fig. 8-left). The replacement of phosphate by carbonate creates a
negative charge deficit which is compensated by the creation of a vacancy in a
nearby calcium site and a vacancy in a close monovalent ion site (OH- or F) [1,4,24],
as shown in figure 8-left. These carbonate ions are thus far from F and OR ions, and
417

the eXlstmg ionic vacancies could be


occupied by water molecules. This
assumption can, however, be rejected,
because in this case the protons of the
water molecules, which would be
located in monovalent (OR or F) or
oxygen vacancy sites, should be more
strongly coupled to the carbonate ions at
168.5 ppm than the OH protons located in
the channel to the carbonate ions at 170
ppm. This is not the case, as is revealed by
the DCP-MAS measurements (the cross
polarization time constant of the carbonate
ions at 168.5 ppm is slightly shorter than
d)
that of those at 170 ppm). The water
174 Ih 170 168 molecules would be mainly in the adsorbed
(ppmITMS) form. However, on the contrary, if a
phosphate is replaced by C03,F association
Figure 7 : 13C NMR I -D spectra (Fig. 8-right), as previously mentioned
obtained by summation of the 2-D
13C HetCor/MAS row data [1,4,24], although such a substitution
con'esponding to the C0 3 has been questioned [26], there is no
interacting mainly with protons i local charge deficit and the cationic and
range: 10 -4 ppm (a) ; 2-0 ppm (b)
anionic vacancies do not form. The fluoride
and with fluoride ions in range: 70
- 50 ppm (c) ; 75- 85 ppm (d). ion (F3 site in fig . 8-right) in the C03,F
association, which may correspond to the
peak at 78 ppm in the 19F MAS spectrum, should be slightly more strongly coupled to
the corresponding C0 3 than the fluoride and/or OR ions located in the apatitic channel
(F2 site). Another point confirming the attribution of the peak at 78 ppm to the F3 site
is its relative intensity. From chemical data the amount of C03,F which corresponds
to the fluoride ion excess, would be close to approximately 30 % of the total amount of
fluoride. This value is consistent with the relative integrated intensity of the peak at
78 ppm deduced from a rapid peak fitting of the 19F MAS spectrum (result not
shown).
The peak corresponding to the fluoride ions located in the apatitic channel (65 ppm),
appears only as a shoulder in the projection in the 19F dimension of the 2-D HetCor
measurements (Fig. I-a). In addition to the shoulder we notice another unresolved
peak at "" 70 ppm. The overall intensity of both peaks is dominated by that of the peak
at 70 ppm. Although both peaks correspond, as previously indicated, to fluoride ions
in the apatitic channel, this result reveals that the fluoride ions at 70 ppm are more
strongly coupled to the carbonate ions at 170 ppm, than those at 65 ppm. We
attributed the latter to the F 1 site (Fig. 8-right). It corresponds mainly to fluoride ions
in a phosphate environment. The peak at 70 ppm, which we assign to fluoride ions in a
carbonate environment (F2 site), should also be also present in the 19F MAS spectrum.
That this peak did not appear clearly in this spectrum can be explained by the fact
418

that the number of P04 (hence the number of fluoride ions close to P04) is greater
than that of the C03 (hence the number of fluoride ions close to C03). Chemical
analysis indicates that only 20% of the phosphate sites were replaced by a carbonate
ion . This is in good agreement with the fact that the intensity of the 19F MAS
spectrum is dominated by the peak centered at 65 ppm (FI site) compared to that of
the peak at 70 ppm (F2 site) .

o P
• C
• Ca
eo
o Vacancy

----. Z=1/4-----

Figure 8: Site occupancy oftluoride ions in carbonate tluoroapatite as revealed by NMR.

Another interesting result revealed by the 2-D I3 C{ 19F} HetCor MAS experiment is
the existence of a third fluoride ion site. This site, which is represented by the peak at
90 ppm in the 19F spectrum obtained by the projection in the 19F dimension of the 2-
D HetCor measurements, is at the same chemical shift as that of ammonium fluoride
(NH4F). This ammonium fluoride is an impurity of the synthesis and could be located
on a calcium-deficient site, as previously mentioned [9,48]. This perfectly explains
the cross-polarization of the NH4F fluoride ions by the C03 carbon and vice versa.
However, the amount of NHF4 remains ve~ small since no significant signal
corresponding to this species was detected by 9F MAS (Fig. I-c). This means that
all the oxygen vacancies would not occupy by F3 sites, as discussed above, but some
of them (same amount as that of NH4F which is very weak) remain unoccupied. It is
the same for the vacancies at the monovalent ion site (F"2 or OH) Similarly, this
result suggests that the projection in the IH dimension of the 2-D 13C{ IH} HetCor
MAS measurements (Fig. lob and 5-left) contains a contribution of NH4F whose
protons resonate at"" 7.2 ppm.

3.4. 31p{ IH} AND 31p tF} HetCor MAS NMR


31 1 31 19
The contour plots of 2-D P{ H} and P { F} HetCor MAS NMR and the
projections in the F2 e1p) and FJ (IH or 19F) dimensions are shown in figures 9 and
10, respectively. The 2-D 3IP{IH} HetCor (contours and projections) shows that the
419

P0 4 groups are connected to the same protons as the two type B C03. However, the
peak at 0.2 ppm, which we assigned to OH in OH"'OH configurations, appears more
clearly in the projection in the IH dimension. Furthermore, the intensity of both
peaks (0.2 and 2 ppm) is higher than that observed for the same peaks obtained by
the projection in the IH dimension of the 2-D \3C {IH} HetCor (see figures I-b and
I-c). This.is explained by the fact that there are more P0 4 groups than C0 3 species.
The protons of water molecules seem to be less cross-polarized by phosphorus than
by the labile C0 3, although the number of P04 groups is greater than that of the latter
species. Furthermore, the peak intensity of these water molecules contains a
contribution from NH4F whose protons resonate at 7. 2 ppm. This .means that the P0 4
groups are more distant from the water molecules than the labile C03, confirming
then that the water molecules are adsorbed (strongly) at the surface. We could also
note that the peaks at 0.9 and 1.2 ppm observed in the 'H MAS spectrum and
previously discussed in detail are absent from the projection in the 'H dimension
(Fig. l-c and 9). This confirms the attribution of the two peaks to water molecules
undergoing rapid reorientation at the apatite surface.

o
U p dmlC'1'l510n (ppm ' ~ttPOl)

Figure 9: Con tour plots and projections of the 2-D


Figure 10 : Contour plots and project ions of the 2-
l'p{ 'H} I fetCor MAS NMR
D l'p{ '~F} HctCor MAS MR.

The surprising result in the 31 p {19F} HetCor is that the projection in 19F dimension
leads to a spectrum very similar to that obtained by 19F MAS. This means that the
fluorine sites at 78 and 65 ppm are practically at the same distance from the P0 4
groups. This is a good agreement with a location of the fluoride ions at 78 ppm at
oxygen vacancy sites (F3 sites), confuming then the interpretation of the 2-D \3C{19F}
HetCor measurements.

4. Conclusion
We have shown in this work that the combination of I-D and 2-D double nuclear
magnetIc. resonance. mvo
. Ivmg
' d'ffi . (IH , 19F , 13C, 31p) can prOVl'de more
1 erent spms
detailed information on the local structure of carbonate fluorohydroxyapatites , and
more generally on materials of complicated structure as bone minerals.
420
In particular, the results show the existence of two type B carbonates and of three
fluorine sites in weakly hydroxylated carbonated fluoroapatite, and allow us to clearly
identify their local environments.
References
I. Labarthe J. e., Bonel G. and Montel G. (1973) Sur la structure et les proprietes des apatites
carbonatees de type B phospho-calciques, Annales de Chimie 8, 289-301.
2. Legeros R. Z. (1967) Crystallographic studies of the carbonate substitutions in the apatite
structure, PhD Thesis, New York University.
3. Elliott J. e., Bonel G. and Trombe J. e. (1980) Space group lattice constants ofCalO(P0.)6C03,
Journal of Applied Crystallography 13,618-621.
4. Bonel G. (1972) Contribution a I'etude de la carbonation des apatites I. Synthese et etudes
proprietes physico-chimiques des apatites carbonatees de type A, Annales de Chimie 7, 65-88.
5. Wilson R. M., Elliott J. C. and Dowker S. E. P. (1999) Rietveld refinement of crystallographic
struture of human dental enamel, American Mineralogist 84, 1406-1414.
6. Morgan H., Wilson R.M ., Elliott J. e., Dowker S. E. P. and Anderson P. (2000) Preparation and
characterization of monoclinic hydroxyapatite and its precipitated carbonate apatite, Biomaterials
21, 617-627.
7. Merry J. e., Gibson I. R., Best S. M. and Bonfield W. (1998) Synthesis and characterization of
carbonate hydroxyapatite, Journal of Materials Science: Materials in Medicine 9,779-783.
8. Doi Y, Moriwaki Y., Aoba T., Okazaki M., Takahashi J., Joshin K. (1982) Carbonate apatite
from aqueous and non-aqueous media studied by ESR, IR and X-ray diffraction: effect of NH. +
ions on crystallographic parameters, Journal of Dental Research 61, 429-434.
9. Xu G., Aksay I. A. and Groves J. T. (2001) Continuous crystalline carbonate apatite thin film. A
biomimetic approach, Journal of the American Chemical Society 123, 2196-2203 .
10. Rey e. and Hina A. (1995) Surface reactivity of bone mineral crystals, a model for bioactive
orthopaedic materials. Bioceramics 8, 55-60.
II. Elliott J. C . (1964) The crystallographic structure of dental enamel and related apatites, PhD
Thesis. Unversity ofLondon.
12. Elliott J. e., Holcomb D. W. and Young R. A. (1985) Infrared determination of degree of
substitution of hydroxyl by carbonate ions in human dental enamel, Calcified Tissue
International 37, 372-375.
13. Rey e., Renugopalakrishman Y., Shimizu M., Collins B. and Glimcher M. J. (1991) A
resolution- enhanced Fourier transform infrared spectroscopic study of the environment of CO/'
ion in the mineral phase of enamel during its formation and maturation. Calcified Tissue
International 49. 259-268.
14. Suetsugu Y., Shimoya I. and Tanaka J. (1998) Configuration of carbonate ions in apatite
structure determined by polarized infrared spectroscopy, JOllrnal of the American Ceramic
Society 81, 746-748.
15. Nelson D. G. A. and Williamson B. E. (1982) Low temperature laser Raman spectroscopy of
synthetic carbonated apatites and dental enamel, Australian Journal of Chemistry 35. 715-727.
16. Pekauskas R. A. and Pullman I. (1978) Radiogenic free radicals as molecular probes in bone,
Calcified Tissue Research 25.37-43.
17. Beshah K., Rey e., Glimcher M. J., Schimizu M. and Griffin R. G. (1990) Solid state carbon-13
and proton NMR studies of carbonate containing apatite and enamel, Journal of Solid State
Chemistry 84, 71-81.
18. Boivin G., Chavassieux P., Chapuy M.e., Baud C.A. and Meunier P. J. (1989) Skeletal
fluorosis: histomorphom~tric analysis of bone changes and bone fluoride content in 29 patients,
Bone 10,89-99.
19. Boivin G., Chavassieux P., Chapuy M.e., Baud e.A. and Meunier P. J. (1990) Skeletal
fluorosis: histomorphometric findings, Journal of Bone Mineral Research, 5 , 185-189.
20. Teotia S. P. S., Teotia M. and Teotia N. P. S. (1976) Skeletal fluorosis: roentgenological and
histopathological study, Fluoride 9, 91-98.
21. Boivin G., Chapuy M.C., Baud C. A. and Meunier P. J. (1988) Fluoride content in the human
illiac bone, results in controls, patients with fluorosis, and osteoporotic patients treated with
fluoride. Journal of Bone Mineral Research 3, 497-502.
421

22. Mamelle N., Meunier P. J., Dusan R., Guillaume M., Martin J. L., Gaucher A., Prost A., Zeigler
G. and Netter P. (1988) Risk-benefit ratio of sodium fluoride treatment in primary vertebral
osteoporosis, Lancet 2,361-365.
23. Riggs B. L., Hodgson S. F., O'Fallon W. M., Chao E. Y. S., Wahner H. W., Muhs J. M., Cedel
S. L. and Melton L. J. (1990) Effect of fluoride treatment on the fracture rate in postmenopausal
women with osteoporosis, New England Journal of Medicine 322,802-809.
24. a
Vignoles M. (1984) Contribution I'etude des apatites carbonarees de type B, These d'Etat INPT,
Toulouse, France.
25. Charlot G. (1966) Les methodes de la chimie analytique , Masson, Paris.
26. MacConnel D. (1973) Apatite, its crystal chemistry, mineralogy, utilization and geologic and
biologic occurrences. 'Springer Verlag" New York.
27. Wu Y., Glimcher M. J., Rey C. and Ackerman J. L. (1994) A unique protonated phosphate group
in bone mineral not present in synthetic calcium phosphate: Identification by phosphorus-31
solid state NMR spectroscopy. Journal of Molecular Biology 244,423-435.
28. Wu X., Zhang S. and Wu X. (1988) Selective polarization inversion in solid state high resolution
CP MAS NMR,Journal of Magnetic Resonance 77, 343-347.
29. Wu X. and Zilm K. W. (1993) Complete spectral editing in CP-MAS NMR, Journal of Magnetic
Resonance AI02, 205-213.
30. Sangil R., Rastrup-Andersen N., Bildsoe H. Jakobsen H. J. and Nielsen N. C. (1994) Optimized
spectral editing of llC MAS NMR spectra of rigid solids using cross-polarization methods,
Journal of Magnetic Resonance AI07, 67-78.
31. Zumbulyadis N. (1987) IHP9Si cross polarization dynamics in amorphous hydrogenated silicon,
Journal of Chemical Physics 86,1162-1166.
32. Gory D. G. and Ritchey W. M. (1989) Inversion recovery cross polarization NMR in solid
semicrystalline polymers, Macromolecules 22, 1611-1675.
33. Ernst R. R. , Bodenhausen G. and Wokaun, G. (1994) Principles ofnuc\ear magnetic resonance
in one and two dimensions, Oxford University Press. New York.
34. Schmidt-Rohr K. and Spiess H. W. (1994) Multidimensional solid-state NMR and polymers,
Academic Press. San Diego.
35. Vega A. J. (1988) Heteronuclear chemical shift correlations of silanol groups studied by two
dimensional cross polarization/magic angle spinning NMR, Journal of the American Chemical
Society 110,1049-1054.
36. Fyfe C. A., Zhang Y. and Aroca P. (1992) An alternative preparation of organofunctionalized
silica gels and their characterization by two-dimensional high resolution solid-state heteronuc\ear
NMR correlation spectroscopy, Journal of the American Chemical Society. 114. 3252-3255.
37. Janicke M. T., Landry C. c., Christiansen S. c., Kumar D., Stucky G. D. and Chmelka, B. F.
(1998) Aluminum incorporation and interfacial structures in MCM-41 mesoporous molecular
sieves, Journal of American Chemical Society 120, 6940-6951.
38. Fisher L., Harle Y., Kasztelan S., Man P. P. and d'Espinosse de la Caillerie J. B. (2000)
[uentilication of fluorine siles at the surface of fluorinated y-alumina by two-dimensional MAS
NMR, Solid State Nuclear Magnetic Resonance 16, 85-91.
39. Santos R. A., Wind R. A. and Bronnimann C. E. (I994)IH CRAMPS and IH_ 3I p HetCor
experiment on bone, bone mineral and model calcium phosphate, Journal of Magnetic
Resonance BI05, 183-187.
40. Freund F. and Knobel R. M. (1977) Distribution of fluorine in hydroxyapatite studied by infrared
spectroscopy, Journal of Chemical Society; Dalton Transactions 10, 1136-1140.
41. Yesinowski J. P. and Eckert H. (1987) Hydrogen environments in calcium phosphates: IH MAS
NMR at high spinning speeds. Journal of the American Chemical Society 109,6274 - 6282.
42. Yesinowski J. P, Wolfang R. A. and Mobley M. J. (1984), New NMR methods for the study of
hydroxyapatite surface, Adsorption on/and Surface Chemistry of Apatites, ed. Miskra D.N..
Plenium Press. New York. 151-175.
43 . Yesinowski J. P. and Mobley M. J. (\983) 19F MAS NMR of fluoridated hydroxyaptite surfaces"
Journal of the American Chemical Society 105,6191 - 6193.
44. Kreinbrink A. T. , Sazavsky C. D. , Pyrz J. W., Nelson D. G. A. and Honkonen R. S., (1990) Fast
magic angle spinning 19F NMR of inorganic fluorides and fluoridated apatitic .surfaces, Journal
ofMagnetic Resonance 88,267-276.
45. Braun M. and Jana C. (1995) 19F NMR spectroscopy of fluoridated apatites, Chemical Physics
Lellers 245, 19-22.
422
46. Braun M. , Hartmann P. and Jana C. (1995) I~ and lip NMR spectroscopy of calcium apatites,
JOl/rnal ojMaterials Science ,' Materials in Medicine 6, 150-154.
47. Iijima M., Nelson D.G.A., Pan Y., Kreinbrink A. T., Adachi M., Goto T and Moriwaki Y. (1996)
Fluoride analysis of apatite crystals with a central planar OCP inclusion: Concerning the role of
F ions on apatite/OCP/apatite structure fonnation, Calcified Tissue International 59, 377-384.
48. Vignoles M., Bone1 G. and Young R. A. (1987) Occurrence of nitrogenous in precipitated B-type
carbonated hydroxyapatites, Calcified Tissue International 40, 64-70.
SPATIO-TEMPORAL CORRELATIONS IN GRAVITY-DRIVEN
AND PRESSURE-DRIVEN FLUID TRANSPORT PROCESSES

S. STAPF, C. HEINE, S. HAN, B. BLUMICH


Lehrstuhl fUr Makromolekulare Chemie, ITMC,
RWTH Aachen, D-52056 Aachen, Germany

NMR pulsed-field gradient velocity encoding techniques and NMR imaging have
been combined for the investigation of flow behaviour for two fluid systems of greatly
differing geometry, the free falling planar liquid film and counterflow in a matrix of
porous cylindrical membranes, respectively. Both systems have in common that ma-
terial transport is predominantly influenced by the dynamics near the interfaces. At
the liquid/gas interface of the falling film, the occurrence of waves above a critical
threshold volume flow rate was demonstrated; transport within the wavy film was
investigated in terms of spatially dependent velocity profiles which allow the compu-
tation of the velocity field at arbitrary positions within the film. In flow through thin
porous capillaries, the efficiency of trans-membrane fluid exchange was determined
by multi-encoding techniques which allow both a statistical description of the veloc-
ity distribution and the quantification of exchange processes in comparison to purely
diffusion-governed transport.

1. Introduction

The investigation of interfaces by a variety of NMR techniques has focused during


the preceding decades mainly on their influence on restrictions of molecular motion;
contrast parameters such as lineshifts, enhanced relaxation or reduced diffusivities
were exploited to describe the interaction of molecules with interfaces and the conse-
quence for the total dynamics of the system. Most techniques demand large surface-to-
volume ratios and are consequentially often applied to internal surfaces such as those
present in porous media. However, an understanding of the dynamics at macroscopic
interfaces is of utmost importance for a wide range of applications such as reaction ki-
netics, fluid transport, evaporation/condensation processes, etc. Applying the NMR
toolbox to such macroscopic surface phenomena provides an opportunity to expand
the traditional applications of NMR to fields like engineering sciences, rheology and
biomedical research.
423
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 423-432.
© 2002 Kluwer Academic Publishers.
424
A powerful recent development of NMR pulsed-field gradient (PFG) experiments
is found in the extension of conventional imaging or diffusion measurements toward
a combination of position and displacement encoding modules within the same pulse
sequence. Such complex sequences can be applied to the multiple encoding of position
and velocity, as well as the measurement of space-dependent diffusivities or average
velocities. The influence of macroscopic interfaces on the transport properties of fluids
can thus be visualized either directly by mapping the local flow velocity as a function
of position, or indirectly by comparing statistical quantities such as the probability
density of velocities or accelerations with analytical descriptions or numerical simu-
lations of the particular boundary conditions investigated in the experiment.
The free falling film is one example of fluid transport which is of considerable indus-
trial importance. Falling films find applications in condensers, evaporators, gas/fluid
reactors and in the context of chemical extraction whenever an optimization of the
kinetics asks for the presence of large interfaces. Typical implementations of falling
film flow are annular flow along the internal surface of a tube or flow along an in-
clined plane [1] with film thicknesses usually being on the order of 1 mm or less. One
characteristic feature of falling films is the spontaneous generation of surface waves;
above a critical flow rate, waves are observed which either leave an overall laminar
flow profile, or lead to chaotic turbulence effects [1]. The frequency and magnitude
of these waves greatly influence mass transport and increase the efficiency of mixing
and reaction processes. Spatially resolved NMR velocimetry provides a unique tool
to determine the velocity distribution parallel and normal to the axis of gravity.
Optimizing the efficiency of fluid transport by the presence of interfaces is also
the goal in the case of the clinical hemodialyzer where the cleaning of human blood
is achieved by counterflow of blood and dialysate through capillaries consisting of
semipermeable membranes. Trans-membrane flow through this porous barrier has the
effect of reversing the direction of flow velocities and can be investigated by double-
encoding of velocities employing suitable multi-PFG pulse sequences [2] . Modelling
of different membrane porosities and a quantitative comparison between transverse
displacements with and without flow allow the discrimination of the relative influ-
ences of Taylor dispersion and mechanical dispersion from flow interactions with the
structured interfaces.

2. Experimental

Experiments on fal'ling liquid films were performed employing a 4.7 T 150 mm


vertical bore magnet driven by a Bruker DSX-200 console. The device used for the
generation of films is sketched in Fig. la. It is constructed of polymethylmethacrylat
(PMMA) and consists of a tube of 64 mm outer diameter which is stabilizing a 40
mm wide vertical plate of length 75 cm. The fluid is flowing on one of the surfaces
of this plate. The sensitive volume is located 50 to 60 cm below the beginning of
the plate, allowing maximum time for the equilibration of the film flow. The average
film thickness is adjustable by varying the separation of a blade from the PMMA
plate. There are two different fluid reservoirs seperated from each other in order to
425

avoid pulsatile flow components from direct connection to a recirculation pump, and
to maintain a constant pressure in the second reservoir and therefore at the position
of tpe film initiation. The device includes a loudspeaker that can be used to excite
waves of a controlled frequency in the film. However, in this study only spontaneous
waves without acoustic excitation have been investigated. As sample fluids, silicon
oils BaySilone MIO and M100 produced by Bayer with viscosities of 17 = 10 mPas
and 17 = 97 mPas, respectively, have been chosen.
All experiments on counterflow through porous membranes were performed em-
ploying a 7T horizontal 200 mm bore magnet driven by a Bruker DMX-300 console.
Mini hemodialyzer modules were used which consisted of a bundle of about 310 cap-
illaries of 200 JLm inner diameter and 30 JLm membrane thickness, sealed in a 6 mm
ID tube. Water was chosen as the flowing liquid both inside the capillaries and in
the dialysate compartment, and average flow velocities comparable to the situation
in clinical use were maintained by two separate pumps. For comparison, analogous
measurements were performed with larger clinical hemodialyzers of typically 15000
capillaries.

3. Velocity determination in falling fluid films

The occurrence of spontaneous waves, i.e. waves in the absence of coherent ex-
citation which could be achieved by a loudspeaker, depends on small fluctuations of
the velocity field and on the boundary conditions given in particular by the blade
defining the initial film thickness. For a single-component system without tempera-
ture gradients, perfectly laminar flow and the total absence of stable waves is only
expected below a critical film Reynolds number. This quantity is defined as
pvzh
Ref =- , (1)
17
where p and 17 are density and viscosity of the fluid, respectively, V z the velocity
component in the direction of gravity, and h the film thickness. The critical Reynolds
number Rec is derived from energetic considerations [3] and one finds Rec = 0.9 and
Rec = 1.9 for the silicon oils M100 and MIO, respectively. Under these conditions,
one expects a parabolic profile of the velocity [4]:
2 (x
pgh
vz(x) ="2r} 2X - hX2)
2 • (2)

The validity of this prediction can be verified by measuring the vertical (z) component
of the velocity across the film thickness for a slice in the center of the film where
boundary effects can be neglected. The velocity is encoded by a bipolar gradient
filter preceding the spin-warp imaging part of the pulse sequence, resulting in a phase
shift which is proportional to the displacement along z during the encoding time [5].
Fig. 2a shows the result of an experiment in the sub critical flow regime; a perfect
agreement with theory is observed.
With increasing Reynolds number, one expects the occurrence of spontaneous
waves in the so-called laminar-wavy regime [1]. It has been found that these waves can
426

(a) (b)

-
dialysa te out
v>O

,lw~ 6 5

Figure 1. (a) Drawing of the falling film device. The film is falling along a 40mm wide plate
(5) made from PMMA and is generated using a straight adjustable blade (4) that allows to
control the film thickness. The device contains two separated reservoirs (1, 3) and a loud-
speaker (2) which allows the excitation of waves with defined frequencies on the film. The
inset shows a transverse section through the lower part of the device with the tube (6) and
the plate (5), as well as the definition of the axes used in this work: x through the film, y
across the film, and z along the falling direction.
(b) Top: Principle of the hemodialyzer. Fluid on the blood side is distributed into membranes
of identical circular cross-section, the dialysate compartment fills the space between the mem-
brane capillaries. For the experiments of this study, both blood and dialysate compartments
were filled with water. Bottom: Schematic drawing of the cross-section of the module.

be described by a second parabolic profile superimposed onto the unchanged profile


of the residual film [6]. The NMR technique used, however, is not able to accumulate
signals from waves coherently. One rather collects signals from waves which travel at
random within the sensitive volume of the resonator during the time of the acquisition.
Still, this signal contains the correct phase information of the moving fluid under the
assumption that the relation between plate distance and velocity remains valid. This
is indeed the case for Reynolds numbers just above the critical value Rec = 0.9, where
one observes a superposition of two parabolic velocity profiles (Fig. 2b).
For higher values of Ref, the deviation from this behaviour becomes more pro-
nounced, and for the given number of averages the random waves lead to a more
disturbed profile while at the same time, the residual film becomes thinner due to
the higher amplitude of the wave undulation [7]. Fig. 3a shows the velocity profile of
V z for Ref = 2.5 along with the results for Vx and v y . There is no net displacement
normal to the flow direction, however, the scatter of data is larger in the region of the
wave.
427
(a) (b)
80 ISO

Ref =0.5 Ref = 1.0


~ro ~
E 100
§. §.
>" 40 >"
so
20

0 0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.5 1.0 1.5
x [mm] x [mm]

Figure 2. Velocity component along the flow direction, V z , as a function of the distance from
the PMMA plate for silicon oil MI00. (a) Ref = 0.5: fully laminar film with a thickness of
h ~1.0 mm. (b) Ref =
1.0: laminar-wavy film with a residual film thickness of h re • ~O . 9
mm. Parabolic functions have been fitted to the experimental data.

Rather than measuring the velocity directly from the phase shift, variation of
the intensity of the bipolar gradient and subsequent Fourier transformation allows
the determination of the probability density of displacements, or velocities, the so-
called propagator [8,9] . At the cost of replacing spatial information by a statistical
distribution function, this method is often more robust and allows a quantification of
the mass fraction of spins moving at a particular velocity. The distribution of velocities
along the flow direction, P(v z )' is presented in Fig. 3b. While the maximum velocities
obtained from both methods are comparable (~ 70 mm/s), the main difference is given
by the reduced probability of finding high velocities. The plot of V z (x) is derived from
the signal phase directly; the propagator representation, on the other hand, proves
that waves are not present during each signal acquisition but contribute with a reduced
probability. From the intensity drop at intermediate velocities one can estimate that
only one out of five acquisitions was performed while a fluid wave was present in the
sensitive volume of the resonator.
A more direct visualization of the flow process is achieved if velocity encoding
experiments are performed employing a second imaging dimension; two such experi-
ments can then be combined to obtain a spatially resolved vector plot of the velocity
distribution. An example of such a four-dimensional representation is shown in Fig. 4
where the longitudinal and the normal velocity components, V z and v x , are plotted in
a side view of the falling film. The extension along z is given by the sensitive volume
of the resonator which is about 40 mm. At this low Reynolds number (Ref = 1.5), the
presence of waves is seen only from the increased thickness of the film but streamlines
are still parallel to each other, indicating no net flow normal to the plate surface.
The situation becomes more complicated for larger Reynolds numbers, even al-
ready below the transition towards turbulent behaviour which is expected at Ref = 75
[10]. In Fig. 5, both the longitudinal and the normal components ofthe flow velocities
are plotted as a function of the x coordinate for silicon oil MlO flowing at Ref = 53.
428

(a) (b)

100 • Vz o.osor----r~__:_'~~--~__,

0.040
~
E 0.030

...
60
E ;::
';'40 Q:' 0.020

20 0.010

0.000
-20L.-.~~~_-'-_~_~---' -O.olO=-_-=,.--~-:-~_-::-_-:-:
0.0 0.2 0.4 0.6 0.8 1.0 1.2 -100 -so 0 so 100
x [mm] vz[mm/s]

Figure 3. Vertical flow of silicon oil Ml0 with Ref = 2.5. (a) Velocity components parallel
(v z ) and perpendicular (vx, v y ) to the flow direction as a function of the distance from the
PMMA plate. (b) Probability density of velocity components along the flow direction, P(v z ).

~ scale:
180mmls

dl!
E
.§.
N

0 2 4 6
x [mm]

Figure 4. Vector plot of the longitudinal and the normal velocity components, Vz and V x , for
flow of silicon oil M 10 at Ref = 1.5. The plate is drawn at the right-hand side.

A clear attribution to either the residual film or the wave region is not possible
any more, a parabolic function of velocities is not found. Moreover, a net fl~w in the
direction normal to the plate seems to be present within the slice of 5 mm thickness
in the center of the film which has been encoded in this experiment (note the different
orders of magnitude of the velocity amplitudes). Because no total flux in this direction
is allowed due to conservation of mass, a distribution of regions of predominantly
upward and downward motion across and along the film is to be expected, a feature
which can only be explained by the presence of large-scale, three-dimensional waves.
This notion is corroborated by the strong dependence of the velocity distribution on
the width of the excited slice, corresponding to the averaging volume.

4. Counterflow in hemodialyzers

Similar to the situation in film flow, the nature of the transport process through
429
(a) (b)
..
~
E
400
Uj'
E 6646666
..-.
-
3
.s .s
300
-
>" 200 >" 2
..
-
----
100
o
o
~O O~ OA O~ O~ 1~ 0.0 0.2 OA 0.6 0.8 1.0
x [mm) x [mm)

Figure 5. Vertical film flow of silicon oil MI0 with Ref = 53. (a) Velocity component parallel
(v z ) to the flow direction as a function of the distance from the PMMA plate. (b) As in (a),
but for the velocity component perpendicular (v x ) to the flow direction.

a structured system is determined by the value of the Reynolds number. The film
thickness in eq. (1) is replaced as the characteristic length scale by the pore or
capillary diameter. For the velocities investigated in this study of several mm/s,
Reynolds numbers of the order of unity are observed while the critical value for the
onset of turbulent flow is Rec = 2300 for a cylindrical pipe. Flow within the capillaries
of radius R is thus strictly laminar and one observes a parabolic velocity profile

v(r) = vmax [l- (r/R)2] , (3)

leading to a probability density of displacements which is constant between v = 0 and


v = V max and only broadened by additional random contributions of self-diffusion. In
the presence of impermeable walls, Taylor dispersion due to radial self-diffusion would
be the only mechanism leading to a change of the longitudinal velocity. For example,
the rms displacement of water during an interval of 50 ms amounts to 14 /.Lm which is
still relatively small compared to the capillary radius R = 100 /.Lm so that only small
changes of velocities are to be expected.
One method to determine the change of velocities during a given mixing time T m
is the velocity exchange spectroscopy experiment [2] which consists of two steps of
velocity encoding by pairs of magnetic field gradients; independent variation of the
strength of each gradient pair allows, after Fourier transformation, the presentation of
the joint two-time probability density of velocities before and after the mixing time.
The results of such a VEXSY experiment are shown in Fig. 6a. At first sight,
the fluid velocities remain mostly unchanged as expected, a fact which can be derived
from the lack of off-diagonal intensity in the plot of VI vs. V2 which would indicate
changes in the velocity component along the flow direction. A projection onto the
secondary diagonal, however, renders directly the distribution of velocity differences
[10,11] which are clearly non-negligible (see Fig. 6b) .
The average displacement spectrum is obtained from a projection onto the main
diagonal of this plot and is shown in Fig. 6c. On the blood side, a rectangular
propagator is observed which is in agreement with the notion of laminar (Poiseuille)
430

(a)

'm=40 ms

(b) (c)

)\,
.1].s ·16.1 0.0 IU l1J
average ICGewation (mmls1 aver9gO veloclty [mmI'.$)

Figure 6. Counterflow of water in a clinical hemodialyzer module. (a) VEXSY plot of


longitudinal velocities for two different mixing times; initial and final velocities are shown
along the two axes in the plane, respectively. (b) Distribution of velocity changes as obtained
from a projection onto the secondary diagonal of the VEXSY plot. (c) Distribution of average
velocities as obtained from a projection onto the main diagonal of the VEXSY plot.

flow with identical maximum velocities for all capillaries. The exponentially decaying
propagator for counterflow in the intercapillary space is characteristic for randomly
structured media with a distribution of pore sizes and cannot be described analytically.
In order to estimate the influence of trans-membrane flow on the shape of the 2D
VEXSY spectrum, numerical simulations have been performed for comparable condi-
tions, but assuming cylindrical geometries in both directions for simplicity. Molecules
hitting the wall during the mixing time are allowed to cross the membrane with a
probability c which leads to an inversion of their flow direction. The results of the
simulations are shown in Fig. 7. An increased broadening is observed for the larger
mixing time; this effect is even observed for the case of impermeable walls (c = 0.0)
and is a consequence of Taylor dispersion due to radial self-diffusion. For finite proba-
bilities to cross the membrane wall (c > 0), the broadening becomes more pronounced
and is particularly marked for small velocities. This contribution is brought about by
particles remaining in the vicinity of the wall where the probability of wall collisions,
and thus also of a change of the flow direction, is largest.
The agreement between simulation and experiment is reasonably good if only the
blood sides (lower left quadrant in the experimental plots) are compared (see Fig. 8).
This remains true for all investigated mixing times up to Tm ~ 100 ms. However, the
deviations are most obvious for short mixing times. In this case, the experimental
431

£=0 ,0 £=0.3 e=1.0 t m= 45 ms

r,5,'
r
,5,"
>"
-. >"
-,

-. ,
v, (mmls)
-, "
v, (mmls)
-, ,
v, (mmls)
t m =135ms

.. ,
E
,5"
~
,5"
.
,;- >"
-, -, -,
-, -, -,
v, (mmls) v, (mmls) v, (mmls)

Figure 7. Numerical simulations of a VEXSY experiment with 500000 particles undergoing


counterflow in cylindrical capillaries. The flo w parameters were chosen comparable to the
experimental conditions ('lima", = 2.5 mm/s, R =0.1 mm, assuming the self-diffusion coeffi-
cient of water). Particles hitting the tube wall were allowed to cross the wall and change the
sign of their ve/ocity with a probability c .

(a) (b)
20 20

10 10
:wE :w
~
->
0 IO
->
-10 -1 0

-20
tm =11 ms -20
tm =11 ms
-20 -10 0 10 20 -20 -1 0 0 10 20
v, [mm/s) v, [mm/s )

Figure 8. Comparison of VEXSY joint probability densities of velocities from (a) experimen-
tal results, and (b) numerical simulations as described in Fig. 7. Arrows point to regions
where a sign change of the veloci ty, equivalent to trans-membrane flow, can be identified.

VEXSY results differ from the simulated ones by additional off-diagonal intensities.
The arrows in Fig. 8a point to regions where a sign change between VI and V2 is
observed. This is not found, as expected from the simulations, symmetrically around
zero velocity, but rather for a particular range of finite values of either VI or V2.
The only possible conclusion is that diffusive processes cannot exclusively account
432

for trans-membrane exchange but that additional convective motion contributes to


the net mass transport between the two counterflowing compartments. Exchange is
enhanced due to mechanical dispersion at the structured interface and the exchange
efficiency of the dialyzer is increased.

5. Summary

Combinations of NMR k-space and q-space encoding techniques have been applied
to two flowing systems of greatly varying geometry, namely the free falling liquid film
and counterflow in a hemodialyzer. Both systems are examples of a new class of appli-
cations of engineering and biomedical importance where understanding the influence
of interfaces is vital for an optimization of mass transport parameters. Correlating
velocity information with spatial resolution or monitoring the evolution of the velocity
as a function of time, both options being achieved from suitably designed multi-PFG
experiments, were found to represent powerful means to quantify the properties of
complex flows.

Acknowledgments

We are grateful to Klaus KupferschUiger for the construction of the film device and
to Peter Bliimler for developing the concept of these projects. The work beoofitted
from helpful discussions with Ute Gorke, Christa Gehlen and Oliver Marseille. Fund-
ing by the Deutsche Forschungsgemeinschaft (SFB 540) is gratefully acknowledged.

References

1. Alekseenko, S. V., Nakoryakov, V. E., and Pokusaev, B. G. (1994) Wave flow of liquid
films, Begell House, New York.
2. P. T. Callaghan and B. Manz (1994) Velocity exchange spectroscopy, J. Magn. Reson. A
106, 260-265.
3. Kapitza, P. L. (1948) Wave flow of a thin layer of viscous fluid, Zhurn. Eksper. Teor. Fiz.
18,3.
4. Nusselt, W. (1916) Die Oberflachenkondensation des Wasserdampfes, Z. VDI60, 541.
5. Fukushima, E. (1999) Nuclear Magnetic Resonance as a tool to study flow, Annu. Rev.
Fluid Mech. 31, 95-123.
6. Adomeit, P., Renz, U. (2000) Hydrodynamics of three-dimensional waves in laminar
falling films, Int. J. Multiphase Flow 26, 1183-1208.
7. Tibatong, H. (1955) Uber die Unterscheidung der Wellenformen in frei flieBenden Gewiissern,
Z. Nat. Pump. 24, 234-252.
8. Karger, J., Heink, W. (1983) J. Magn. Reson. 51, 1.
9. Callaghan, P. T . (1993) Principles of Nuclear Magnetic Resonance Microscopy, Clarendon
Press, Oxford.
10. Ishigai, S., Nakanisi, S., Koizumi, T., Oyabi, Z. (1972) Hydrodynamics and heat transfer
of vertical falling liquid films, Bull. J8ME 15, 594.
11. Bliimich, B. (2000) NMR imaging of materials (2000), Clarendon Press, Oxford.
12. Han, S., Bliimich, B. (2000) Two-dimensional representation of position, velocity and
acceleration by PFG-NMR, Appl. Magn. Reson. 18, 101-114.
PULSE GRADIENT SPIN ECHO MEASUREMENT OF FLOW
DYNAMICS IN A POROUS STRUCTURE: NMR SPECTRAL
ANALYSIS OF MOTIONAL CORRELATIONS

JANEZ STEPISNIK, ALES MOHORlC AND ANDREJ DUH


Physics Department, FMF, University of Ljubljana, and J.
Stefan Institute, Ljubljana and University of Maribor, Faculty
of Electrical Engineering and Computer Science, Institute of
Mathematics and Physics, Maribor, Slovenia

1. Introduction

NMR measurement of spin migrations by the spin-echo [I] and magnetic


resonance microscopy [2] have important implication on understanding of
the molecular transport in porous media. Among various approaches, the
analysis of self-diffusion and flow through a porous structure by Modulated
Gradient Spin Echo method (MGSE) demonstrated that a properly shaped
gradient sequence can be is a powerful non-invasive probe for study of
molecular dynamics by measuring the low frequency features of the velocity
self-correlation function (VCF) [3, 4]. The method provides new informa-
tion that might be relevant to a wide range of scientific, technological and
medical inquiries such as oil reservoir appraisal and management, aquifers
behaviour, distillation and filtration processes, heterogeneous catalyst bed
design and performance, ion channelling through membranes, cell migration
in biological processes etc.
Although the experiments have confirmed a great potential of spec-
tral analysis in revealing details of translation motion through a porous
structure, it also opened a puzzling problem, when treating results of flow
measurement through the porous structure by the two Pulse Gradient Spin
Echo (PGSE) sequence in a traditional way. Namely, in contrast to the
MGSE measurement, which shows an increase of the effective diffusion
constant with the flow, it decreases when measure by the PGSE sequence as
shown in the figure 1. This phenomena can be explained by the analysis of
433
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Col/oid and Inteiface Science. 433-444.
© 2002 Kluwer Academic Publishers.
434

the spectral relation between the gradient spin echo NMR and the molecular
motion in porous media as follows.

2. Spin echo and restricted self-diffusion

Whenever in NMR a non-uniform magnetic field is used to encode the spin


magnetization for motion rather than position, it is appropriate to refocus
any spin phase shift due to absolute spin position by means of a spin echo,
so that the time integral of the effective gradient G{t) is zero. Therefore
we can write small perturbations of spin-echo phase, due to molecular dis-
placements in the non-uniform magnetic field, as O{r) = J;
F{t) v{t) dt,
where F{r) = ,J; G{t)dt and r is the time of phase refocusing. The
propagators of Fick's diffusion equation are commonly used to average
the O( r) fluctuation [5, 6, 2]. Although the Fick's diffusion equation may
describe a process on a cruder level as needed for molecular motion in small
compartments, this method provides quite satisfactory explanation of the
spin echo attenuation in porous media when PGSE sequence with a sharp
gradient pulses is applied, 8 «: ~. However, the spin phase average with
diffusion propagator is not unique and other ways are possible, in which the
molecular motion can be described in a different manner. Namely, according
to theorems of probability theory, a stochastic process is fully described
either by the probability distribution function, i.e. the propagator, or by
the correlation functions [7]. An exact distribution function or all correla-
tions of spin phase fluctuation are hardly ever available and one has to use
approximations. Since NMR does not detect the individual spin, .but rather
a coherent superposition of small signals that arise from the induction of
immense number of spins ( » 106 ), the spin phase fluctuations can be
considered as a Gaussian process [7], if certain conditions are fulfilled. In
the case of the gradient spin echo, the molecular mean free path l must be
short compare to the length of gradient spin phase-grating, F.l «: 1 [8] .
This condition is met at the most practical applications of the gradient spin
echo for the diffusion measurement. The average of spin phase fluctuations
by the cumulant expansion method in the Gaussian approximation, permits
to neglect all cumulants higher than the second moment. Thus, the spin
echo attenuation is related to the VCF (V(tl)V(t2)) [9, 10]. With the Fourier
transforms of the VCF tensor V(w), and the Fourier transform of spin
dephasing F(w, r) = J; F(t)eiwtdt the spin attenuation is written as

(3j(r) = - 11
IT 0
00
F(w,r)·1J(w)·F(w,r)dw (1)
435

.... 4
":'
1/1
1: 3
'I'

..• 2
()
~
....... PGSE (WGSE-Z)
Q
--WGSE-SS

0
0 20 40 60 80 100
Flow [mIIh]

Figure 1. The effective self-diffusion coefficient of flow through porous media as mea-
sured by MGSE sequence at the frequency 400 Hz and PGSE (MGSE-Z) sequences with
the gradient perpendicular to the flow.

Here the overlap of the gradient spectrum F(w, T) and the motional spec-
trum V(w) can be used to probe the molecular dynamics as shown in the
references [3, 4].

2.1. VELOCITY CORRELATION FUNCTION

In simple fluids without restriction to motion, the velocity correlation func-


tion decays exponentially to zero over the correlation time Tc ~ 10- 12 -
10- 1°8, which corresponds to the average collision time of molecules. The
resulting diffusion spectrum is relatively constant for low frequencies and is
decreasing for frequencies above w ~ Tc- 1 • Since T c- 1 is much greater then
the highest frequency component of the gradient modulation, Tc « 6, the
Torrey's formula [11] can be used to describes the spin echo attenuation.
However, the theory and the computer simulations of fluid hydro-dynamics
reveal the existence of slow molecular motion that appears as a long time
tail of the velocity correlation function superposed on the fast exponential
decay. This occurs with the molecular motion in the complex fluids, in
confined fluids as the characteristic negative decay, in entangled polymers
as the tube disengagement times etc. These times are more closely related
to the structure than to local motion of molecule, and may correspond to
the low frequencies regime accessible by NMR.
The delta function is a reasonable approximation for the VCF of re-
stricted diffusion only when the rate of intermolecular collisions is much
436

higher than the rate of molecular impacts with walls, Tc « t « Tw. For
longer times, the calculation of VCF with the probability distribution func-
tion from Fick's diffusion equation provides a better approximation [12] .
With a full set of characteristic eigen functions '¢k(r), which characterizes
the motional restriction, the generalized form of probability distribution
function can be written as P(r, tiro, to) = L:k '¢k(r)'¢k(r o)e- k2 Dlt-tol. In
the case of isolated pore, it provides the average of VCF spectrum over the
pore volume as

Drest(w) = D(2: Bkk2 1 +T;k~2


k Trkw
2) (2)

Here D is the constant of unbounded diffusion, Tr,k = kiD are the character-
istic correlation times for the restricted motion and Bk are the parameters
of porous structure

Bk = ~ [ [[f. (r - ro)f'¢k(r)'¢k(r o)) d3r d3 r o , (3)


ivJv o

where f is the unit vector along the applid gradient. An identical re-
sult is obtained by solving the Langevin equation for diffusion between
parallel planes [13]. The low frequency limit of VCF is the long time dif-
fusion constant, D co , which tends to zero in the case of isolated pores,
liIIlw->O D(w) => o. It imposes the following condition: L:k Bkk2 = 1.
In the case of diffusion in structures of interconnected pores, with the
tortuosity constant a defined by the ratio between the long range fluid
diffusivity Dco = D p , and the local molecular self-diffusion coefficient D, the
low frequency limit of the diffusion spectrum has to be liIIlw->o D(w) => D.a.
It changes the structure terms in a way that L:k Bkk2 = 1 - a.Thus, the
characteristic VCF spectrum of motion in the structure of interconnected
pores is

Drest(w) = D(a+ 2: Bkk21+T;k~2


k TrkW
2). (4)

The spectrum at the zero frequency is shifted upward for Dp = D.a, but
retains the characteristic lowering in the proximity of zero frequency. Since
the spectrum of spin dephasing, created by the PGSE sequence, exhibits a
peak around zero frequency as well, Torrey's formula needs to be replaced
with the more generalized expression, where the effect of VCF on the spin
echo time dependence is taken into account.
437

2.2. SPIN-ECHO AND VELOCITY CORRELATION

However, Eq.1 can be used to describe the spin-echo of restricted diffusion


in the system of interconnected pores as long as the diffusion displacement
is short or comparable to the inter-pore distance. When the displacement
is much larger that spin jumps between pores, in addition to the effect
of motional correlation, the pore structure imposes a restriction to the
spatial definition range of spin variables. It means that prior performing
the average of spin phase, one has to incorporate the spatial limit of spin
phase structure created by applied gradients (Le. the phase grating). It
can be described by the discrete Fourier components Sk(F a) in the inverse
space, where the wave vectors k defines the volume occupied by spins and
F a is an effective spin dephasing created by app1ied gradients. As shown in
reference [14], it gives the spin echo in a very general form as
1 2 2
E(r) = I: EOjSk.(Faj)ei(Faj - k)rj(O)e -'2 k Rgj(r) , (5)
j,k

where j denotes the summation over the sub-ensembles of spins for which
the effect of applied fields may be different. In the case of PGSE with narrow
gradient pulses Fa = ')'G8, but for the sequences with finite pulses or long
gradient waveform, the general form of effective spatial spin dephasing (or
phase grating) is

Fa = f (6)

Here R~(r) is the mean squared displacement along the gradient [14,8].
In the short time limit, Eq. 5 reduces to that for the unbounded diffusion
122
E(r) ~ Eoe -'2 Fa (r)R g(r) = E oe-{3(r) , (7)
where information about a restricted molecular motion is hidden in {3(r).
At long displacements, when R~(r) becomes much larger than the pore
size, only the zero-th Fourier component of the spin phase structure is
retained in Eq. 5, giving the spin echo attenuation independent of time
(8)
which is for a weak gradient, Fad « 211'
{3(r)
- R2(r) Bo
E(r) ~ e 9 , (9)
438

MGSE-Z

l!

PGSE
time frequency
Figure 2. MGSE-SS, PGSE and MGSE-Z sequences and their dephasing spectra.

where Bo is defined by Eq.3. It can be considered as an averaged second


moment of the pore volume and its interconnections along the direction of
applied gradient.
The spectrum of two-pulse PGSE gradient sequence

(10)
is dominated by the zero-frequency lobe with the width 1/6.. Therefore,
it is unsuitable to extract high-frequency information about D(w). In the
case of restricted diffusion, it covers the range of the VCF spectrum that
has a dip in the proximity of w = O. It means that Torrey's formula has to
replaced by the expression for (3(6.) in Eq.l, in which the spectral overlap
is taken into account. For the PGSE sequence with the finite pulse widths,
the calculation gives the spin echo attenuation in the form

It simplifies in the limit of narrow gradient pulses as


439

We need to emphasize that these result differs from the known relations
with respect to the term with Do:. It also confirms a breakdown of the
narrow pulse approximation [15] for 0 ~ Trl = TTl which imposes the upper
limit to the gradient pulse width 0 « T r , to where the spin echo attenuation
is in proportion to the mean squared displacement of particles.
If the structure factor in Eq.3 is written as Bk = Cf! k 4, because
the constants Ck exhibit a weak dependence on wave vector at least for
simple geometries (planar, cylindrical and spherical) [18], we can obtain
the early-time dependence of the spin-eco. Since the terms in the sum of
Eq.12 monotonically decrease with the increasing k, the Cauchy formula
permits to substitute the summation with the integration. In the short
time approximation (~ < T r ) , it gives

(13)

This result is the same as that already obtained with a more sophisticated
theories such as the "hearing the shape of drums" [16] or the probability
"returns to the origin" [17], if assuming C = 3~~ ' Here ~ is the surface-
to-volume ratio of porous media. The squared root early time dependence
has been verified experimentally in liquids and gases imbibed in a variety
of porous media as shown and quoted in the references [19].
In the intermediate regime 2D ~ ~ d2, when the collision frequency
with walls increases, all correlation times but the longest one Tr , can be
neglected. It transforms Eq.12 into
~

!3(~) = ''Y2G202[Do:~ + Bkl (1 - e Tr) + L Bk], (14)


k#l

where the spin echo attenuation exponentially approaches the linear time
dependence that has the slope proportional to inter-porous diffusion rate
Dp = Do:.
At still longer times, when R~ > d2 is getting much larger than the pore
size, the number of molecules colliding with the boundaries prevails over
those experiencing free diffusion, the discord of the spin phase structure
must be taken into account Eq. 9.

2.3. FLOW DISPERSION IN A POROUS MEDIA

In the case of the flow, the velocity field in the porous structure can be
regarded as an array of streamlines between which the molecule moves
because of the Brownian motion. Averaging over all molecules confined in
440
the matrix leads to a statistical description of the flow pattern and to the
dispersion of displacement. The dispersion of an incompressible fluid within
a porous structure in the presence of flow and self-diffusion can be described
by the convection-diffusion (or Focker-Planck) equation [20J by assuming
that the probability profile of tagging molecule becomes smoother in the
long time limit. Depending on the magnitude of mean flow velocity v, it
describes the flow dispersion with the effective coefficient D' that can be
very much greater than D of molecular diffusion. With the flow along z-
axis in a porous media, the parallel and transverse components of dispersion
tensor

~J. ~~ o 1,
D' 0
V' =
[ (15)
Dr,
are related to the zero frequency tensor of VCF spectrum as

v' = 1 00
([v - v(t)][v - v(O)]) dt = V(O). (16)

The flow dispersion obeys the convection-diffusion equation that allows us


to apply the same approach as used for the restricted self-diffusion when
analysing the spin echo experiment on the porous media. The probability
density of flow dispersion gives the transverse frequency spectrum of VCF
as before

(17)

but with the distinction that Tr,k are the characteristic correlation times
of flow dispersion, and D~ denotes the asymptotic transverse dispersion
coefficient. The measurement of the that dispersion coefficient needs to
reach a constant value might clarify the understanding of the flow dispersion
phenomena, which is important in different field of science and technology;
for example to model the pollutant transport in the ground water.

3. Measurement and discussion

The spin echo decay caused by the flow motion through the porous structure
was measured on the 300 MHz Bruker spectrometer/micro-imager, which
has a specially constructed quadrupole gradient coil. Porous media was a
column of ion-exchange resin with poly-dispersed beads (100 - 30 J.L) packed
in a capillary of 2.0 mm inner diameter. With the magnetic field gradient
of 4.5 T / m applied perpendicular to the capillary axis, the spin echo of
441

8 /

7 •• • •.. .!/)V . . . .
6 / •• .......K
S
w ,/ • xX
"0 5 /0' • )f' ..... )( .....
w r-----------,
"6i 4
, • ,)C"/
,0 X" • MGSE-Z ftow 1.1 mm/s
.Q 3 /+ )fa/

,~ ....'*/ 0 MGSE-Z ftow 0


2 / T
,If/x
,~K xMGSE-SSftow 1.1mm/s208Hz

-1 +-~~~--~--~--~~==~==~~--~
-0,05 0,05 0,15 0,25 0,35 0,45 0,55
time NT= [s]

Figure 3. The MGSE-Z attenuation with fJ = 70 /1-s, G = 4.52 Tim and the modulation
frequency above 600 Hz of stationary water and water trickling through porous structure
with the average velocity 1.1 mmls. MGSE-SS attenuation is given for comparison to
demonstrate a clear linear rise of attenuation.

stationary water, and water trickling though the media was measured.
The reciprocating piston pump is used to control the flow velocities. In
the high magnetic field, the susceptibility difference of the heterogeneous
porous structure can spoil the anticipated time dependence of the spin
echo. Commonly, these effects are drowned by the use of large external
gradients, but it also shortens spin echo decay and makes impossible to
trace a long-time inter-pore diffusion. In order to eliminate the susceptibility
attenuation, we used another method: the fast spin phase cycling with the
modulated gradient spin echo sequence [21]. For this purpose, we employed
the MGSE-Z sequence instead of PGSE. MGSE-Z sequence brings about
the identical lobe of diffusion spectrum in the proximity of zero-frequency,
as the PGSE sequence (Fig.2), but the modulation frequency has to be
high enough to remove susceptibility attenuation and simultaneously shifts
the side peaks of MGSE-Z spectrum outside characteristic frequencies of
motion. It brings about a spin echo attenuation almost identical to that of
PGSE sequence. According to previous measurement [3], the modulation
frequencies of 600 Hz is high enough to shift the side peaks in the range
that is not affected by flows as long as velocity is below 3 mm/ s. Above this
range the motional spectrum levels into a plateau equal to the local diffusion
constant D. Considering the length of the MGSE-Z sequence NT as the
interval between two pulses,.!)., the expression for the spin echo attenuation
is the same as that of PGSE sequence, but with the addition of the linear
442
TABLE I. The parameters of the fit by Eq.18,
error 5%.

I parameters I 1.1 mm/ s 2 .2mm/s


Q 0.59 0.62
Tk 88ms 46ms
Bl 1.5·1O- 10 m 2 1.25 . 1O-lOm2
Eo 7.1 · 1O- lO m 2 7.7. 1O-lOm2

term equal to h·2a262 DD.. At intermediate times, the only distinction


with respect to the PGSE attenuation is in the addition of one half to the
tutousity factor as

D.
+L
1 --
f3(D.) ,2a262[D(a + 2)D. + Bkl (1 - e T r ) BkJ. (18)
k;tl

The MGSE-Z sequence with the gradient pulse width 6 = 70/-Ls was
used to measure the spin echo attenuation as a function of the sequence
duration NT for the stationary water and water trickling with the velocity
1.1 mm/s and 2.2 mm/s through porous structure.
The result for the flow rate of 1.1 mm/s is shown in the figures 3.
The spin echo attenuation displays a clear transition from the exponential
increase into the linear time dependence, which levels into almost time
independent asymptote at long times. This time evolution agrees with the
theoretical predictions according to Eq.14 and Eq.9. The fit to Eq. 18
provides almost identical slopes of the linear part for both flows but with
a shorter Tr for faster flow as shown in Tab.I. It reads that the flow motion
enlarges the bouncing rate with the walls, while leaving the inter-pore
motion perpendicular .to the flow less affected, at least for weak flow rates.
It demonstrates that the flow dispersion is more effective within a pore
than between pores. The quantitative analysis with the fit of Eq.18 to the
experimental results provides the parameters as shown in TableI In figure
3 the spin echo attenuation obtained by the MGSE-SS sequence is added
to show the distinct contrast between its linear time dependence and the
non-linear dependence of the MGSE-Z (or PGSE) attenuation in different
time regimes.
443
Conclusion

The approach with the spectral analysis of spin and molecular motion con-
firms the non-linear time dependence of the PGSE attenuation in the fluid
trickling through a porous structure. Apparently, the flow motion shorten
the time-of-flight between boundaries, so that the spin echo decay displays
three distinct time regimes from which different properties of the porous
structure can be revealed. The experiments demonstrate that the flow as-
sisted pulsed gradient spin echo, extends the range of NMR measurements
into multi-pore length scale, and is able to disclose the parameters of the
molecular dynamics and of the porous structure at once. In combination
with the modulated gradient sequence, it can be a powerful non-invasive
probe for studying diverse porous structures, which appear in nature. The
new method may be useful to a wide range of scientific, technological and
medical inquiries such as oil reservoir appraisal and management, aquifers
behaviour, distillation and filtration processes, heterogeneous catalyst bed
design and performance, ion channelling through membranes, cell migration
in biological processes etc.

References

1. E. L. Hahn. Spin-echoes. Phys. Rev., 80, 580-94, 1950.


2. P. T . Callaghan. Principles of Nuclear Magnetic Resonance Microscopy. University
Press (Oxford), Oxford, 1991.
3. P.T. Callaghan and J. Stepisnik. Frequency-domain analysis of spin motion using
modulated gradient nmr. J. Magn. Res. A, 117, 118-122, 1995.
4. J. Stepisnik and P. T . Callaghan. The long time-tail of molecular velocity correlation
in a confined fluid: observation by modulated gradient spin echo nmr. Physica B,
292, 296-301, 2000.
5. J . Kiirger and W . Heink. The propagator representatin of molecular transport in
microporous crystallites. J. Magn. Res on, 51, 1-7, 1983.
6. A. Coy and P. T . Callaghan. Pulsed gradient spin echo nuclear magnetic resonance
for molecules diffusing between partially reflecting rectangular barriers. J. Chern.
Phys., 101 , 4599-609, 1994.
7. N.G. van Kampen. Stochastic Processes in Physics and Chemistry. North-Holland
Publishing Company, Amsterdam, 1981.
8. J. Stepisnik. Validity limits of gaussian approximation in cumulant expansion for
diffusion attenuation of spin echo. Physica B, 270,110-117,1999.
9. J. Stepisnik. Analysis of nmr self-diffusion measurements by density matrix
calculation. Physica B, 104, 35G-64, 1981.
10. J. Stepisnik. Measuring and imaging of flow by nmr. Progress in NMR spectr., 17,
187-209, 1985.
11. H. C. Torrey. Bloch equations with diffusion terms. Phys. Rev., 104,563-565, 1956.
12. A. Duh, A. Mohoric, and J. Stepisnik. Computer simulation ofthe spin-echo spatial
distribution in the case ofrestricted self-diffusion. J. of Mag. Res., 148, 257-66, 2001.
444
13. E. Oppenheim and P. Mazur. Brownian motion in system of finite size. Physica,
30, 1833-45, 1964.
14. J. Stepisnik. Spin echo attenuation of restricted diffusion as a discord of spin phase
structure. J. Magn. Res., 131,339-346, 1998.
15. L.Z. Wang, A. Caprihan, and E. Fukushima. The narrow-pulse criterion for pulsed-
gradient spin echo diffusion measurement. J. Magn. Res .. A, 117,209-19, 1995.
16. P. N. P. P. Mitra, L. M. Sen, Schwartz, and P. Le Doussal. Diffusion propagator as
a probe of the structure of porous media. Phys. Rev. Lett., 68, 3555-8, 1992.
17. M. D. Hurlimann, L. M. Schwartz, and P.N. Sen. Probability of returns to the
origin at short times: A probe of microstructure in porous media. Phys. Rev. B,51,
14936-40, 1995.
18. J. Stepisnik. Time dependent self-diffusion by nmr spin-echo. Physica B, 183,
343-50, 1993.
19. R.W. Mair, G.P. Wong, D. Hoffmann, M.D. Hurlimann, S. Patz, L.M. Schwartz,
and R.L. Walsworth. Probing porous media with gas diffusion nmr. Phys. Rev.
Letters, 83, 3324-27, 1999.
20. C. van en Broeck. Taylor dispersion revisited. Physica A, 168,677-96, 1990.
21. J . Stepisnik. Measurement of susceptibility magnetic field gradient in a porous media
by modulated gradient spin echo. Proceedings 0/ the EENe 2000, http://eenc.uni-
leipzig.de/Stepisnik2b.pd/, 00, 1-5, 2000.
MESOPOROUS TRANSITION METAL ALUMINOSILICAS:
INCORPORATION OF ALKYLPHENOTmAZINES AND
THEIR PHOTOIONIZATION

SUNSANEE SINLAPADECH AND LARRY KEVAN


Department of Chemistry
University of Houston
Houston, Texas 77204-5003

1. Abstract

Photoionization of N-alkylphenothiazines in mesoporous Me-AIMCM-41


containing ion-exchanged transition metal ions Me = Ni(I1), Fe(III) and
+.
Cu(I1) was investigated. N-alkylphenothiazine cation radicals (PC n ) are
produced by 320 nm light at room temperature and characterized by electron
spin resonance and ultraviolet-visible diffuse reflectance spectroscopy. Me-AIMCM-
41 materials are shown to be efficient heterogeneous hosts for the
+.
photoinduced formation of long-lived PCn cation radicals indicating efficient
photoinduced charge separation. Ni-AIMCM-41 shows the highest photoionization
efficiency compared to Fe-AIMCM-41 and Cu-AIMCM-41. The photoionization
efficiency depends on the metal ion type and concentration ion-exchanged into
mesoporous Me-AIMCM-41 molecular sieves. Also, as the alkylphenothiazine
alkyl chain length increases from methyl to hexadecyl, the photoionization yield
decreases.

2• Introduction

Photoionization with net electron transfer has been a subject of much


research . One interesting application involves light energy storage and
conversion. 1,2 In a photoredox system, the stored chemical energy of
photoproducts serves as an energy source to drive chemical reactions. Back
electron transfer must be minimized to achieve a net production of
photoproducts. Several photoredox systems have been designed to improve the net
. effilClency.
energy converSlOn ' 3-6
Generally, a photoredox system consists of a photosensitive electron donor and
an electron acceptor to generate a pair of radical ions? ,8 It is desirable to
extend the lifetime of the photoinduced radical ions to be able to utilize the
stored energy before back electron transfer occurs. Net photoionization
efficiency is enhanced in porous materials such as silica gels9 and zeolites 10 by
trapping the electron inside the porous materials.
445
J. Fraissard and O. Lapina (elis.), Magnetic Resonance in Colloid and Interface Science, 445-454.
© 2002 Kluwer Academic Publishers.
446

MCM-41 mesoporous silica molecular sieves were ftrst synthesized by Mobil


scientists. l1 MCM-41 materials have uniform pore sizes from 15 to 100 A,
· h can be controIIed dunng
W hIC . syntheSlS.
. 12,13 Al ummum
. can su b'stltute lor
4'

some silicon in MCM-41 to form AIMCM-41 which has a negatively


charged framework and net ion-exchange capacity.14-16The structures of MCM-41
and AIMCM-41 are hexagonal as characterized by x-ray powder diffraction (XRD)
and nitrogen adsorption techniques. 17 - 19 One advantage of AIMCM-41 materials
over siliceous MCM-41 is that they exhibit greater ion-exchange capacity and
aCI'd'lty. 15,20-23
MCM-41 and AlMCM-41 are effective catalysts for oxidation reactions24-27 and
7,28-30
are promising host systems for photoredox reactions. Previous studies
showed that MCM-41 and AIMCM-41 can achieve long-lived photoproduced charge
730
separation.' Incorporation of reducible transition metal ions into silica and
aluminosilica porous materials further impedes back electron transfer by acting as a
7,3037,38
more stable electron acceptor. '
. . metaIs suc h as tItanIum
T ransltlOn . . 7,30 , manganese 31 , copper32 or vanad'lUm33
have been incorporated into either framework or extraframework (ion-exchange)
sites of MCM-41. Titanium ion [Ti(IV)] has been successfully incorporated
into framework sites of MCM-41 and shows high photoionization efftciency for
.
Incorporate . . bl e mo Iecu Ies.7,30 However, ph
d photolOnIZa ' ..
otolOnIZatlOn 0f
photoionizable molecules in AIMCM-41 ion-exchanged with different transition
metal ions has not yet been studied.
In this work, AIMCM-41 with transition metal ions incorporated by
ion-exchange were prepared and denoted as Me-AIMCM-41 [Me = Ni(II),
Fe(III) or Cu(II)]. They are used as heterogeneous hosts for photoinduced
electron transfer from N-alkylphenothiazines (PC n , n=l, 6,10 or 16 where n is
the number of carbon atoms in the alkyl chain). The structures, pore sizes and
surface areas of AIMCM-41 with different amounts of Al were characterized
by powder x-ray diffraction and N2 adsorption before ion-exchanging with
transition metal ions. The efftciency of photoionization of N-alkylphenothiazines
incorporated into Me-AIMCM-41 was monitored by electron spin resonance (ESR) and
ultraviolet-visible diffuse reflectance spectroscopy. The experimental results show
that the photoyields depend on the nature and the amount of metal ions in
Me-AIMCM-41 materials. The photo ionization of N-alkylphenothiazines with
different alkyl chain lengths was also compared.

3• Experimental Section

Commercial trimethylammonium hydroxide (Aldrich), sodium silicate


solution (27 wt % Si02 ; Aldrich), cetyltrimethyl ammonium bromide (Aldrich). fumed
silica (Aldrich) and aluminum sulfate (Fluka) were used as received for AIMCM-41
synthesis. AIMCM-41 materials with different Al contents (Si/AI = 15. 30 and 60)
447

were prepared hydrothermally using cetyltrimethyl ammonium bromide as the organic


template following an earlier procedure. 15
Powder x-ray diffraction (XRD) patterns of AIMCM-41 were obtained with a
Philips PW 1840 diffractometer using Cu Ka radiation of wavelength 1.541 A
over the range 1.50 < 2q < 150 . Structure type, crystallinity and phase purity of
AIMCM-41 with different Al contents were confirmed by XRD and the Al
content was measured by electron microprobe analysis on a JXA-8600
spectrometer.
Nitrogen adsorption isotherms were measured at 77 K using a Micromeritics
Gemini 2375 analyzer. The volume of adsorbed N2 was normalized to standard pressure
and temperature. Prior to the experiments, samples were dehydrated at 250 °c for I h.
The specific area, ABET, was determined from the linear part of the BET equation.
34,35
The Barrett-Joyner-Halenda (BJH) method was used to determine the cumulative
pore surface area (ABJH), pore volume (VBJH) and pore sizes (DBJH) of AIMCM-41
samples. ABJH and VBJH were obtained from the pore size distribution curves
whereas DBJH was calculated from 4 VBJH I ABJH .
The ESR spectra were recorded at room temperature at X-band frequency
using a Bruker ESP 300 spectrometer with 100 kHz field modulation and microwave
power low enough to avoid saturation and distortion of the spectrum. The
+.
photoproduced PCn (n=l, 6,10 or 16) radical yields were determined by double
integration of the ESR spectra using the ESP 300 software. Diffuse reflectance
ultraviolet-visible spectra were recorded before and after different times of 320
nm photoirradiation at room temperature using a Perkin-Elmer model 330
spectrophotometer with an integrating sphere accessory. Thermogravimetric analysis
(TGA) of the samples were performed using a TGA 2050 analyzer from T A
Instruments.
Cu(N0 3 )2 (Fisher-Scientific), Ni(N0 3)2 (Acros) and Fe(N0 3)3 (Fisher-
Scientific) were used as received for ion-exchange. In order to study the effect
of transition metal ions on the photoyield, 0.5 M of metal ion solutions of
Cu(N0 3)2' Ni(N03)2 and Fe(N0 3)3 were liquid state ion-exchanged with calcined
AIMCM-41 (Si/AI = 30). The mixtures were stirred near 90 °c for I h
according to procedures described?2 To explore whether the photoyield depends
on the amount of the transition metal ion contained in Me-AIMCM-
41, different ratios of AIMCM-41 (SiiAI = 15,30 and 60) were liquid state ion-
exchanged with 0.5 M Ni(N0 3)2. The nickel ion contents in Ni-AIMCM-41 were
determined by electron microprobe analysis using a JEOL JXA-8600
spectrometer.
Methylphenothiazine was used as received from Aldrich. Hexylphenothiazine
(PC6)' decylphenothiazine (PClO) and hexadecylphenothiazine (PCI6) were
synthesized following a reported procedure.8 The ion-exchanged solids Me-
AIMCM-41 were heated at 550 0 C for I h to remove water and organic
templates. They were then transferred to a glovebox under N2 flow.
The ion-exchanged solids (Me-AIMCM-41) were impregnated with N-
448

alkylphenothiazines in a glovebox at room temperature by mixing Me-


AIMCM-41 (0.1 g) with 1 ml of 0.01 M N-alkylphenothiazines in benzene
for 30 min in the dark. The benzene was evaporated by flowing dry
nitrogen gas over the samples in a glovebox for 2 h. For electron spin
resonance measurements the solid powders were introduced into 2 mm Ld. x 3
mm o.d. Suprasil quartz tubes and sealed with parafilm.
For dlffuse reflectance measurements the samples were filled into a
cylindrical quartz sample cell (20 mm diameter by 1 mm path length) which
was evacuated to below 1 Torr and sealed with parafilm. All samples were
handled in the dark to minimize exposure to light.
Each powder was photoirradiated at room temperature with a Cermax ISO
W xenon lamp (ILC-LX1S0 F). The light was passed through a 10 cm water filter
and a Coming glass filter 7-S1 with 90 % transparency at 240 and 400 nm and a
maximum at 320 nm. The samples were rotated during photoirridiation for uniform
exposure to the light. ESR was used to detect the photoproduced N-
+.
alkylphenothiazine cation radicals. The cation radical yields (PC n ,where n
= 1,6,10 or 16) produced from the reaction were determined by ESR and diffuse
reflectance ultraviolet-visible spectroscopy.

4. Results

The XRD patterns of calcined AIMCM-41 with different Si/AI ratios and
siliceous MCM-41 are consistent with previous ,reports. 11 ,12,lS The BET surface
areas, BJH surface areas, volume absorbed (VB]H) and pore sizes of AIMCM-41 with
different Al contents all indicate mesoporous materials. The increase of Al
content in AIMCM-41 results in a decrease of the ABET and ABJH surface areas.
However, the pore sizes (DBJH) for different SiiAI ratios in AIMCM-41 are similar
(27 - 29 A) except for the highest Al content which shows a larger pore size (38
A). It is not expected that the metal ion exchanged into AIMCM-41 affects
the pore size or structure.
Methylphenothiazine and its derivatives are important molecules due to their
antioxidant and electron donor properties. The diameters of PC1, PC6, PCI0 and
30
PC16 molecules are about 6.5, 10.5, 14.5 and 20 A, respectively. This means
that the cage openings of AIMCM-41 seem big enough to be occupied
by N-alkylphenothiazines (PC n where n=l, 6,10 or 16). But thermogravimetric
analysis results (see below) indicate that this is not completely true for all
these alkyl chain lengths. Incorporation of methylphenothiazine (PC1) into Me-
AIMCM-41 [Me = Ni(II), Fe(III) or Cu(II) produces a background ESR signal
+.
of methylphenothiazine cation radicals (PC1 ). The samples are light pink
before irradiation and tum dark pink after irradiation. The pink color is
characteristic of PC1 +. cation radicals. 8 ,9 The unresolved signal with a g value

of 2.004 is assigned to PC1 +. following previous literature. 8 ,9 The ESR


+.
signal intensities of the PC 1 radical ions dramatically increase during the
449
first 10 min and reach a plateau in about 60 min by 320 nm
photoirradiation. Cu-AIMCM-41 produces the highest dark reaction
compared to Ni-AIMCM-41 and Fe-AIMCM-41. The highest net photoyield
is obtained for Ni-AIMCM-41 followed by Fe-AIMCM-41, Cu-AIMCM-41 and
AIMCM-41. The photoyields of Me-AIMCM-41 decrease during the first three
hours and then remain relatively stable for several days.
The diffuse reflectance ultraviolet spectra of AIMCM-41 (SiiAI = IS) with
+. 89
impregnated PCI at SIS nm indicate some PCI cation radicals' formed
+.
before irradiation. So, some PCI is ionized to PCI during sample preparation.
+
Additional PC I is produced by photoirradiation and the photoyield rapidly
increases during the first 10 min of irradiation by 320 nm light at room
temperature. For AIMCM-41 with impregnated PCI6 only a weak absorption at
+.
SIS nm is observed prior to irradiation. The PCI6 intensity in AIMCM-
411PCI6 slowly increases as the irradiation time increases. For Ni-AIMCM-41
(Si/Ni = 64), a weak absorption peak was also observed at SIS nm before
photoirradiation. The peak intensity significantly increases after 3 min irradiation
and then gradually increases until 20 min irradiation. This confirms that
+.
PCI is photoionized to PCI radical ions by 320 nm light. As compared
to AIMCM-41 impregnated with PCI' the net photoyield of Ni-AIMCM-41
is about 1.5 times larger than that of AIMCM-41 after 20 min. irradiation
time at room temperature.
+.
The photoyield of PCI in Ni-AIMCM-411PCI varies with different Ni(II)
concentrations and different irradiation times. The highest Ni(II) content yields the
+.
highest dark reaction prior to irradiation. The photoyield intensity of PCI increases
with increasing irradiation time. Moreover, it is also dependent on the Ni(II)
concentration in Ni-AIMCM-41. At low Ni content (0.15% Ni), the photoyield
seems rather constant after 5 min photoirradiation at room temperature. For higher
Ni contents, the photoyields still increase after 5 min and reach a plateau
after about 30 min photoirradiation. After 60 min irradiation, the photoyields
+.
of PCI in Ni-AIMCM-411PCI increase 68 % for 0.15% Ni, 106 % for
0.32% Ni and 72 % for 0.47 % Ni.
The TGA curves for Ni-AIMCM-41 (SilNi =99) with different alkyl
o
chain lengths of N-alkylphenothiazines show peaks near 50-80 C that are
assigned to weight losses due to water. The TGA curves show an additional broad
peak near 475 0 C that is assigned to oxidative decomposition of
phenothiazine derivatives inside the Ni-AIMCM-41 channels.3 6-38 Oxidative
decomposition of PCI, PC6, PCIO and PCI6 is observed near 475 0 C
+.
although the peak for PCI6 is very small.
450

5• Discussion

XRD and N2 adsorption experiments suggest that the cage sizes of


AIMCM-4l are big enough to be occupied by N-alkylphenothiazines (PC n) for
n up to 16. This conclusion is modified by the interpretation of the
broad TGA peaks near 475 °C, which are assigned to oxidative decomposition
of PCn inside Ni-AIMCM-4l,36,37 observed for all N-alkylphenothiazines.
From these data, we confirm that most of the PCl and PC6 penetrate
into the Me-AIMCM-4l channels while PClO and PC16 only partly
penetrate into Me-AIMCM-4l channels with the rest of the PClO and
PC16 likely being adsorbed on the external surface of Ni-AIMCM-41.
ESR (g = 2.004) and diffuse reflectance (515 nm) signals of
methylphenothiazine cation radicals are observed before photoirradiation. This
implies that there is some electron transfer from Pen to Me-AIMCM-4l
during the sample preparation process 7,9,30 which is defined as a dark reaction
+.
prior to irradiation. The photo yields of PCl in Me-AIMCM-4lIPCl increase
after photo irradiation based on both ESR and diffuse reflectance, indicating that
methylphenothiazine (PCl) is photoionized to methylphenothiazine cation radicals
+.
(PC1 ) by 320 nm photoirradiation.
+.
Relative mobilities of the PCn cation radicals can be estimated by
the relative resolution of the ESR spectra.8 ,9 The observed unresolved ESR
+.
signals of PCn cation radicals in Me-AIMCM-41 [Me = Ni(II), Pe(III) and
+.
Cu(U)] indiate that the mobility of PCn cation radicals is restricted
which is consistent with them being partly located inside the channels of
the Me-AIMCM-4l materials.
One can conclude that the transition metal ions act as electron
acceptors from the fact that there is more net photoionization from
PCn in the metal-containing Me-AIMCM-4l materials than in AIMCM-
41 materials. Me-AIMCM-411PCl gives higher photoyields as compared with
+.
AIMCM-411PCl' The photoyield of PCl in Ni-AIMCM-411PCl is the
largest as compared with Fe-AIMCM-4l and Cu-AIMCM-41. This suggests
that Ni(II) is the most efficient electron acceptor among Ni(II), Pe(III) and
Cu(II).
+.
The stabilities of PCI in Ni-AIMCM-41, Fe-AIMCM-41 and Cu-
AIMCM-4l are comparable. They all show higher stability as compared with
AIMCM-41. During the photoionization of Me-AIMCM-4l [Me = Ni(II),
Fe(III) and Cu(II)], it is suggested that the transition metal ions accept
electrons and are reduced to Ni(I), Fe(II) and Cu(I). Attempts to detect
paramagnetic Ni(I) were not successful apparently because of overlap with
the ESR spectra of PCn +'. However, in a similar photoionization system,39
451

ESR has directly shown the photoreduction of V(V) to paramagnetic V(IV)


indicating that V(V) acts as an electron acceptor.
+.
The photoyield of PCl in Ni-AIMCM-411PCI depends on the
Ni(II) content in Ni-AIMCM-41. The higher the Ni(II) content the more dark
reaction occurs. The photoionization efficiency of PCI in Ni-AIMCM-41
materials increases with increasing irradiation time. The highest Ni(II) content
shows an increased photoyield for a longer irradiation time (- 10 min) at
room temperature compared to the lowest Ni content (- 5 min.). The role
of transition meal ions as electron acceptors is evident since there is net
photoionization of PCn in Me-AIMCM-41 in addition to the dark
reaction yield at 0 min photoirradiation observed which varies with the
amount of the transition metal ion. The dark reaction, which is
highest for the highest Ni loading, limits the photoionization efficiency
of Ni-AIMCM-41 since the highest net photoyield is obtained for the
0.32 % Ni in Ni-AIMCM-41 which is an intermediate Ni concentration.
Longer alkyl chains on PCn decrease the photoionization efficiency
of N-alkylphenothiazines. A longer alkyl chain length makes the molecules
more bulky and more difficult to penetrate into the Ni-AIMCM-41 channels.
This is confirmed by TGA analysis in which broad peaks near 475 0 C assigned
+.
to PCn radical ions inside the AIMCM-41 channels are weaker with
increasing alkyl chain length. With a sufficiently long alkyl chain, it
is expected that the rate of diffusion of the PCn into Me-AIMCM-41
will be reduced and yield less dark reduction. This explains the
photoionization results which show decreased photoyield with longer
akyl chain length. The photoyield decreases with increasing alkyl chain
length may also be partially due to an increase of molecular aggregate
generation with alkyl chain length . 36,40 The formation of such
aggregates will further lower the rate of diffusion of PCn into Me-
AIMCM-41 and lead to diminished photoyields.

6. Conclusions

The experimental data clearly reveal that N-alkylphenothiazines


incorporated into Me-AIMCM-41 [Me = Ni(II), Fe(III) or Cu(II)] can be photoionized
+.
and form alkylphenothiazine cation radicals (PC n ). Electron transfer seems to
occur from PCn inside the Me-AIMCM-41 channels to the incorporated metal
ion according to ESR and diffuse reflectance spectroscopy. Back electron
transfer is efficiently retarded in Me-AIMCM-41. Ni(II) is the most efficient
electron acceptor among Ni(II), Fe(III) and Cu(II) ion-exchanged into
AIMCM-41. The photoyield depends on the nature and amount of the
transition metal ion-exchanged into AIMCM-41, and also on the size of
the electron donor molecules.
452
7. Acknowledgment. This research was support by the Division of Chemical
Sciences, Office of Basic Energy Sciences, Office of Energy Research, U.S .
Department of Energy, the Texas Advanced Research Program and the
Environmental Institute of Houston.

References

1. Kavarnos, G.L. and TUITo, NJ. (1986) Photoionization by Reversible


Electron Transfer: Theories, Experimental Evidence and Samples, Chern.
Rev. 86,401 - 449.
2. Connolly, J .S. (1981) Photochemical Conversion and Storage of Solar
Energy; Academic: New York.
3. Vermeulen, L.A. and Thompson, M.E. (1992) Stable Photoinduced
Charge Separation in Layered Viologen Compounds, Nature , 358,
656 - 658.
4. Nakato, T., Kazuyuki, K and Koto, C (1992) Syntheses of Inercalation
Compounds of Layered Niobates with Methylviologen and Their
Photochemical Behavior, Chern . Mater. 4,128-132.
5. Kang, Y.S ., McManus, HJ.D and Kevan, L. (1993) Comparative Electron
Spin Resonance and Electron Spin Echo Modulation Studies of the
Photoionization of Positively and Negatively Charged and Neutral
Alkylphenothiazines in Cationic Dioctadecyldimethyhlammonium Chloride,
Neutral Dipahnitoylphosphatidylcholine and Anionic Dihexadecylphosphate,
J. Phys. Chern. 97, 2027 - 2033.
6. Yonemoto, E.H., Kim, Y.L, Schmehl, R.H ., Wallin,J.O., Shoulders, B.A.,
Richadson, B.R., Haw, J.F. and Mallouk, T.E. (1994) Photoinduced
Electron Transfer Reactions in Zeolite-Based Donor-Acceptor and Donor-
Donor-Acceptor Diads and Triads, JArn .ChernSoc. 116,10557 - 10563.
7. Sung-Suh, H. M., Luan, Z. and Kevan, L. (1997) Photoionization of
Porphyrins in Mesoporous Siliceous MCM-41, AIMCM-41 and TiMCM-
41 Molecular Sieves, J. Phys . Chern. BIOI, 10455 - 10463.
8. Krishna,R.M., Kurshev , V. and Kevan,L. (1999) Photoinduced Charge
Separation of Phenothiazine Derivatives in Layered Zirconium Phosphate
at Room Temperature, Phys . Chern. Chern. Phys. 1, 2833 - 2839.
9. Xiang, B. and Kevan, L. (1994) Photooxidation of Phenothiazine
Derivatives in Silicas Different Pore Sizes, Langmuir 10, 2688 - 2693.
10. Ledney, M.; and Dutta, P.K. (1995) Oxidation of Water to Dioxygen by
Intrazeolitic Ru(bpY)33+, J. Am. Chern. Soc. 117,76877695.
11. Kresge, C. T., Leonowicz, W. J., Roth, W. J., Vartuti, J. C. and Beck, J.S.
(1992) Order Mesoporous Molecular Sieves Synthesized by a Liquid-
Crystal Template Mechanism, Nature 359, 710 - 712.
12. Beck, J. S., Vartuli, J. C., Roth, W. J., Leonowicz, M. E., Kresge, C. T.,
Schmitt, K.D., Chu, C.T., Olson, D.H., Sheppard, E.W., McCulley, S.
B., Higging, J. B. and Schlenker, J. L. (1992) A New Family of
Mesoporous Molecular Sieves Prepared with Liquid Crystal Templates,
J. Am. Chern . Soc. 114, 10834 - 10843.
453

13. Ying, J. Y., Mehnert, C. P. and Wong, M. S. (1999) Synthesis and


Applications of Supramolecular-Templated Mesoporous Materials, Angew.
Chem. Int. Ed. 38,56 - 77 .
14. Boger, T., Roesky, R., Glaser, R., Ernst, S., Eigenberger, G. and
Weitkamp, J. (1997) Influence of the Aluminum Content on the
Adsorptive Properties of MCM-41, Microporous Mater. 8,79 - 91.
15. Luan, Z., Cheng, C-F., Zhou, W. and Klinowski, J. (1995) Mesoporous
Molecular Sieve MCM-41 Containing Framework Aluminum, J. Phys.
Chem. 99, 1018 - 1024.
16. Biz, S. and White, M. G. (1999) Syntheses of Aluminosilicate
Mesostructure with High Aluminum Content, J. Phys. Chem. B 103,
8432 - 8442.
17. Anderson, M. T., Martin, J. E., Odinek, J. E. and Newcomer, J. G. (1998)
Effect of Methanol Concentration on CTAB Micellization and on the
Formation of Surfactant-Templated Silica (STS) Source, Chem. Mater.
10, 1490 - 1500.
18. Franke, 0., Schulz-Ekloff, G., Rathousky , J., Starek, J. and Zukal, A.
(1993) Unusual Type of Adsorption Isotherm Describing Capillary
Condensaion Without Hysteresis, J. Chem. Soc., Chem. Commun.
724 - 726 . .
19. Beck, J. S ., Vartuli, J. C., Kennedy , G. J., Kresge, C. T., Roth, W. J. and
Schramm, S.E. (1994) Molecular or Supramolecular Templating: Defining
the Role of Surfactant Chemistry in the Formation of Microporous and
Mesoporous Molecular Sieves, Chem. Mater. 6,1816 - 1821.
20 . Reddy, K. R., Araki, N. and Niwa, M. (1997) Generation and
Identification of Strong Acid Sites in AIMCM-41 Prepared by Gel
Equilibrium Adjustment Method, Chem . Lett. 7,637 - 638.
21. Corma, A., Fornes, V., Navarro, M. T. and Perez-Pariente, J. (1994)
Acidity and Stability of MCM-41 Crystalline Aluminosilicates, J. Catal.
148, 569 - 574.
22. Mokaya, R., Jones, W., Luan, Z., Alba, M.D. and Klinowski , J.
(1996) Acidity and Catalytic Activity of the Mesoporous Aluminosilicate
Molecular Sieve MCM-41, Catal. Lett. 37,113 - 120.
23. Mokoya, R. and Jones, W. (1997) Post-Synthesis Grafting of Al
onto MM-41 , J. Chern. Soc., Chem . Cornmun. 22, 2185 - 2186.
24. Climent, M. J, Corma, A., Iborra, S, Miquel, S., Primo, J. and Rey,
F. (1999) Mesoporous Materials as Catalysts for the Production of
Chemicals: Synthesis of Alkyl Glucosides on MCM-41, J. Catal. 183,
76 - 82.
25. Kageyama, K., Ogino, S., Aida, T. and Tatsumi, T. (1998)
Mesoporous Zeolite as a New Class of Catalyst for Controlle
Polymerization of Lactones, Macromolecules 31, 4069 4073.
26. Chaudhari, K., Das, T. K., Chandwadkar, A. J. and Sivasanker, S.
(1999) Mesoporous Aluminosilicate of the MCM-41 Type: Its Catalytic
Activity in n-Hexane Isomerization, J. Catal. 1, 81 - 90.
27 . Long, R. Q. and Yang, R. T. (1999) Selective Catalytic Reduction of
Nitric Oxide with Ethylene on Copper Ion-Exchanged AI-MCM-41
Catalyst, Ind. Eng. Chern . Res. 38, 873 - 878 .
454

28. Conna, A., Fornes, V., Garcia, H., Miranda, M. A. and Sabater, A.
(1994) Highly Efficient Photoinduced Electron Transfer with 2,4,6-
Triphenylpyrylium Cation Incorporated Inside Extra Large Pore Zeotype
MCM-41, J. Arn. Chern. Soc. 116, 9767 - 9768.
29. Cano, M. L., Cozens, F. L., Garcia, H., Marti, V. and Scaiano, J. C.
(1996) Intrazeolite Photochemistry. 13. Photophysical Properties of Bulky
2,4,6-triphenylpyrylium and tritylium Cations Within Large and Extralarge
Pore Zeolites, J. Phys. Chern . 100, 18152 18157.
30. Krishna, R. M., Prakash, A. M. and Kevan, L. (2000) Photoionization of
N-Alkylphenothiazines in Mesoporous SiMCM-41, AIMCM-41 and
TiMCM-41 Molecular Sieves, J. Phys. Chern. B 104, 1796 - 1801.
31. Xu, J., Luan, Z, , Wasowicz, T. and Kevan, L. (1998) ESR and ESEM
Studies of Mn-Containing MCM-41 Materials, Microporous Mesoporous
Mater. 22, 179 - 191.
32. Luan, Z., Xu, J. and Kevan, L. (1997) Cupric Ion Exchange into
Tubular Aluminosilicate MCM-41 Material with Variable Framework
Si/AI Ratios, Nukleonika 42,493 - 504.
33. Luan, Z., Xu, J., He, H., Klinowski, J. and Kevan, L. (1996)
Synthesis and Spectroscopic Characterization of Vanadosilicate Mesoporous
MCM-41 Molecular Sieves, J. Phys. Chern. 100,19995 - 19602.
34. Ravikovitch, P.I., Wei, D., Chueh, W.T., Haller, G.L. and Neimark, A.V .
(1997) Evaluation of Pore Structure Parameters of MCM-41 Catalyst
Supports and Catalysts by Means of Nitrogen and Argon Adsorption,
J. Phys. Chern. B. 101, 3671 - 3679.
35. Barrett, E.P., Joyner, L.G. and Halenda, P.P. (1951) The Detennination
of Pore Volume and Area Distributions in Porous Substances. I.
Computations from Nitrogen Isothenns, J. Arn. Chern. Soc. 73,
373 - 380.
36. Krishna, R. M., Chang, Z., Choo, H., Ranjit, K. T. and Kevan, L.
(2000) Electron Paramagnetic Resonance and Diffuse Reflectance
Spectroscopic Studies of the Photoionization of N-Alkylphenothiazines in
Synthetic Microporous M-Clinoptilolite (M = Na+ + K+, H+, Li+, Ni+,
K+, Ni 2+, Co2+, Cu2+) Molecular Sieves at Room Temperature, Phys.
Chern. Chern. Phys. 2, 3335 - 3339.
37. Kurshev, V., Prakash, A.M., Krishna, R. M. and Kevan, L. (2000)
Photoionization of Methylphenothiazine in Transition Metal Containing
Silicoaluminophosphates, Microporous Mesoporous Mater. 34, 9 -14.
38. Ranjit, K. T., Chang, Z., Krishna, R. M., Prakash, A. M. and Kevan,
L. (2000) Photoinduced Charge Separation of Methylphenothiazine in
Microporous Metal Silicoaluminophosphate M-SAPO-n (M = Cr, Fe and
Mn, n = 5,8, 11) Materials, J. Phys. Chern. B 104, 7981 - 7986.
39. Chang, Z., Ranjit, K. T., Krishna, R. M. and Kevan, L. (2000)
Photoinduced Charge Separation of Methylphenothiazine in Vanadium and
Titanium-Containing AIPO-5 and AIPO-11, J. Phys. Chern. B 104,
5579 - 5585.
40. Cavanaugh, J. (1959) Peroxidase Catalyzed Oxidations in Essentially
Non-Aqueous Media: The oxidation of Phenothiazine and Other
Compounds, J. Arn. Chern. Soc. 81, 2507 - 2515.
99 Tc NMR of Technetium and Technetium - Ruthenium Metal Nanoparticles

V. P. Tarasov* , Yu. B. Muravlev*, N. N. Popova** and K. E. Guerman"

* Kurnakov Institute of General and Inorganic Chemistry,


Russian Academy of Sciences. Leninskii pro 31. Moscow.
I 1999 I Russia
** Institute of Physical Chemistry. Russian Academy
of Sciences. Leninskii pro 3 I . Moscow. I 19991 Russia

The properties of metals are related to their electronic structure and crystal structure. Small clusters
of metal atoms exhibit extraordinary physical and electronic properties, caused by size effects,
namely, by the surface-to-volume ratio and discreteness of electronic levels [I). Bulk technetium
metal has a hexagonal close-packed lattice with parameters a = 2.735 and cia = 1.6047; technetium
films less than I SO A thick are characterized by a fcc lattice with a = 3.68 A [2, 3). Also, bulk
ruthenium metal has a hcp lattice with a = 2.704A and cia = 1.5809. Tc-Ru alloys are infinite solid
solutions [4). One of the most important characteristics of the metal electronic structure is the
density of states at the Fermi level N(EI' ). For the two most probable states of technetium, (4d' Ss' )
and (4d5 S/ ), the calculated N(EF) values are 12 .2S and 11.87 states/(Ry atom), respectively [S).
The bulk densities of states in Tc and Ru metal are the same (6). The e'xperimental characteristics
that retlect the metal electronic state and structure are NMR parameters, such as the Knight
isotropic shift (K), its anisotropy (K. n ), spin-lattice relaxation time (T, ), line width (ilv),
quadrupole coupling constant (CQ ), and asymmetry parameter 11 of the electric field gradient
tensor. We have recently determined these parameters for a technetium metal powder with a grain
size of 50-100 f.l.m : K = 6872 ppm, Kan = -400 ppm, (T, x 1)- ' = 3.23 s- ' K " ,CQ = 5.74 MHz, and
11 = 0 [7). For the bulk ruthenium metal at 4.2 K, the Knight shift is 4900 ppm (8). We are interested
in comparing these characteristics with those for technetium nanoparticles. Here. we present the
results of studying technetium and technetium-ruthenium oxide-supported catalysts by <)')Tc NMR
using supports with different crystal structures and specific surface areas. As is known, small
technetium mono- and bimetallic particles on different supports are active catalysts [2). Metallic
active states at the inert oxide support surface are believed to be the cause of catalytic properties of
the material fonned. In the case of bimetallic catalysts, an increase in catalytic activity (synergism)
is due to the fonnation of intetmetallic compounds.
455
J. Fraissard and O. Lapina (eds.). Magneric Resonance in Colloid and Interface Science. 455-468.
© 2002 Kluwer Academic Publishers.
456

EXPERIMENT AL
Three types of supports with basic properties were used, namely, y-AI203, MgO, and Ti0 2 (supports

are conventionally c1assitied as acid, basic or neutral (9)). The structure, specific surface SSP' and
pore size of these supports are given in Table 1.

Table 1. Characteristics of the supports

Support Structure Specific surface Pore size, A


area, S(m2/g)

i y-A1 01
1 spinel 189 ; 320 and 40
!
I

MgO fcc 46 20
I
1

Tetragonal 1-
I
Ti0 2 60% rutile 7
,I
40% anatase

The dispersion and particle size of technetium metal were determined on an EM-301
transmitting electron microscope with a resolution of 3.5 A [10,11]. Fig. la shows the electron
micrograph of a 1% Tc/y-AI 203 catalyst, prepared by impregnation; The magnification is 75000.
Light spots are technetium nanoparticles. The particles differ in size. Large particles are too
infrequent to be visible in the size distribution diagrams, but they may give a non-negligible
contribution to the NMR spectrum.
The bar diagrams of size distribution of particles indicate that on a y-A1203 support with the highest
specific surface area, the size of technetium particles ranges from 10 to 80 A (the average particle
size is 23 A for a 1% Tc/AI 20 3 catalyst) (Fig. Ib). This distribution is skewed toward larger
particles with an increase in technetium concentration. For the MgO and Ti0 2 supports with a
smaller specific surface area, the size distribution is wider, the average particle size being above 40
A. Catalysts were prepared by procedures described elsewhere [10,11]: support samples were
impregnated with aqueous solutions of NH4 Tc04 and RuCIJ· H 20 or H 2PtCI 6, dried at 80-90°C, and
reduced in hydrogen flow for 2-12 h at 700°C'. The calculated amount of the deposited metal was
-0.01-20 wt % for Tc and 1-10 wt % for Ru. Catalysts (0.7-0.8 g) were placed in Teflon tubes 10
mm in diameter and 30 mm in length for recording NMR spectra. 99 Tc NMR spectra were recorded
457

at 293 K on a Bruker MSL-300 spectrometer in a magnetic field of7 .04 T at a frequency of67.55
MHz.

80
-

60

40
B
20

0
I
0 40 80 120 160 200
A

Fig. ). (a) Electron micrograph of 1% Tc/y-AbO) catalyst, prepared by impregnation (x 75000);


(b) Approximate size distribution ofTc particles determined from TEM micrograph in the 1%
Tc/y-AhO) sample
458

The spin echo pulse sequence was used. The width of the first exciting pulse was 3.22 ~s, repetition
time 0.5 s, and the number of scans 64000 to 250000. It is important that the intensity of the signal
noticeably depends on the concentration of the initial NH4 Tc04 solution used for impregnation: for
a 0.01 wt % Tc/A1203 catalyst, a signal-to-noise ratio of -2 :1 was achieved after 2250000 scans
with a repetition time of 0.2 s. This spectrum was acquired in a week (Fig. 2).

i
I

Fig. 2. 99Tc NMR spectrum of a 0.01 % Tc/y-AI203 catalyst at 295 K, SW 250 kHz, NS 2250000,
Do=0.2 s

Thus, the plots of Knight shift and line width versus temperature were measured only for a
20%Tc!AI 20J catalyst. For the same sample, the spin-lattice relaxation time T,(l9Tc) was measured
using the standard saturation-recovery technique. The dependenc.e of the peak amRlitude of an
NMR signal on the delay time is described by a one-exponential function. At 295 K, the TI is equal
to 204 ms, which is about 200 times larger than that for the bulk metal (7). Chemical shifts were
referenced to a 0.1 M KTc04 solution as the external standard. Support samples were spherical
grains 1.5-2.0 mm in diameter.

RESULTS AND DISCUSSION


The 99Tc NMR spectra of all the catalysts under consideration showed signals in the region of

technetium metal shifts (-7000 ppm) and in the region of the external standard (-0 ppm) (ionic
fonn) (Fig. 3a).The integrated intensity of the low-field signal that arose from technetium metal was
roughly one order of magnitude weaker than that of the signal of the ionic form. In the general
459

spectrum these signals are hard to observe. The upper spectrum (Fig. 3b) shows this region in more
detail.

Fig. 3. 9QTc NMR spectra of a 5% Tc/y-Ah03 catalyst at 295 K; SW 1.7 MHz, NS 250000; (b) SW
250 kHz, NS 64000

The stronger signal from small particles has a shift of 7400 ppm. The weaker signal from
large particles has a shift of about 6950 ppm. The 99Tc NMR shift and line shape for nanoparticles
differ considerably from those for the bulk technetium sample (Fig. 4).
The shift is 7406 ppm, which is about 600 ppm larger than that for the bulk sample. The line with a
width at half-maximum of -I kHz is Lorentzian and lacks the satellite structure caused by first-
order quadrupole interactions, typical of the hexagonal close-packed lattice. For technetium foil
20~lIn thick the 99Tc NMR spectrum shows that the position of the central component is very close

to its position in Tc metal powder sample and 8 satellites are not clear due to the highly defective
crystal cell caused by a mutual consecutive mechanical treatment. The missing quadrupole structure
clearly points to the cubic lattice of the nascent technetium phase. The considerable increase in the
Knight shift can reflect a change in the density of states at the Fenni level, compared to the bulk
technetium sample with the hexagonal close-packed lattice [7). Figure 5 shows the . temperature
dependences K( 7) of Knight shifts for the bulk sample and technetium nanoparticles.
460

(oom)

Fig. 4. 99Tc NMR spectra of (a) 20% Tc/y-AI20) catalyst at 295 K; SW 500 kHz, NS 191000, Do
0.5s; (b) bulk Tc metal, SW 2.5 MHz, NS 50000, Do 0.5s .

7400 •.. ·.·.11 .........•


7300 a
7200
E
Q..
Q.. 7100
!fi
.c
(I)
7000
E
.2'
~ 8900

6800

6700

100 150 200 250 300 350 400
Temperature. K

Fig. 5. Temperature dependence of the Knight shifts of (a) 20% Tely- AhO) catalyst; (b) bulk Tc
metal.
461

28 ••• • •• ••• ••••••


24
• • •• B

~
~_ 20
~
~ 16

12

Fig. 6. Temperature dependence of the line widths of (a) 20% Tc/y-AI 20) catalyst; (b) bulk Tc
metal.

Compared to the K(7) = 7268 - l.35T for the bulk sample, the temperature. dependence of
the Knight shift for nanoparticles is noticeably weaker, K(7) = 7360 + 0.16T, and has the opposite
sign. The plots of line width versus temperature for the small particles and bulk technetium are
given in fig. 6. These dependences are similar and rather weak. For the bulk technetium sample,
K(7) is determined by d-polarization interaction on the background of temperature-independent
contact and orbital terms [5) . We may assume that for the cubic lattice also, a temperature change in
the d-polarization contribution dictates a change in the Knight shift. Since the density of d-states for
nanoparticles is smaller than for bulk samples [12), the contribution of Kd to the total shift is also
smaller in absolute value. However, the reasons behind such a weak temperature dependence K(7)
and its reverse sign for nanoparticles, as compared to the bulk technetium sample, are not
conclusively established. Tables 2-3 presents the 99Tc NMR parameters (Knight shifts for the metal,
chemical·shifts for the ionic form, line widths, and the metal-to-ionic form ratio).
462
Table 2. ""Tc NMR parameters for y-AI 20 J - supported monometallic Tc catalysts

! Annealing NMR shift, ppm NMR line width, Integrated


Tc time at Hz±5% intensity
content, 700°C, h Te metal TC04 Tc metal TC04 ratio
% K ±1.8 S±0.2 TeITc04
0.01 12 7409.9 2.8 921.5 937.0 1/14
I
O.OS 112 7411.7 3.6 121S.9 lOIS . 1 117
0.1 1 12 7411.7 3.1 1361.8
I
898 .0
I
1/4 I
I
I I 16 7409.9 2.0 1361.8 Ilos3 .6 III 7 I
2 6 7411.7 1.3 1848.2 1443.9 III 2
3 6 7409.9 0.7 1614.7 77S.1 1116
5 12 7409.9 0.2 16S3 .6 1287.8 116
j lO 2 7408.1 O.S 1848.2 1639.8 1113
I

7409.9 1.6 136S.7 771.2 1/ 10

1
1 0
20 J~ 7408.1 0.9 960 878.3 118

Table 3a. 99Tc NMR parameters for MgO-supported monometallic Ie catalysts

NMR shift, ppm NMR line width, Integrated


IT,
Annealing
time at Hz±S% intensity
content, 700°C, h Tc metal Te04 Tc metal TeO. ratio
i% K ±1.8 S±0.2 Tc/TcO.
!I 12 7402 .7 -6.0 1611.3 1171.3 118

I;
!4
12
6
12
7406.3
7400.9
7409.9
-2 .2
-8.8
- 3.8
2140.0
1945 .S
1848.2
2420.6
1834.1
2498.7
1/13
1112
1111
1 10 12 7406 .3 -S .2 301S.S 4060.S 1/8

Table 3b. 99Tc NMR parameters for Ti0 2- supported monometallic Tc catalysts

1 Annealing NMR shift, ppm NMR line width, Integrated


I Tc time at Hz±S% intensity ratio
I con- 700°C, h Tc metal Tc04 Tc metal Tc04 Te/Tc04
tent K ±1.8 8±0.2
1 2 7431.6 -16.3 4863.7 3434.1 1117
3 2 7401.2 -13 .7 4873.2 3746.3 114.7
S 2 7402.7 -1.1 2727.7 1874.0 I/S.S
10 2 7408.1 -1.4 2343.7 1326.8 117.7
463

These data pennit the following conclusions: (I) The K shifts are independent of the type of
support, within the experimental error; in all the samples technetium metal has a cubic structure. (2)
The line widths for the metal and technetium ionic are 1- 5 kHz; the lowest values are observed for
the y-AI20) support, and the highest values for Ti0 2. (3) The shifts for the ionic technetium form
are slightly different for the three supports. (4) The metal-to-ionic fonn ratio slightly depends on the
initial technetium concentration, the largest ratio being observed for the Ti0 2 support. The high
content of the ionic fonn in the catalyst studied evidently points to the incomplete reduction of the
initial salt to the metal, presumably, because of the "capsulation" effect. The stepwise application of
the salt to a support would thus be expected to increase the metal content of the catalyst. To verify
this hypothesis, we studied three catalysts obtained by layer-by-layer coating of technetium onto a
support. The initial sample was 5%Tc/y-AhO) annealed at 700°C for 6 h. This catalyst was then
impregnated two times with a 5% salt solution and reduced under identical conditions. The data in
Table 4 show that we obtained the result opposite to that expected; i.e. the content of the ionic fom)
increased. In addition, the signal due to metal particles became considerably broader and somewhat
up field shifted . This may be due to a wide size distribution of metal particles.

Table 4. 9t>Tc NMR parameters for three catalysts on y-AI20) supports

Annealing NMR NMR Line width Line width Integrated


Tc time at shifts,ppm shifts,ppm Tc-metal, Tc02, Hz intensity
content, % 700°C Tc metal Tc02 Hz ratio,
ITc/he02
5 6 7411.7 3.2 2529 1249 1/3
5+5 6 7406 .3 3.8 7805 5935 1114
5+5+5 6 7406.3 -2 .7 4683 3809 IllS
I I

The 99Tc NMR line shape for nanoparticles is represented by an asymmetric contour with a small
shoulder at the high-field wing. The degree of asymmetry and the change in line width depend on
the technetium concentration and annealing time of the catalyst. However, these changes are
irregular (Tables 2-3). The bar diagrams of distribution point to a wide size distribution of particles:
from 10 to 80 A fo r 1%Tc/y-AI20) (the average diameter is 23 A) and from 10 to 200 A for 2%
Tc/MgO (the average diameter is 40 A). The influence of nanoparticle size on the line shape and K
has been found for rhodium and platinum [12-14] . Thus, we assumed that the experimental 99Tc

NMR line is a composite one because of the size effects of nanoparticles, and decomposed this line
464

into components. Simulation of a contour with the LineSim program resulted in five components of
the experimental line shape for the 2% Tc/AI 20) catalyst, in eight components for the bimetallic
(3% Tc-I % Ru)/Ti0 2 catalyst and seven components for the 10%Tc-I 0%Ru/y-AI 20) catalyst.
Figures 7-9 show the 99Tc NMR spectra and the results of decomposition of experimental lines into
components. Each of the components has a Lorentzian shape and a width of 0.5-1.0 kHz. The area
under a component corresponds to the relative concentration of a definite technetium metal foml.
Table 5 shows the results for two Tc-Ru catalysts and a 3%Tc-3%Pt/MgO catalyst.
As can be seen, neither the type of support nor the nature of the second metal has much effect on
the 99Tc NMR line width and shift. This may be an indication of the absence of any intermetallic
compounds in the catalysts studied.

Table 5. 99Tc NMR parameters for y-AI 20)- and TiOrsupported bimetallic Tc catalysts

!I Tc
Annealing NMR shift, ppm NMR line width, Integrated
time at Hz±5% intensity
con- 700°C, h Tc metal TC04 Tc metal TC04 ratio
tent K ±1.8 8±0.2 Tc/Tc04
I
for 10% Ru-lO% Tc on y-AI 2O)
!
I

I 10 I 6 7410.2 0.9 2075.2 1281.7 1/4.5

for 1% Ru-3% Tc on Ti0 2

3 I 2 I 7395.5 I -15 .2 I4263 .8


for 3% Pt-3% Tc on MgO
I 3118.5 I 114.6

3 12 7399.4 -11.7 3440.0 4018 .5 IIl3

Consideration of the size distributions of particles in combination with the intensities and
shifts of resonance lines permits the tentative and qualitative conclusion that the smaller the
technetium nanoparticles, the larger the Knight shift (downfield shift). In a small particle, the
technetium positions are not equivalent, in contrast to the bulk sample where translational
symmetry results in equivalent technetium positions. Site nonequivalence implies that the densities
of states N, and Nd change on switching from one technetium posi.tion to another. In this case, N,
and Nd are related to the local density of states [13] . Therefore, each technetium position gives rise
to an individual relatively narrow « 1 kHz) line, and the experimental spectrum is an unresolved
superposition of these individual lines. The model for describing the Knight shift is based on the
465

concept of layer nonequivalence of atoms [14]. Each technetium layer is treated as a spherical shell
2.5-3 A thick. For the cubic lattice, the smallest particle contains 13 atoms: one atom is at the
center, and 12 atoms are at the surface. The next layer contains 42 atoms, etc. The overall number
of atoms in a particle containing (m + 1) layers is NT (m + 1) = NrCm) + Ns (m + 1), where N,(m) =
1013 m3 - 5m 2 + I 113m - 1 is the number of atoms in the interior layers, and Ns = 10m! + 2 is the
number of atoms at the surface of a layer [11]. For spherical technetium nanoparticles with an
average diameter of 20 to 40 A, the number of atoms is 100 to 2000, which corresponds to 4 to 8
layers. For each layer, the Knight shift Kn is the same. The layer with n = 0 corresponds to the
surface, the layer with n =. 1 corresponds to the subsurface layer, etc. The Knight shift for the nth
layer Kn is described by the formula [15): KII - K", = (Ko - K "') exp (-nlm ), where K"" = 7350 ppm is
the limiting shift of the technetium position in the bulk, Ko = 7430 ppm is the technetium shift at the
surface of a particle with a given diameter, m is a dimensionless constant, which has the meaning of
the depth at which the layers have distinguishable Knight shifts. With allowance for these data, the
Knight shifts KII in the layers of a five-layer particle were estimated to be K, = 7417, K] = 7410, K3
= 7397, K. = 7384, and K5 = 7365 ppm, at the average value m = 5. The calculated Kn values are
consistent with the data obtained upon decomposition of the experimental line shape (table to Fig.
7). Note that the layer model for a monodisperse sample implies that the most intense signal with
the maximal shift Ko should arise from the surface. The signals that arise from interior layers will be
less intense. The real sample is polydisperse, which may lead to a change in Ko depending on the
particle size, so that the intensity distribution is disturbed. For the binary (3% Tc-I % Ru)/Ti0 2 and
10%Tc-I 0%Ru/y-AI203, the experimental line shape points to the multi component character of the
signal (Fig. 8 and 9).
A possible reason for this observation may be the narrower individual lines due to dilution
of technetium with ruthenium . Since the width of each individual line is detennined by dipole-
dipole interaction between the magnetic moments of technetium spins, substituting ruthenium,
characterized by low natural abundances of magnetic isotopes with small magnetic moments, for a
fraction of the technetium will lead to line narrowing.
466

(a)

Peak Knight shift, Line width, Peak Signal area,


No ppm Hz intensity %
1 7431.2 1955.0 11.5 11.4
2 7411.4 1309.5 93.3 65.4
3 7399.5 892.1 27.5 13.1
4 7385.0 420.3 9.2 2.1
5 7365.6 1538.9 9.8 9.1

Fig. 7. ~9Tc NMR spectrum of a 2%Tc/y-Alz03 catalyst and its decomposition


into Lorentzian components; the NMR parameters of the components and
their intensities are given in the table.

Peak Knight shift, Line width, Peak Signal area,


No ppm Hz intensity %
I 7427.3 885.1 22.3 7.1
2 7418.1 925 .0 39.8 18.2
3 7408.2 872.1 65 .5 20.6
4 7396.2 909.2 54.5 17.9
5 7384.8 909.8 62.2 20.4
6 7378.3 791.5 29.2 8.3
7 7364.1 916.8 29 .9 9.9
8 7351.1 919.8 7.7 2.6

Fig. 8. 99Tc NMR spectrum of a binary 1% Ru - 3%Tc/TiO z catalyst and its decomposition
into Lorentzian components; the NMR parameters of the components and their intensities
are given in the table.
467

Peak Knight shift, Line width, Hz Peak intensity Signal area, %


Number ppm
1 7429.3 994.0 8.5 4.3
2 7414.5 1007.0 68 .9 35.4
3 7406.3 1037.9 56.0 29.7
4 7393.4 1071.5 17.9 9.8
5 7389.9 665.8 7.8 2.6
6 7381.0 814 .9 24.9 10.3
7 7367 .6 762.9 20.1 7.8

Fig. 9. 99 Tc NMR spectrum of a binary 10% Ru - 10%Tc/y-Al z0 3 catalyst and its


decomposition into Lorentzian components; the NMR parameters of the components and
their intensities are gi'{en in the table.

A possible reason for this observation may be the narrower individual lines due to dilution
of technetium with ruthenium. Since the width of each individual line is determined by dipole-
dipole interaction between the magnetic moments of technetium spins, substituting ruthenium,
characterized by low natural abundances of magnetic isotopes with small magnetic moments, for a
fraction of technetium will lead to line narrowing.
As was mentioned above, the spectra show the signals of the ionic form, along with the signals due
to the metal. The 99 Tc NMR chemical shift of this foml corresponds to the shift of the pertechnetate
ions, and the counter-ion may be ammonium or the positive charge of the support. To decide
between these possibilities, we studied the IH and 14N NMR spectra of a 5%Tc/Al zO] catalyst,
depending on the annealing time in a hydrogen atmosphere. We found that the integrated intensities
of IH and 14N NMR signals decreased with an increase in annealing time from 2 to 12 h at 700°C. A

IH NMR signal was observed for all the catalysts and initial supports. The signal was a two-
468

component line with a width of -4 kHz. The signal became weaker when.a sample is annealed. The
14N NMR chemical shift was 300 ppm from the signal of ammonium in an aqueous NH 4NO)
solution. Therefore, we assigned the signal of the ionic form to residual unreduced ammonium
pertechnetate. The smallest amount of unreduced NH4 TcO. was found for the TiO z support, which
may be due to specific features of its surface (the pore number and size). For Ti0 2, the specific
surface is two orders of magnitude lower than for the remaining supports (Table I). Therefore, the
latter may exhibit the "encapsulation" effect when a fraction of the initial component (NH4 Tc04) is
caught in pores and, thus, is not reduced. As follows from Tables 2-4, the content of unreduced
technetium (ionic form) exceeds the content of the metal phase roughly tenfold.
It should be noted that the question of the nature of the ionic form in the catalysts under
consideration is still open (16J, since the 99 Tc NMR shifts are the same for Tc02 and NH4TcO.
powders.

REFERENCES
I. Halperin, W.P., Rev. Mod. Phys., 1986, vol. 58, no. 3, p. 533.
2. Spitsin, V.l., Kuzina, A.F., Pirogova, G.N., and Balakhovskii, O.A., [togi Nauki Tekh., Ser ..·
Neorg. Khim.,1984, vol. 10.
3. Golyanov, V.M., Elesin, L.A., and Mikheeva, N.M., Zh. Eksp. Teor. Fiz., 1973, vol. 18, p. 5'12.
4. Zaitseva, L.L., Velichko, A.V., and Vinogradov, l.V., [togi Nauki Tekh .. , Ser. : Neorg. Khirn.,
1984, vol. 9.
5. Faulkner, J.S., Phys. Rev., 1977, vol. 16, no. 2, p. 736.
6. Van der Klink J.1. Adv. Catal., 2000, vol. 44, p. 1
7. Tarasov, V.P., Muravlev, Yu.B., and Guerman, K.E., J. Phys.: Condens. Matter 2001, vol. 13 (in
press).
8. Burgstaller A., Ebert H., Voitlander J. /1 Hyperfine Interactions, 1986, vol. 89, p. 1015
9. Stakheev, A.Yu. and Kustov, L.M., Appl. Catal. A, 1999, vol. 188, p. 3.
10. Pirogova, G.N., Popova, N.N., Matveev, V.V., and Chalykh, A.E., [zv. Akad. Nauk SSSR, Ser.
Khim., 1990, no. II, p. 2486.
II. Pirogova, G.N., Popov a, N.N., Voronin, Yu.V., et aI., Zh. Fiz. Khirn., 1990, vol. 64, no . II, p.
2933.
12. Tong, Y.Y., Yonezawa, T., Toshima, N., and Van der Klink, J.J., J. Phys. Chern., 1996, vol.
100, no. 2, p. 730.
13. Vuissoz, P.A., Yonezawa, T., Yang, D., et al., Chern. Phys. Lett., 1997, vol. 264, p. 366.
14. Bucher, J.P., Buttet, J.J., and Van der Klink, J.J., Surf Sci., 1989, vol. 214, p. 347.
IS. Makowka, C.O., Slichter, c.P., and Sin felt, J.H., Phys. Rev. B.' Condens. Matter, 1985, vol. 31,
p.5663.
16. Chinenov, P.P., Ronyi, S.l., Ershov, V.V., Mezenzev, V.A., Kryutchkov, S.V., Peretrukhin,
V.F. Radiokhimiya (Sov. Radiochem.), 1997, vol. 39, no. 3, p. 219.
SIZE EFFECTS ON THE NUCLEAR MAGNETIC RESONANCE OF SODIUM
METAL CONFINED IN CONTROLLED PORE GLASSES

V.V. TERSKIKH, I.L. MOUDRAKOVSKI, C.1. RATCLIFFE* and


J.A. RIPMEESTER
Steacie Institute for Molecular Sciences, National Research Council of Canada
100 Sussex Drive, Ottawa, Ontario, Canada, KIA OR6

C.J. REINHOLD, P.A. ANDERSON andP.P. EDWARDS


School of Chemistry, The University of Birmingham
Edgbaston, Birmingham, UK, BIS 2TT

1. Introduction

The growing demand from industry for functional nanoscale devices has stimulated a
tremendous interest in the design and manufacture of low-dimensional nanostructures,
including extremely thin semiconductor and metal wires · of nanometer dimensions
(nanowires). Since such nanowires are much smaller than the characteristic wavelength
of the valence electrons they possess unique electronic properties, which are
fundamentally different from corresponding bulk materials. These changes in the
electronic properties of small particles are often referred to as a quantum size effect
(QSE). The QSE is not only of academic interest, but is also practically important. For
example, due to the quantized nature of the electronic states in metallic nanowires their
electronic conductance can vary in a stepwise manner (quantized conductance) [1, 2], at
the same time the increased surface-to-volume ratio implies a greater impact of the
surface on the properties of the entire nanowire.
The experimental fabrication of nanomaterials has exploited several synthetic
techniques, ranging from manipulation of individual atoms through SEM or AFM [3] to
molecular beam epitaxy [4], arc discharge [5] and template synthesis [6]. The latter
utilizes organic and inorganic porous solids as templates for directed growth of
nanowires or their precursors. Whenever the size and orientation of the template pores
are under control, as in porous anodic aluminium oxide membranes (AAO), well-defined
and uniform nanowires can be synthesized either electrochemically (noble metals,
sulfides) [7, 8] or by sol-gel deposition (oxides) [6]. Such metals as mercury, gallium or
lead can be forced into a template from the liquid phase under high pressure [9], while
alkali and alkaline-earth metals can be transferred into a template through the gas phase
(vapor deposition). Among other porous templates are zeolites and controlled pore
glasses (CPG). Although the latter have a three-dimensional network of randomly
oriented interconnected channels, their mean pore size can be controlled by the
preparative procedure over a wide range, while the pore size distribution remains quite
narrow. This makes CPG a convenient matrix to simulate QSE in confined nanosized
species. In this work we have used CPG templates to prepare thin sodium filaments via
vapor deposition.
469
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 469-475.
© 2002 Kluwer Academic Publishers.
470

2. Experimental

Controlled Pore Glasses (CPGs) were supplied by CPG Inc. (USA). One sample of
porous glass was Vycor from Corning Glass Company. Structural characteristics of these
templates are summarized in Table 1. Before use, all glasses were cleaned by a
conventional procedure [10] and dehydrated in vacuum at 450°C for 48 hrs. Loading
with sodium was performed by exposing the dehydrated glasses to known amounts
(calculated for maximum filling of the pores) of sodium vapor in an evacuated quartz
vessel at 200-300°C, for about 30 min. followed by rapid cooling. Metal content in the
samples was confirmed by measuring the volume of evolved hydrogen after reaction
with water/methanol mixture. Since all samples are extremely air-sensitive they should
be handled and stored under vacuum or in an argon atmosphere. All measurements were
done with samples sealed in evacuated quartz tubes.

TABLE I. Morphology of Vycor and Controlled Pore Glasses (CPG)

D,A S {m2/g} V~ {cm3/g}

Vycor-44 44 198.0 0.22


CPG-115 115 119.5 0.49
CPG-251 251 82.1 0.96
CPG-585 585 57.4 1.66

23 Na NMR spectra were measured for static samples on a Bruker DSX-400 spectrometer
at 105.9 MHz (magnetic field 9.4 T) and a sweep width of300 kHz. A simple one pulse
sequence was applied with a pulse of 1 IlS and a relaxation delay of 0.1 s. Chemical
shifts are referenced to solid NaCI taken as zero ppm. Spin-lattice relaxation was
measured with a standard saturation-recovery technique. Temperatur.e was controlled
with a Bruker BVT-3300 unit.

3. Results and discussion

On contact with the sodium vapor at elevated temperatures the porous glass powder
assumed a dark brown to shiny black color depending on the amount of added metal,
temperature and duration of the synthesis. After prolonged heat treatment the internal
surface of the quartz reaction vessel also turned brown. The process of blackening of
vitreous silica on contact with sodium vapor is a well known phenomenon, which,
although it has been studied over several decades, is still not well understood [11]. In
earlier works the blackening was attributed to the formation of colloidal metallic sodium
particles on the glass surface [12]. However, later studies [11, 13] as well as our NMR
results favour a chemical interaction of the sodium with the glass.
Representative static 23Na NMR spectra of the samples under study are shown in
Fig.l. The common feature of these spectra is a broad line in the region of zero ppm. The
integral intensity of this line increases with reaction time and temperature. The position
of the signal suggests that the corresponding sodium species is cationic. Since the
471

pnstme porous glasses were sodium-free,

1051 A the appearance of this line indicates a


chemical interaction between the surface of
_---.-A_A
___ X24_~ ~... silica glass and the hot sodium vapor. The
likely products of such a reaction could
1095
include a mixture of sodium silicates,
nonstoichiometric silicon oxides SiO x and
NalCPG·115 silicon. The latter is clearly seen in 29Si
MAS NMR spectra as a narrow line with a

~
!l '_119_ _ ~
__ "
__
characteristic chemical shift of - 77.0 ppm
(not shown).
NalCPG· 251
23Na NMR quadrupolar nutation
experiments show that the sodium cations
1 124 ' __ are involved in strong quadrupolar
~ ./'.... NalCPG.$15
interactions and the observed broad line
represents only a central transition. Such
lines are typical for quadrupolar nuclides in
~
1124

glasses and are not narrowed by MAS at


Na metal moderate rates; we measured 3-4 kHz MAS
, spectra for selected samples, however no
1200 100 400 00400
(ppm) significant line-narrowing was observed.
FIGURE 1. Room temperature static 23Na NMR Below room temperature the position of this
spectra of sodium in controlled pore line is temperature independent. Under
glasses. heating from room temperature to 400K the
line narrows (from 9.0 to 4.3 kHz) and
shifts to higher field by as much as 20 ppm. Similar but smaller shifts to lower frequency
have been observed in bulk sodium silicate glasses [14]. It is likely that these are caused
by structural changes occurring in the glass and by increased diffusion of sodium
cations. At temperatures above ambient spin-lattice (T 1) and spin-spin (T2*) relaxation
times have been found essentially the same as reported previously in bulk sodium
silicate glasses, for example in Na2Sh07 [15]. We can suggest, therefore, that similar
sodium silicate compounds are formed in our case, and that during the course of addition
the first portions of sodium are consumed by the surface reaction, followed by formation
of metallic species as can be seen in the 23Na NMR spectra:
The very narrow 23Na NMR line measured for bulk Na metal (Fig. 1) reflects its cubic
crystalline symmetry (bcc). This line has a characteristic chemical shift of 1124 ppm
(Knight shift). A line with the same chemical shift is observed for Na/CPG-S8S . This
line confirms the presence of metal, yet it is broader than in bulk Na metal (Fig. 1). This
broadening becomes much more pronounced in Na/CPG-2S1 and Na/CPG-l1S, where
gradual upfield shifts of the line are also observed. In the 23Na NMR spectrum of
Na/Vycor-44 a very broad, low intensity line can be seen that has shifted considerably to
high field (Fig. 1).
The observed shifting, broadening and eventual disappearance of this metallic 23Na
NMR signal in sodium loaded glasses can be associated with the characteristic diameter
of the sodium filaments, which in our experiments was limited to the pore size of the
CPG templates. Similar NMR effects in small metal particles have been found for
472

example in the 207Pb NMR of lead embedded in porous glasses [16], 69CU NMR in silica-
supported colloidal copper particles [17], 19Spt NMR of small platinum particles [18],
and 109Ag NMR of silver clusters supported on silica and alumina [19, 20]. Neither 71Ga
nor 199Hg NMR signals were found in gallium [21] and mercury [22] confined within 40
- 70 A porous glasses (unless the samples were heated above the melting point when a
narrow signal from liquid metal appeared). Theoretical and practical aspects of NMR in
small metal particles have been discussed thoroughly in a recent review paper [23]. To
the best of our knowledge the results presented here represent the first systematic 23Na
NMR study of the nanosized sodium structures, although relevant conduction-electron
spin resonance (CESR) results were reported long ago [24, 25].
The reason for the observed NMR line broadening in metal nanoparticles lies in the
high surface-to-volume ratios. It is known that near a metal surface the electron density
distribution does not exhibit a sharp edge, since the surface interrupts the periodic crystal
potential felt by the conduction electrons. This results in electron density oscillations
near the surface (Friedel oscillations), and the role these play increases with a decrease
in the particle size. Recent DFT calculations [26, 27] have shown that in cylindrical
sodium nanowires, electronic density oscillations elongate over several atomic layers
from the surface, and could cover the entire nanowire if it has a diameter less than a few
nanometers. In thicker nanowires, however, the core may remain untouched by the
oscillations and thus retain the properties of the bulk material. The ratio between surface
and core regions would therefore determine the nanowire properties and QSE
phenomena as well. From the NMR point of view oscillations of the electron density
lead to oscillations of the Knight shift, since the latter is proportional to the square of the
conduction electrons' wave functions [28, 29]. Knight shift oscillations near the surface
will broaden the NMR line in small metal particles with respect to the bulk. Additional
broadening could also come from second-order quadrupolar interactions in the case of
quadrupolar nuclides.
If the broadening of the resonance line is mostly due to distribution of the Knight
shift (magnetic broadening) the linewidth should scale linearly with the magnetic field,
whereas for the second-order quadrupolar broadening an inverse field dependence is
expected. Neither pure behavior has been found for NalCPG samples in our variable
field experiments. The linewidths do increase almost linearly with the magnetic field, but
at 9.4 T, however, they are only 30-50% broader than at 4.7 T, indicating that both
broadening mechanisms could simultaneously contribute to the observed lineshape.
As might be expected, we found the largest broadening for Vycor-44 (Fig. 1), which
represents the limiting case where almost all the Na atoms are subjected to the Friedel
oscillations. The NMR line is also shifted upfield of the bulk metal position. We believe
that most of the metallic sodium in this sample is not detected by NMR because of
severe line broadening. With an increase in the filament/pore size the share of atoms in
the core region increases, leading to narrowing and shifting of the line (Fig. 1). Yet even
for the largest particles studied (CPG-585) where the Knight shift attains the bulk metal
value, the residual line broadening is still substantial. Indeed, this room temperature
spectrum originates mostly from the core region, which has a more or less regular
structure, while the line from the surface layers is obviously broadened out of detection.
At low temperatures Na diffusion is slow and exchange between the two regions is
473

negligible. We can expect that


Knight shift, K (ppm)
such exchange could become
1170 more important at high
temperature when sodium
1160
diffusion increases [30].
1150
We have measured tempe-
rature dependent 23Na NMR
1140 spectra for bulk sodium and two
sodium loaded porous glasses.
1130 The temperature dependencies of
the Knight shift are shown in
1120
Fig.2. The almost linear increase
1110
in the Knight shift for bulk
sodium as well as the sharp jump
1100 at the melting point (T m=371 K)
are due to an increase in the
200 250 300 350 400 T,K paramagnetic susceptibility of the
FIGURE 2. 2JNa Knight shift for Na and Na/CPG samples conduction electrons caused by
as a function of temperature. thermal framework expansion
[31]. It is remarkable, that below
270 K the Knight shifts are almost the same in the bulk metal and in the loaded samples.
At higher temperatures, however, marked differences occur, such as the decreasing
Knight shift in Na/CPG-2S1 (Fig.2), which is nevertheless interrupted by a step increase
at the melting point, similar to the bulk. This effect probably can be attributed to
increased diffusive mobility of the Na atoms, which can then exchange between core and
surface regions. The surface layers have a reduced Knight shift and very broad NMR
lines with respect to the bulk material. Heating increases the exchange of sodium atoms
between core and surface layers. The resulting exchange-averaged NMR line is therefore
shifted to high field and broadened relative to the bulk metal. Due to extreme broadening
in the upper surface layers the integral intensity of the line decreases with increasing
temperature as more and more sodium atoms become NMR invisible. The temperature
dependence of the Knight shift for Na/CPG-S85 is, as might be expected, intermediate
between Na/CPG-251 and bulk metal.
As with the Knight shifts, the nuclear spin lattice relaxation rates in small metal
particles and in the bulk are different. At lower temperatures T I tends towards the values
for bulk metal (recall the similar tendencies in the Knight shifts). At higher temperatures,
when diffusive exchange between core and surface becomes significant, much faster
relaxation we found implies greater contribution of competing relaxation mechanisms
including quadrupolar relaxation and relaxation due to paramagnetic species on the
sodium-silica interface.

4. Conclusion

Controlled pore glasses are convenient porous templates to prepare alkali metal
filaments via vapor deposition, with a range of diameters in the nano-regime. Care
should be taken, however, in choosing the appropriate reaction conditions, since there is
474
partial reduction of some of the glass by the alkali metal and this is exarcerbated by
prolonged heating. The variations in the 23Na NMR spectra for a range of nanosized
filamentary sodium structures in CPG reported here is a vivid example of the Quantum
Size Effect in metals. As the characteristic diameter decreases, the Knight shift decreases
and the linewidth increases drastically from that of pure sodium metal. These effects can
be explained in terms of Knight shift oscillations near the surface of the metal filaments,
the increasing ratio of surface to core atoms, and increasing exchange between these
regions as the temperature increases. Experiments to determine the conductivity of these
materials are currently underway.

5. Acknowledgement

The authors would like to thank Dr. M. Fergusson for PXRD measurements and Mr. J.T.
Bennett for expert technical assistance. This work was supported by a Cooperative
Research Project Grant from the National Research Council of Canada - British Council
Science and Technology Fund.

6. References

I. Yannouleas, e. and Landman, U. (1997) On mesoscopic forces and quantized conductance in model
metallic nanowires,1. Phys. Chern. BIOI, 5780-5783.
2. Guo, H. and Wang, J. (1998) Wires, dots, and tunnel junctions at the atomic scale, Physics in Canada,
137-145.
3. Eigler, D.M. and Schweizer, EX. (1990) Positioning single atoms with a scanning tunneling microscope,
Nature 344, 524-526.
4. Herman, M.A. and Sitter, H. (1989) Molecular bearn epitaxy, Springer, New York.
5. Ebbesen, T.W. and Ajayan, P.M. (1992) Large-scale synthesis of carbon nanotubes, Science 269,966.
6. Lakshmi, B.B., Dorhout, PK, and Martin, C.R. (1997) Sol-gel template synthesis of semiconductor
nanostructures, Chern. Mater. 9,857-862.
7. Hulteen, l.e., Patrissi, C.l., Miner, D.L., Crosthwait, E.R., Oberhauser, E.8., and Martin, e.R. (1997)
Changes in the Shape and Optical Properties of Gold Nanoparticles Contained within Alumina Membranes
Due to Low-Temperature Annealing, J. Phys. Chern. B 39, 7727-773 \.
8. Routkevitch, D., Bigioni, T., Moskovits, M., and Xu, J.M. (1996) Electrochemical fabrication of CdS
nanowire arrays in porous anodic aluminum oxide templates, J.Phys. Chern. 100,14037-14047.
9. Bogomolov, V.N. (1978) Liquids in ultrathin channels (filaments and cluster crystals), Sov. Phys. Usp. 21,
77-83.
10. Elmer, T.H. (1992) Porous and Reconstructed Glasses, Engineered Materials Handbook 4, 427-432.
I\, Lau, l. and McMillan, P.W. (1982) Interaction of sodium with simple glasses, 1. Mater. Sci. 17,2715-
2726.
12. Fonda, G.R. and Young, A.H. (1934) The A-c. sodium-vapor lamp, General Electric Rev. 37,331-337.
13. Stryjak, A.l. and McMillan, P.W. (1979) Color centre formation due to alkali metal vapour exposure and
X-ray ilTadiation of spinel transparent glass ceramics, Glass Tech. 20,53-58.
14. George, A.M. and Stebbins, l.F. (1996) Dynamics of Na in sodium aluminosilicate glasses and liquids,
Phys. Chern. Minerals 23, 526-534.
15. Sen, S., George, A.M., and Stebbins, l.F. (1996) Ionic conduction and mixed cation effect in silicate
glasses and liquids: 13Na and 7Li NMR spin-lattice relaxation and a multiple-barrier model of percolation,
1. Non-Crystalline Sol. 197,53-64.
16. Charles, R.l. and Harrison, W.A. (1963) Size effects in nuclear magnetic resonance, Phys. Rev. Lett. 11,
75-77.
17. Williams, M.l., Edwards, P.P., and Tunstall, D.P. (1991) Probing the electronic structure of small copper
particles: 63CU NMR at 1.5 K, Faraday Discuss. 92, 199-215.
18. Bucher, J.P., Buttet, l., van der Klink, J.J ., and Graetzel, M. (1989) Electronic properties and local
densities of states in clean and hydrogen covered Pt particles, Surf. Sci. 214,347-357.
475
19. Bercier, J.1., Jirousek, M., Graetzel, M., and van der Klink, U. (1993) Evidence from NMR for
temperature-dependent Bardeen-Friedel oscillations in nanometer-sized silver particles, J. Phys.: Condens.
MatterS, L571-L576.
20. Mastikhin, V.M., Goncharova, S.N., Tapilin, V.M., Terskikh, V.V. , and Balzhinimaev, B.S. (1995) Effect
of particle size upon catalytic and electronic properties of supported Ag catalysts: combined catalytic,
109Ag NMR.and quantum chemistry studies, J. Molecular Catalysis 96, 175-179.
21. Borisov, B.F., Chamaya, E.V. , Loeser, T., Michel, D., Tien, C., Wur, C.S., and Kurnzerov, Yu.A. (1999)
Nuclear magnetic resonance, resistance and acoustic studies of the melting-freezing phase transition of
gallium in Vycor glass, J. Phys.: Condens. Matter 11, \0259-10268 .
22. Borisov, B.F., Chamaya, E.V., Plotnikov, P.G., Hoffmann, W.-D., Michel, D., Kurnzerov, Yu.A., Tien, c.,
and Wur, C.S. (1998) Solidification and melting of mercury in a porous glass as studied by NMR and
acoustic techniques, Phys. Rev. B58, 5329-5335.
23. van der Klink, U. and Brom, H.B. (2000) NMR in Metals, Metal Particles and Metal Cluster Compounds,
Progress in NMR Spect. 36,89-201.
24. Winter, J (1971) Magnetic Resonance in Metals, The Clarendon Press, Oxford; and references therein.
25. Gordon, D.A. (1976) Conduction-electron spin resonance in small particles of sodium, Phys. Rev. B 13,
3738-3747.
26. Zabala, N., Puska, MJ., and Nieminen, R.M. (1999) Electronic structure of cylindrical simple-metal
nanowires in the stabilized jellium model, Phys. Rev. B 59,12652 - 12660.
27. Smogunov, A.N., Kurkina, L.I., and Farberovich, O.V. (2000) Electronic structure and polarizability of
quantum metallic wires, Physics of the Solid State 42,1898-1907.
28. Knight, W.D. (1956) Electron paramagnetism and Nuclear Magnetic Resonance in metals, Solid State
Phys. 2, 93-136.
29. Abragam, A. (1961) The Principles of Nuclear Magnetism, Oxford University Press, N.Y.
30. Nachtrieb, N.H., Catland, E., Well, J.A. (1952) Self-diffusion in solid sodium, J. Chern. Phys. 20, 1185-
1188.
31. McGarvey B.R. and Gutowsky H.S. (1953) Nuclear magnetic resonance in metals. n. Temperature
dependence of the resonance shifts,J. Chern. Phys. 21,2114-2119.
Poster presentations
EPR INVESTIGATION AT 4K OF CERIA AND Cu-Ce OXIDE
UNDER S02 AND H2 ATMOSPHERE

E. ABI-AAD*, J. MATTA, R. FLOUTY, C. DECARNE,


S. SIFFERT, A. ABOUKAiS
Laboratoire de Catalyse et Environnement E.A. 2598, Universite du
Littoral - Cote d'Opale, MREID, 145 avenue Maurice Schumann, 59140
Dunkerque, France. * abiaad@univ-littoraljr

1. Introduction

Ceria (Ce0 2) is often used as a catalyst for exhaust emissions elimination [I, 2]. The
most important function of ceria is the Oxygen Storage Capacity (OSC) that allows to
perform the oxidation and reduction reactions [3]. Moreover, nowadays, some
alternative to precious metal supported catalysts are considered. Cu-Ce-O is one of such
catalysts [2]. However, it is important to know its behaviour towards exhaust gases.
Indeed, some of them can poison catalysts such as S02 that can form metal sulphates
(Ce~(S04)3' Ce(S04b CuS04) [1] . Moreover, the reduction behaviour of catalyst is
important because exhaust gases often lead to reducing atmosphere. The aim of this
work is to study the behaviour of ceria and Cu-Ce-O catalysts towards S02 or Hz
atmosphere by Electron Paramagnetic Resonance (EPR) measurements at 4K.

2. Experimental

2.1. SOLID PREPARATION

Ceria (CeO~) was prepared by precipitation of cerium hydroxide from Ce(N03h,6HP


with a NaOH solution. The solid is calcined at 773 K for four hours under a flow of
dried air. Subsequently, copper-cerium catalyst is prepared by incipient wetness
impregnation of copper nitrate on ceria with an atomic ratio of CulCe= I. The sample is
dried at 373 K and calcined up to lO73 K for 6 hours in a flow of dry air. Samples are
denoted by Ce0 2 and ICulCelO73 (the digits preceding the atomic symbols represent
their molar ratio in the solid and the last number represents the calcination temperature
1073 K). EPR characterisation of these solids has been published elsewhere [4, 5].

2.2. ELECTRON PARAMAGNETIC RESONANCE

The Electron Paramagnetic Resonance (EPR) measurements were performed at 4 K on


a EMX BRUKER spectrometer with a cavity operating at a frequency of -9.5 GHz (X
band). The magnetic field was modulated at 100 kHz. The g values were determined
from precise frequency and magnetic field values. The calculated spectra were
simulated with the BRUKER « Simfonia » program based upon perturbation theory [6].
479
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 479-484.
© 2002 Kluwer Academic Publishers.
480

The EPR parameter values of the different signals were determined from the calculated
spectra. Simulated signals were calculated using the effective spin Hamiltonian:
J.I = ~ Hj Sj gj + Aj Sj Ij (eq. /)
where j is the component along one of the three axis x, y and z, H is the applied
magnetic field, S is the total spin electron, A is the hyperfine interaction, I is the spin
nuclear, g is the spectroscopic factor and ~ is the Bohr magneton. A polyoriented
sample EPR signal was simulated by generating 9000 random orientations of magnetic
field and by summing the corresponding 9000 absorption signals. The final signal was
obtained by performing a convolution (Gaussian or Lorentzian line shape) of each
transition line, adding all contributions, and by calculating the first derivative of the
signal; the line width ~Hj for convolution was optimized in order to obtain the best
accordance with the observed experimental values.

3. Results and Discussion

3.1. POISONING EFFECT OF S02 ON CERIA AND COPPER OXIDE PHASES

The presence of sulphur in diesel exhaust gases or particles has to be considered as a


poisoning agent for the catalysts used in soot combustion reactions. Ceria and copper
oxide have been reported to be sensitive towards sulphur dioxide [7] which implies a
deactivation of the solid and then eventual modifications of its surface properties. In
this way, Ce02 and ICuiCel073 samples were treated in a microflow reactor under S02
flow (2 L.h· l ) at room temperature for 30 minutes.

CeO, treated under SO,

Ce(SO.h calcined at 873K

~ \------.-:..--
/\ Simulated spectrum ~~
~

o 2000 4000 6000 8000 10000 12000 J4000


H (Gauss)

"Figure I." EPR spectra o/the CeO} after SO} treatment and Ce(S04h sample

After this treatment, Ce(S04h is formed on ceria surface and evidenced by thermal
analysis and Raman spectroscopy [8]. Simultaneously, the EPR spectrum exhibits a
complex superimposition of different signals "Figure I. ". An EPR line at g=2.013 is
evidenced and has been already attributed to O2 species [4]. In addition, an axial
symmetry signal (g,;.=1.965 and gli =1.942) characteristic of an interaction between
481

conduction electrons and 4f orbitals of Ce 4+ ions in the Ce0 2 matrix [4] is observed. The
EPR signal at g=4.27 is the well-known EPR transition for high spin Fe 3T ions;
electronic configuration (3d 5). Moreover, different EPR lines in the range of g,.,8.14,
g,.,2.18 and g,.,0.57 have been observed. In order to elucidate the attribution of these
signals and due to the presence of the Ce(S04h phase, revealed elsewhere [8], the EPR
analysis of a Ce(S04)2 sample has been undertaken at T=4 K "Figure 1. ". It is a known
feature that the calcination of the Ce(S04h oxide at high temperature (>773 K) leads to
the formation of Ce 2(S04h where the cerium is Ce(III) [9]. An asymmetric EPR
spectrum with gl=3.9702, g2=0.6256 and g3=0.5285 is observed for a Ce(S04h sample
calcined at 873 K "Figure I.". The EPR parameters of this signal are confirmed by
simulation and are typical for the ions having the electronic 4[1 configuration. The
intensity of the spectrum and values of EPR parameters allow us to suppose that this
spectrum is due to Ce3+ ions [10, 11]. When the ceria sample treated under S02 is
calcined at 873 K, the EPR spectrum reveals the only presence of conduction electrons
interacting with 4f orbitals of Ce 4+ and the EPR signal relative to Fe3+ ions.
"Figure 2." shows the EPR spectra of the 1Cu ICe 1073 catalyst before and after this
treatment. Compared to the spectrum of 1Cu 1Ce 1073 sample, a new signal, centred at
g=2.166 with a linewidth of LlH=280 Gauss appears in the second spectrum without
affecting the intensities of signals attributed to Cu 2+ ion monomers and dimers [2, 5] .
The new signal is better observed after a simple subtraction of the first spectrum from
the second one. A similar signal is recorded when a flow of S02 is introduced into CuO
particles at room temperature. Moreover, the same signal is detected for a CuS0 4,5HP
sample. From these results, it is then evident to attribute the new signal to the formation
ofCuS0 4 phase when the lCulCel073 catalyst is treated with S02 at room temperature
[2]. Indeed, it was already demonstrated that CuO phase is present in such a catalyst
previously calcined at 1073 K [5]. Therefore, when S02 is in contact with the catalyst,
the CuS0 4 phase is formed following the mechanism [2, 12] :
CU 2T + 20(surface) + S02(g) ~ CU 2T , S04 (ads)

Gain

1Cu1Ce1073

1Cu1Ce1073 + so,

subtraction

CuSO.,5H,O

1000 1500 2000 2500 3000 3500 4000 4500


H (Gauss)

"Figure 2" EPR spectra of the IClIICelOl3 after SO: treatment (ref2!
482

3.2. EFFECT OF H2 TREATMENT ON COPPER OXIDE PHASE

The 1Cu 1Ce 1073 catalyst reduced at 873 K under a mixture of 90 vol. % He + 10 vol.
% Hz, exhibits an EPR spectrum, recorded at T=4 K and reported on "Figure 3.". This
spectrum is the superimposition of different signals characteristic of isolated Cu z+ ions,
Cu"+ dimers and CuO agglomerates described elsewhere [5]. However, two EPR lines at
a magnetic field of -2255 Gauss and -5236 Gauss are observed for the first time and
only at 4K and they will be discussed in this work. In addition, another EPR signal
centred at g;so=6.79 can be observed on "Figure 3. /1. It consists of eight well-resolved
equidistant HFS lines (A=74 Gauss) and one high intensity transition with the same g-
factor. It has only been observed at 4K. The intensity ratio of the central transition to
the sum of the eight HFS lines is equal to -3. These results allow to assign this
spectrum to one of the fine transitions of E~+ isotopes (1=0 for evenEr and 1=7/2 for 167Er
with a natural abundance of77.1% and 22.9% respectively) [11). The ground state of
these ions (41 15/2) is, as for Ce 3+, the result of a strong spin-orbit interaction and the total
magnetic moment is equal to 1=15/2 (S=312 ; L=6). For the eightfold co-ordination,
either a r6 (E1d or r 7 (E5d doublet is expected to be the lowest energy level, depending
on the ratio of high-order cubic field terms. In our case, E~+ ions should be considered
as impurities since they were not intentionally introduced in the solid, but the chemical
analysis has confirmed the presence of such atoms.
The ground state electronic configuration of Cuo atoms is 3d 104s I. Assuming that the
sites for the Cu atoms retain the octahedral symmetry of the host lattice, the resonance
spectra of these atoms can be described by a spin Hamiltonian of (eq. /).

Cu + dimer
ICuiCel073

reduced at 873K

simulated Cuo dimer

UZ+ monomer

o 1000 2000 3000 4000 5000 6000 7000 8000


H (Gauss)

"Figllre 3. " EPR spectrlln! of the ICilICelOl3 catalyst reduced at 8l3K under a mixture of He + H_,
483

The splitting of these hyperfine levels as a function of the external magnetic field H is
described by Kasai et at. [13] for the particular case of I = 3/2. It is inunediately clear
that the spectrometer frequency v must be larger than the zero-field splitting, (1+V2) A,
in order to observe the « normal» spectrum consisting of (21+ I) hyperfine components
arising from the transition ~M s = ±I , ~M, = O. If, on the other hand, the spectrometer
frequency is less than zero-field splitting, one expects to observe only two transitions
corresponding to the (~Ms = ±I , MI = -I) and (Ms = -1/2, M, = -I ~ -1+1). Note that the
first of these two corresponds to the highest field hyperfine component of the normal
EPR transitions, while the latter can be regarded as an NMR transition. Equations
relating to the resonance positions of these transitions can be exposed on "Figure 4. "
[13]. In addition, for 3d ' 04s ' electronic configuration the EPR signals present isotropic
line shape. However, axial EPR parameters (gl =1.6 and gll=O.8) have been reported in
the literature [14] for copper atoms at an interstitial site of ZnO, having 3d94s~ electron
configuration and 2DJ/2 ground state. In these conditions, the hyperfine coupling
constant A is considerably smaller than hv and four well-resolved HFS lines are
observed for each perpendicular and parallel components.

2
1,8
1,6
1,4

> 1,2
.c
==
c::l.
I:)«) 0,8
0,6
0,4
0,2

°
° 0,2 0,4
Alhv
0,6 0,8

"Figure 4... Dependency oJthe resonance field gPHlh v Jor S = It, and I = 312 system upon the h}perfine
coupling constallt Alh v. II EPR » transitions are labelled with the nuclear magnetic quantllm number
assuming A > O.

In order to have a reliable attribution of the two EPR lines at -2255 G and -5236 G
observed in our case "Figure 3. we have undertaken the simulation of such a signal
If,

with respect to the theoretical aspects exposed above . The simulated spectrum
represents the fine structure of interacting two CUD atoms. The EPR parameters of the
simulated signal are gl =1.7078 and gll =O.8412, the zero-field splitting is D=2820G and
the hyperfine coupling constants are A1 =74G and AII =150G. In this case, the observed
484
EPR lines correspond to the two allowed transitions ~Ms = ±1 "Figure 3.". However,
the hyperfine structure, consistent on seven lines (I,0,=2x3/2=3=> 21,0'+ 1=7) for each
transition, is not revealed on the EPR spectrum due to high dipolar interaction between
copper atoms. In fact, the width of each EPR line (~H"'320 G) is considerably higher
than the hyperfine coupling constants. From these results, the EPR signal can be
attributed to Cuo dimers in Ce0 2 matrix.

4. Conclusion

Ceria and copper-ceria catalysts have been investigated by EPR at 4K after a treatment
under controlled atmosphere. An asymmetric EPR spectrum with g,=3.9702, g2=O.6256
and gj=O.5285 has been observed for a Ce 2(S04h sample. The EPR parameters of this
signal are typical for the ions having the electronic 4[' configuration (Ce 3+ ions). The
copper-cerium oxide sample, treated under S02 atmosphere, exhibits a signal, centred at
g=2.166 with ~H=280 Gauss, relative to the formation of CuS0 4 phase.
Under a mixture of He + H 2, the Cu (II) species of the copper-cerium catalyst are
progressively reduced into Cu+ and Cuo. Two transitions corresponding to the fine
structure of an EPR signal were evidenced and attributed to Cuo dimers in the ceria
matrix. All these results were confirmed by means of computer simulation.

5. Acknowledgments

The authors would like to thank the "Conseil General du Nord", the "Region Nord - Pas
de Calais" and the European Community (European Regional Development Fund) for
financial supports in the EPR and Thermal Analysis apparatus purchase.

6. References

I. Lundgren, S., Spiess, G., Hjortsberg, 0., Jobson, E., Gottberg, I., Smedler, G. (1995) Studies ill SlIIface
Science and Catalysis. 96,763.
2. Courcot, D., Abi-Aad, E., Capelle, S., Aboukai"s, A. (1998) Studies in Swface Science and Calalysis. 116,
625 .
3. Yao, H.C., Yu Yao, Y.F. (1984) JOllrnal of Catalysis, 86, 254.
4. Abi-Aad, E., Bechara, R., Grimblot, J., Abouka'is, A. (1993) Chemistry of Materials,S, 793.
5. Abouka"is, A., Bennani, A., Lamonier-Dulongpont, c., Abi-Aad, E., Wrobel, G. (1996) Colloids and
Surfaces A : Physicochemicnl and Engineering Aspects, 115, 171.
6. Weber, R.T. (1995) WINEPR SlMFONIA manuel, Ver 1.2, Brucker Instruments, Inc, Billerica, Ma and
references therein.
7. Ahltrom, A.F., Odenbrand C.U.1. (1990) Applied Catalysis, 60, 143 .
8. Flouty, R., Abi-Aad, E., SitTert, S., Aboukai"s, A., to be published.
9. Yang, Y., Rudong, Y. (1992) Thermochimica Acta, 202, 30 \.
10. Griscom, D.L. (1980) Journal of Non-Crystalide Solid, 40, 211.
II . Abi-Aad, E., Zhilinskaya, EA, Aboukai"s, A. (1999) Journal de Chimie Physique, 96, 1519.
12. Galtayries, A., Grimblot, J., Bonnelle, J.P. (1996) Surface and Interface Analysis, 24, 345
13. Kasai, P.H., McLeod, Jr., D. (1971)Journal of Chemical Physics, 55 (4),1566.
14. Hausmann, A., Schallenberger, B., Roll, O. (1979) ZeitschriJt fur Physik B, 34, 129.
CHARACTERIZATION OF MESO POROUS MATERIALS BY IH NMR

D. W. AKSNES\ L.GJERDAKERI AND L. KIMTYS2

IDepartment of Chemistry, University of Bergen, N-5007 Bergen, Norway


2Department of Physics, Vilnius University, Vilnius 2734 Lithuania

1. Introduction

NMR is a well-established method for characterising porous materials and for


investigating the behaviour of adsorbed substances [1]. The geometrical restrictions and
surfaces have a large influence on both the phase behaviour and the dynamics of the
confined molecules. Thus, relaxation and self-diffusion studies might give valuable
information on pore geometries, and on the important process of mass transport [1].
In this investigation, it is the dynamics and transformations of adsorbed organic
substances that is the major topic of interest. Thus, acetonitrile confined within two
different mesoporous silica materials, with nominal diameter 60 and 200 A, are studied
by high-field NMR, and the results are discussed with reference to the bulk material.
Acetonitrile melts at 229 K and undergoes a solid-solid phase transition at 217 K [2).
However, no structural information on acetonitrile appears to be available.

2. Experimental

2.1. SAMPLE PREPARATION

Two different silica samples supplied by Unilever, Sorbsil C60-40/60 and Sorbsil C200,
with nominal pore diameters of 60 and 200 A, were used in this work. The N2 sorption
isotherms and BJH desorption pore size distributions were provided by Unilever. The
acetonitrile (>99.7%), obtained from Merck, was re-distilled and dried over molecular
sieves. The two test samples were prepared in 5 mm o.d. NMR tubes, each filled to a
height of approximately 10 mm with the porous silica gels. Physisorbed water was
removed by drying the silica gels in the NMR tubes at 423 K in oven for 12 h. The
porous materials were then slightly overfilled with acetonitrile. The quantity to be added
was calculated from the known specific silica pore volume, and by observing when the
grains became less free-running . In addition, two bulk samples of acetonitrile were
prepared in 5 and 2 mm o.d. NMR tubes. The four NMR tubes were immediately sealed
to prevent evaporation and contamination.
485
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 485-490.
© 2002 Kluwer Academic Publishers.
486

2.2. NMR MEASUREMENT PROCEDURES

The NMR measurements were carried out at 9.4 T on a Bruker Avance DMX 400 WB
spectrometer. The 90° transmitter pulses were carefully calibrated for all samples and
temperatures. The sample temperature was maintained to within ±0.5 K using a Bruker
B-VT 2000 temperature-control unit. In order to avoid undercooling, all measurements
were done by increasing the temperature in steps of ca. 1 K after having cooled the
sample to ca. 150 K. Under the high-resolution conditions used in this work, no
background absorption from physisorbed water and silanol species was seen.
The self-diffusion IH measurements were performed with a Bruker Z-shielded 5 mm
probehead and a BAFPA 40 gradient unit generating field gradients with strengths up to
0.1 T cm· l . The diffusivities (D) of the confined samples were measured using the 13-
interval pulse sequence due to S0rland et al. [3], which largely eliminates the effect of
internal gradients caused by susceptibility changes throughout the sample. All diffusion
measurements in the liquid state of the bulk sample were performed on the 2 mm o.d.
NMR sample to avoid problems with convection caused by a small temperature gradient
along the sample tube [4].
The TI and T2 relaxation times were measured by standard inversion-recovery and
Carr-Purcell-Meiboom-Gill (CPMG) sequences, respectively. A spin-echo sequence was
added at the end of the inversion-recovery sequence to remove the broad-line component
when TI of the narrow-line component was being measured. The estimated errors in the
measured parameters are generally within 3-5% for Tl and within 5-10% for T2 and D.

3. Results and discussion

The melting point depression llT = To - T of a liquid confined within a spherical pore is
given by the Gibbs-Thompson equation [5]

(1)

where To and T are the melting points of the bulk and confined liquid, respectively, R is
the pore radius and kp is a characteristic property of the liquid. The pore size distribution
curves of the investigated mesoporous materials dlldR vs. R, shown in Figure 1, were
determined by measuring the amount of liquid acetonitrile (I) as a function of
temperature by NMR [6] . The average pore size distributions obtained by NMR agree
quite well with those obtained by N2 sorption, both showing quite broad overall
distributions. However, NMR gives a more detailed picture of the distribution, revealing
three relatively well-defined peaks. The peak at the lowest R-value in Figure 1, however,
reflects the surface layer rather than a pore size distribution.
The IH NMR line-shapes of the bulk and confined samples of acetonitrile at 190 K
are shown in Figure 2. The complex line-shape of the bulk sample is typical for a methyl
group undergoing hindered rotations in the solid [7]. The line-shape of acetonitrile
confined within the 200 A pores clearly reveals two distinct components giving rise to a
narrow line superimposed on a broad complex line. The appearance of two components
487

with different T. relaxation times (see below) in the low-temperature region indicates
slow exchange, on the T. time scale, between the liquid-like surface layer, and the solid
at the centre of the pore. For the 60 A pore system, the line-shape is dominated by the
narrow line making the broad line hardly discernible. This is reasonable since the
contribution of the broad component is very low, because of the increased surface-to-
volume ratio relative to the 200 A pores by a factor of ca. 4.

-------
1.2 · 1.2f


60 A Sorbsil 1 1· f'o
200 A Sorbsil
/~\
:::l0.8 · I ~~ :') :::l 0.8 L -
oj oj
\
\ ;; 0.6 l ~, / \
0:: 0.6 0
/':J C ::-
"0 ::!2 ~

~ 0.4 . I ..
e :00.4 ,
\ :l\ /'
0.2· \ 0.2 · ' \j
0
0
?
50
\..
-- 100
0 ,-
0
V

100 200
- 300
-- 400 500
R (A) R (A)

Figure 1. Pore size distributions obtained by measuring the IH NMR intensity of the liquid component of
acetonitrile confined in 60 and 200 A Sorbsil as a function of temperature (solid line) [2]. The circles on the
curves represent the experimental points on the intensity vs. temperature curves. The dashed lines show the
pore size distribution derived from N2 sorption measurements.

bulk

60A

80 40 o -40 -80
kHz

Figure 2. IH NMR line-shapes of acetonitrile in bulk and confined in 60 and 200 ASorbsil at 190 K.
488

The T2 and diffusion data of bulk and confined acetonitrile are shown in Figure 3.
Only T2 and D of the bulk liquid and the narrow-line component of the confined samples
could be measured with the spin-echo and stimulated echo sequences. Both T2 and D of
the confined liquid are considerably smaller than in the bulk liquid. The largest effect is
observed for the 60 A pore system, reflecting increasing restriction to molecular motion
with reduced pore size in accord with the increased surface-to-volume ratio. This
observation can be largely attributed to H-bonding interactions between the nitrogen of
the adsorbate and the silanol surface groups. However, whereas surface interactions
appear to play an important role in slowing down the tumbling motion, the rotation of
the methyl group seems to be little affected [8].
The broad pore size distribution will cause the confined acetonitrile to melt over a
certain temperature region. Actually, the average depressed melting point is ca. 18 and
5.5 K in the 60 and 200 A pore systems, respectively [6]. Thus, the observation that T2
increases rapidly with increasing temperature just below the transition point indicates
that the molecular motion is less hindered in the undercooled pore liquid than in the
liquid-like surface layer. The diffusivity of the confined acetonitrile is practically
continuous across the melting and transition points, whereas the slope is affected
considerably by the change of phase.

.... : • Bulk acetonitrile • Bulk acetonitrile


• 200 ASorbsil • 200 A Sorbsil
• 60 A Sorbsil .. : .60 A Sorbsil
1--'"
.:
....
... . ... ... .. .:-.:
.. :.
' ..

:. ...

'. .
MP TP
.... MP TP

3.5 4.0 4.5 5.0 5.5 6.0 6.5 4.0 4.2 4.4 4.6 4.8 5.0 5.2
1000lT (K- 1 ) 1000lT (K- 1)

Figure 3. Temperature dependence of T2 and D for bulk and confined acetonitrile. MP and TP indicate the
melting and transition points of bulk acetonitrile.

The experimental T, data of the bulk and confined samples of acetonitrile are shown
in Figure 4. If the spin-lattice relaxation of bulk acetonitrile is governed by
intramolecular dipole-dipole interactions modulated by overall tumbling with correlation
time Tc in the liquid phase, and internal methyl rotation (C3 motion) with correlation time
Tm in the solid phase, the relevant relaxation rate is given by Eq. (2).
489

(2)

where Tj='c or 'm


as appropriate, rHH is the internuclear H-H separation within the
methyl group, g( 0) H' 'j) is the spectral density function and A= 1 for the overall
tumbling motion and 3/4 for the C3 motion. The C-H bond length is set to 1.10 A,
giving rHH=1.80 A if the methyl group has tetrahedral synunetry. Then, Eq. (2) gives a
T\ minimum of 234 ms for the C3 motion.
The T\ curve of bulk acetonitrile appears to be approaching a minimum in solid II,
but the actual minimum occurs below the temperature available on our equipment
(Figure 4.). The theoretical fit gives the following Arrhenius parameters: l.32xlO-\4 s
and 11.2 kJ mor\ for the C3 motion. These parameters represent a T\ minimum at 138
K, only 12 K below the experimental point at lowest temperature. The steep increase of
T\ in solid I indicates that the effective correlation time is considerably reduced due to
the onset of overall molecular tumbling. The large reduction of T\ in the liquid as
compared to the solid I phase, and the levelling off of the T\ curve in the high-
temperature region of the liquid phase, presumably reflect an increasing contribution of
the intermolecular dipolar relaxation modulated by translational diffusion .

• Bulk acetonitrile
, ,
, _,0 200 A Sorbsil
.. .. .
~

10 A,I::. 60 A Sorbsil
,'.,, ,,' ,
• I
I a'....
I •••••
a.
• :..·'·0 [J c
• ••
• ~(]a:
-,

, ,
I

, ,

,,
.4
...... 4&1
I I ~ ~~
."'t .C:I
~...F~
,, ,
,, ,,,

MPTP

3 4 567
1000lT (K-1)
Figure 4. Temperature dependence of T\ for bulk and confined acetonitrile. The solid line represents a
theoretical fit to the experimental T\ data in solid II of bulk acetonitrile (£q.(2» . The filled and open symbols
refer to the narrow and broad components, respectively. MP and TP indicate the melting and transition points
of bulk acetonitrile.
490

When acetonitrile is confined to the 200 and 60 A pores, T, is significantly reduced


in the liquid state owing to more restricted reorientational and translational motions (cf.
the T2 and diffusion data). Notice that the T, values of the narrow and broad components
in the solid II region are different, but of comparable magnitude. This is reasonable since
the spin-lattice relaxation in this temperature region is largely governed by the C3 motion
which is rather insensitive to the change of phase.

4. Conclusions

The overall pore size distributions determined by NMR agree fairly well with those
obtained by N2 sorption measurements.
The relaxation times and the diffusivity of the confined samples are significantly
reduced owing to more restricted molecular motions. However, whereas surface
interactions appear to play an important role in slowing down the translational and
tumbling motions, the rotation of the methyl group seems to be little affected.

5. References

1. Barrie, P.J. (1995) NMR applications to Porous Solids, Ann. Rep. NMR Spectrosc.
30,37-95.
2. Putnam, W.E., McEachern, D.M., Jr. and Kilpatrick, J.E. (1965) Entropy and related
thermodynamic properties of acetonitrile (methyl cyanide), J. Chern. Phys. 42, 749-
755 .
3. S0riand, G.H., Hafskjold, B. and Herstad, O. (1997) A stimulated-echo method for
diffusion measurements in heterogeneous media using pulsed field gradients, J.
Magn. Reson. 124, 172-176.
4. Aksnes, D.W. and Kimtys, L. (1994) NMR relaxation and self-diffusion studies of
H-bond exchange and molecular motion in pivalic acid and binary solutions of
pivalic acid and carbon tetrachloride, J. Mol. Struct. 322, 187-194.
5. Jackson, c.L., McKenna, G.B. (1990) The melting behavior of organic materials
confined in porous solids, J Chern . Phys. 93,9002-9011.
6. Aksnes, D.W., Fmland, K. and Kimtys L. (2001) Pore size distribution in
mesoporous materials as studied by 'H NMR, Phys. Chern. Chern. Phys .. In press.
7. Powles, J.G. and Gutowsky, H.S. (1953) Proton magnetic resonance of the CH3
group. I. Investigation of six tetrasubstituted methanes, J. Chern. Phys. 21, 1695-
1703.
8. Zhang, J. and Jonas, J. (1993) NMR Study of the Geometric Confinement Effects on
the Anisotropic Rotational Diffusion of Acetonitrile-d3 , J. Phys. Chern. 97, 8812-
8815.
29 S i
AND 27 AI MAS NMR STUDY OF ALKALI LEACHED KAOLINITE AND
MET AKAOLINITE

N. BENHARRATS 1, A.P. LEGRAND2 and M. BELBACHIR3


ILPPMC, departement de Chimie, Faculte des Sciences, USTG, BP 505
Al M'nouer Gran 31000. Algerie. 2ESPCI, LPQ, FRE 2312, 10 rue
Vauquelin 75005 Paris, France. lLCP departement de Chimie, Faculte
des Sciences, Universite d'Es-Senia BI! 1525 Al M'nouer Gran 31000.
Algerie

ABSTRACT. The alkali leaching of two aluminosilicates, kaolin and metakaolin with
aqueous sodium hydroxide, has been studied. Both of these silicates give
hydroxysodalite with or without the zeolite NaA as intermediate.
The crystallisation kinetics was studied first by X-ray powder diffraction. The
progress of the reactions was also followed by high-resolution 29Si and 27 Al MAS-
NMR spectroscopic monitoring of the signal ratio of the initial materials and the
reaction products.

1. INTRODUCTION

There is a great interest in studying the synthesis of zeolites since they are
formed in systems very far from equilibrium and their evolution is very sensitive to
small variations in the process parameters [1, 2] . For these reasons the understanding
of the process is still incomplete.
Kaolinite, AI4 Si4 0 IO(OH)g, a layered · aluminosilicate, is known to give
hydroxysodalite (HS) on treatment with aqueous alkali at about 100°C [1 , 2]. When
heated, kaolin undergoes several transitions in air. The first one (1) takes place at
about 550°C and produces the disordered metakaolin phase by endothermic
dehydroxilation reaction.
2 A1 2Sips (OH)4 --? 2 A1 6Si0 7 + H 20 (1)
We examine the factors, which control the reaction of the transformation of kaolin
and metakaolin into zeolites using, X-ray diffraction and high-resolution 29Si and 27 Al
NMR. The reactions were followed for a maximum period of 12 hours.

2. EXPERIMENTAL

Kaolin from Tamazert (east of Algeria) was used in these investigations. The clay
491
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 491-496.
© 2002 Kluwer Academic Publishers.
492

was heated at 750°C in a muffle furnace for 8 hours. 10 g of clays, kaolin (K)
metakaolin (MK), were stirred with 100 ml of 2 and 6N of NaOH solutions at 80°C in
a TPX Erlenmeyer for various reaction times. The reaction products were filtered and
washed until pH of the final wash water approximates 7, then dried at 120°C
overnight. The products are denoted as KnN and MKnN, n for concentrations of the
alkali solutions (2N, 6N).
The X-ray diffraction patterns were obtained in a Philips PW 1710 using the CuKa
line (1..=1.54IA). Magnetic Resonance (MAS-NMR) experiments were performed on
Bruker ASX spectrometers in zircon rotors. 27Alone-pulse experiments were
performed at 11.7 T with a spinning rate of 12 kHz, a selective pulse «7tI6) duration
of 0.5 J.lS, a recycle time I s, and 200 acquisitions. The chemical shift was referenced
to a 0.1 M aqueous solution of Al(N0 3h- The 29Si one-pulse spectra were recorded at
7.04 T with a spinning rate of 4.5 kHz. The 7tI2 pulse duration was 2.5f.ls, the recycle
time was 5 sand 500 acquisitions were accumulated. 29Si chemical shifts were
referenced to TMS.

3. RESULTS

3.1. X-RAY POWDER DIFFRACTION


XRD patterns of kaolin (figure 1) show diffraction lines ofkaolinite [3], additional
amount of illite (mainly at 29 = 9° and 14°) and quartz (at 29 = 29 and 21°).
Diffraction patterns of the leached kaolin samples for increasing reaction times are
shown in figures I, 2. Kaolinite reflections disappear progressively as the reaction
proceeds. For K2N samples, specific lines of NaA zeolite appear after 2 h, followed
by HS ones after 6 h reaction time. After 2 h, K6N exhibits peaks which are
characteristic ofHS, without detectable occurrence ofNaA.
Metakaolin is amorphous to X-ray, giving a broad featureless spectrum. The
structural changes occurring as the reaction progresses are shown in figures 3, 4. The
MK2N diffraction pattern is remarkably different from that of K2N. No new
crystalline phases can be observed before 4 h reaction time. After 6 h, the only new
crystalline phase is the NaA zeolite. For MK6N samples both HS and NaA are
detected simultaneously.

3.2. 29Si MAS-NMR


The 29Si spectrum of kaolin shows the usual shifts given by the literature [4, 5]. It
consists of a main resonance at -91.6 corresponding to Q3(OAl) sites with a shoulder at
about -86.6 ppm attributed to Q3(lAl) sites (figures 5 and 6). This observation is
clearly related to the corresponding 27Al spectrum (figure 10). Using the Lowenstein's
rule, the relative intensity ratio gives the average SilAl 1v ratio in tetrahedral sheets. It
is estimated to be 17.
The 29Si spectra of leached kaolin samples, show two resonances at -91.6 and -87
ppm (figures 5, 6). The latter becomes more intense and narrows with a maximum at
-86.6 ppm. In the K6N6h spectra, these changes appear earlier and are more marked.
493

FIGURES I and 2. XRD patterns of K2N and K6N series with reflections of HS and
quartz (Q) and Illite (I).

Kaolin
m g ro _ ~ ~ ~
10 15 20 25 30 35 40 45 50
2 e angle 29 angle

FIGURES 3 and 4. XRD patterns of MK2N and MK6N series with reflections of HS
and NaA(*).

2 a angle

The 29Si NMR spectrum of metakaolin shows (figure 7) a broad single component,
specific of amorphous materials, centred at -90 ppm. Upon alkali leaching, it splits
into two broad resonances at about -89.6 ppm and -86.6 ppm. As for the leached
kaolin sample, they narrow as the reaction proceeds. Also, the maximum of the
downfield contribution shifts to -85 ppm for the 12h MK6N sample (figure 8).
According to [4] the resonance at -85 ppm is attributed to Q4 sites in HS and that at
-89.6 ppm to Q4 in NaA. The resonance at -86.5 ppm could correspond in part to the
resonance of Q3(lAI) sites in residual kaolin. However, the increase in its intensity
upon leaching of metakaolin is contradictory. The observed line-narrowing is an
indication of ordering of the Si environments and reveals that the crystalline structure
of the reaction products evolves with the reaction time. Since it occurs in the MK
samples at 2 h, when no crystalline products were detected by XRD, it might,
therefore, reflect the formation of an amorphous aluminosilicate phase.
494

FIGURES 5, 6, 7 and 8. 29Si NMR spectra of selected samples from K and MK series .
• 86.6 ppm

~ K2H1211

K2_
K6N12h

-
K2N3h
K6H3h
K2N2h

Kaolin KaolIn

-60 -70 -80 -9000m ·100 ·110 -120 -60 ·70 ·80 -90 ppm ·100 ·110 ·120

_,211

_.
MK6N6h

MK6N4h

MK6N2h

-60 ·70 ·BO ·90~m ·100 ·110 ·120 60 -70 -BO -90 ppm -100 ·110 -120

3.3. 27AI MAS-NMR


The 27 Al NMR spectra of selected samples from the K and MK series are given in
figures 9, to, 11 and 12. The kaolin spectrum exhibits mainly the typical resonance at
3 ppm of octahedral aluminium (Alvl) in a phyllosilicate but also a small but
significant (17% of the total intensity) contribution at 70 ppm assigned to tetrahedral
aluminium co-ordinated to oxygen (A1 Iv ). The latter shows that the Tamazert kaolin
does not consist of pure kaolinite but in fact of an interstratified illite-kaolinite.
The 27 Al spectra of K2N samples (figure 9) show, in addition to the Al lv imd A1 v1
resonances of the unreacted kaolin, a line at 60 ppm attributed to Al lv in zeolite [4].
The maximum of this resonance shifts from 59 to 61 ppm, and its width decreases with
increasing reaction time. This reflects structural ordering and changes in composition
from NaA to NaA and HS as observed by XRD. The same spectral' features are
observed for the K6N series (figure 10) except for the line shift. Indeed, only HS was
detected by XRD.
The 27 Al spectrum of metakaolin consists of three lines at 3, 27 and 65 ppm that
are typical of Al vl , pentahedral aluminium co-ordinated to oxygen (Alv) and AI IV ,
respectively (figure 11) [6]. For both MK2N and MK6N samples (figures 11 and 12)
the spectra consist, apart from residual metakaolin resonances, of a resonance at 59
ppm corresponding again to Al lv in NaA. A shoulder at 61 ppm marking the formation
of Al lv in HS is also observed for the MK6N series, again in good agreement with
XRD.
The contributions of Al lv in the zeolite products obviously increase with reaction
time at the expense of Al vl in the starting materials. However, the Al lv resonances are
too broad for the formation of HS and NaA to be quantified independently.
Interestingly, the Al lv contribution in the starting kaolin sample appears to be constant
at 17%.
495

FIGURES 9,10,11 and 12. 27AI NMR spectra ofK2N, K6N, MK2N, MK6N.
60 7 ppm

2.75 ppm

K2N12h
K6N12

K6N6h
K2N6h

K6N3h
K2N4h

Kaolin
Kaolin

·20 ·40 ·60 -80 120 100 80 60 40 20 0 ·20 -40 -61


ppm ppm

MK2N12h 2.97 ppm


AI VI
MK6N12h

MK2N6h MK6N6h

MK2N3h MK6N4h

Metakaolin
Metakaolin

120 100 80 60 40 20 -20 -40 -60 120 100 80 60 40 20 0 -20 -40 -50
ppm

4. DISCUSSION

The relative contributions of the HS and NaA zeolites can be obtained by


decomposition of the spectra into Lorentzian components. The integrated areas,
relative to the total area, are plotted against reaction time in figure 13. The HS
contribution is absent from the MK2N spectra. These plots represent the rate of
transformation of kaolin or metakaolin into the HS zeolite and its amorphous gel
precursor. It appears to be linear with respect to time. This is a clear indication that the
zeolitization process is not diffusion-limited. Moreover, the rate of reaction, as
reflected by the slope, increases with the NaOH concentration. In order to distinguish
the formation of HS from that of the amorphous precursor, the XRD data were also
quantified. The synthesis yield of HS was estimated by quantitative analysis based on
the peak summation procedure. Figure 14 shows the evolution of the fraction of HS
versus the reaction time. As for the NMR data, linear behaviour is observed but HS is
formed faster from kaolin than from metakaolin.
Comparison of the qualitative kinetic data obtained from NMR and XRD,
therefore, leads to the following conclusion: If one includes the amorphous products,
the metakaolin precursor reacts as fast as kaolin, although HS is formed faster for the
latter. This shows that the rate-limiting step is not the formation of the amorphous gel
precursor but its reorganization into a crystalline HS product. Furthermore, it appears,
that this reorganization is easier for the gel formed from kaolin than from metakaolin.
496

The resulting gel is therefore probably more heterogeneous than that resulting from
kaolinite, and the time necessary for its reorganization into HS, the induction period, is
longer.

FIGURES 13, 14. Plot of percentage ofHS versus reaction time obtained by XRD data
(left, K6N 7.5 and MK6N 4.1 % per hour slopes) and MAS-NMR data (right, K6N
7.5, MK6N 5.8 and K2N 1.5 % per hour slopes).
l00rr=~~~~~~~~~~~
linear fits
••• • • K6N

-
.. .... MK6N
• K2N •

• ",,f' .... .. .

...
6 8 10 12 6 10 12
reaction time (hours) reaction time (holXS)

5. CONCLUSION

After activation by alkali leaching for several hours, large amounts of kaolinite or
metakaolinite remain together with the three reaction products: an amorphous
aluminosilicate gel precursor, intermediate NaA zeolite, and the final product, HS
zeolite. For both kaolin and metakaolin, the conversion is greater for high alkali
concentrations. However, because the reaction is not diffusion-limited, the same rate
of conversion is observed irrespective of the choice of kaolin or metakaolin. Although
the ordered structure of kaolinite might hinder the diffusion of active ions from the
solid to the aqueous medium, this step is not rate-limiting. Furthermore, with respect
to HS formation, kaolin is more favourable, as it forms an amorphous gel precursor
that crystallises more easily into zeolites. With different reaction times, alkali
concentrations and the dehydroxylation of kaolin into metakaolin, it is thus possible to
prepare samples with varying degrees of zeolitization.

ACKNOWLEDGEMENT
The authors gratefully thank Dr. 1.B d'Espinose de la Caillerie from LPQ, ESPCI, for
help in the recording and discussions ofNMR characterisations.

6. REFERENCES
I. R.M. Barrer. Hydrothennal Chemistry of Zeolites, Academic Press, London, (1982).
2. D.W. Breck. Zeolite Molecular Sieves. 1. Wiley and Sons, New York (1974).
3. 1. Rocha, 1. Klinowski. J. Phys. Chern. Miner.17, 179, (1990).
4. G. Engelhardt, D. Michel. High resolution solid state NMR study of silicates and zeolites.
1.Wiley and Sons, NewYork (1987).
5. J. Sanz, J. M. Serratosa. 1. Am. Chern. Soc., 106,4790-4793, (1984).
6. 1. Sanz., A. Madani., 1. M. Serratosa, 1.S. Moya, S. Aza. 1. A. Ceram. Soc.,71, C-418, (1988).
Electron accepting properties of Zr02 based catalysts studied by EPR
of paramagnetic complexes of probe molecules.

M.V .Burova, A.V.Fionov, A.O.Turakulova

Department o/Chemistry, Moscow State University, Leninskije Gory, Moscow, 119899,


Russia. Fax: 7-095-9394575. E-mail: fionov@kgemsu.ru

Introduction
Zirconia is of great interest as a catalyst and support. The tetragonal and cubic
modifications are most important for these purposes because of significant specific surface
area. Surface properties of these modifications are defined by the presence of Lewis acid sites
- coordinatively unsaturated zirconium (IV) ions. These modifications are metastable at
ambient conditions, and the use of dopants (metal cations of II or III groups) is necessary. The
most known dopants are yttrium and cerium oxides [1-3]. Other metals of III group (AI, Ga)
are also studied as possible stabilizers of the tetragonal zirconia [4-6]. The change of the
dopant content causes the changes in the nature and concentration of the Lewis acid sites on
the surface [7]. In order to improve catalytic properties of zirconia-based catalysts the
complex systems (containing more than two components) are studied. Nevertheless there are
few data on the surface properties of these systems, especially on the Lewis acidity, which is
very important for predicting the adsorption and catalytic properties.
The main goal of this work was to study Lewis acidity of the zirconia catalysts,
containing two dopants, in order to define how surface properties of such systems depend on
the composition. Electron accepting properties of the surface were studied by EPR of
paramagnetic complexes of probe molecules (2,2,6,6-tetramethylpiperidine-N-oxyl (TEMPO)
and anthraquinone). This method has been successfully used previously to study surface
properties of more simple catalysts, such are alumina [8], gallia [9], zirconia [10), Al20rZr02
and Ga20rZr02 [7].

Experimental

Zirconia samples, containing 20 % mol. of alumina or gallia as well as 10 % mol. of


alumina and 10 mol. % of M20) (M=Ga, Y, In) were prepared by co-precipitation of
497
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 497-505.
© 2002 Kluwer Academic Publishers.
498

hydroxides from corresponding nitrates and Zlrcomum oxynitrate water solutions with
ammonia under pH 9. A precipitate was filtered, washed by weak (1 %) ammonia solution,
dried for 12 h at 293 K, 12 h at 393 K and then calcinated at 873 K during 3 hours.
The specific area was measured chromatographically by the nitrogen adsorption at 77
K and desorption at r.t. The crystal structures of the samples were determined by X-ray
analysis with CuKa-radiation. The Raman spectra were recorded at EQUINOX 55/S (Bruker)
spectrometer with FRA 106/S accessory with Nd: YAG laser source (/... = 1046 nm) and
InGaAs detector.
The synthesized samples are characterized in the Table I.
Table 1.
Specific surface area and phase composition of studied samples.
Sample Ssp' m"/g phase composition*
20 % mol. AI 20 r Zr02 136 A
20 % mol. Ga20r Zr02 99 A+T
10 % mol. Ga20r 10 % mol. AI 20 r Zr02 142 A
10 % mol. Y20r 10 % mol. AI20 r Zr02 90 T+A
10 % mol. In20r 10 % mol. AI20 r Zr02 99 T+A

* A- amorphous Zr02, T- tetragonal Zr02

Before the adsorption of probe molecules all samples were calcinated for 2 h at 743 K
in the air and then for 2 h at the same temperature under vacuum 10.3 Pa. TEMPO was
adsorbed at room temperature from the gaseous phase. Anthraquinone was adsorbed
according to the method described in [8]. The final temperature of anthraquinone adsorption
was 473 K.
EPR spectra were recorded using X-band spectrometer of RE-1306 type. g-values
were determined with a reference to diphenilpicrylhydrazyl standard (g = 2.0036). Spin
concentrations were measured by double integration of the spectrum and comparison with
calibrated samples using an intermediate standard (Cr3+ in corundum). EPR spectra were

recorded at room temperature or at 77 K (for AiNi measurements of adsorbed TEMPO).


499

Results and discussion

Phase composition
The data on the phase composition and specific surface areas of the studied samples
are presented in Table 1. It has been shown that the dopants prevent the monoclinic zirconia
formation. All samples contain the tetragonal zirconia in the mixture with amorphous phase.
The impact of amorphous zirconia is significant in the case of the samples with only alumina
and/or gallia. The presence of yttrium or indium oxide causes more crystallized tetragonal
zirconia formation. Raman spectra are in accordance with X-ray powder data.
It is well known that yttrium oxide forms stable solid solutions with tetragonal
zirconia [1]. On the other hand, aluminum and gallium oxides form only metastable solid
solutions with zirconia [4-6]. On our opinion, this fact can be explained by similar ionic radii
of Zr4+ (0.082 nm) and y3+ (0.097 nm), in contrast to smaller radii of A1 3+ (0.057 nm) and
Ga)+ (0.062 nm). Indium cations (0.092 nm) seem to be similar to yttrium cations.
As a result, the formation of solid solutions in tetragonal zirconia is more preferable in
the presence of yttrium and indium. The interaction between amorphous Al 20 3 and Zr02 as
well as between Ga20) and Zr02 with the formation of metastable solid solutions proceeds
more difficult, especially for Al20rZr02 System, because the aluminum cations have a largest
difference in ionic radius in comparison with zirconium ones. Therefore the growth of
tetragonal zirconia crystallites is impeded in the presence of alumina, and the sample with
high alumina content is the most amorphous.
It could be expect that in the case of two dopants with different nature the more stable
solid solution will form, with the segregation of third component on the surface. Previously
[11] it was found that in the case of the Al20rY20rZr02 system the crystallites of stable
solid solutions (tetragonal Y20rZr02) are surrounded by smaller Al 20 3 crystallites.
It was interesting to study the influence of the sample composition on the Lewis acid
properties of the surface.

EPR spectra of adsorbed TEMPQ,

The Lewis acidity of the oxide surface can be determined by the EPR of adsorbed
paramagnetic molecule - nitroxide, usually 2,2,6,6-tetramethyl-piperidine-N-oxide (TEMPO) .
The nitro xi de, acting as Lewis base, coordinates to an electron acceptor through the non-
bonding electron pair of the nitroxide oxygen. The redistribution of the spin density in the
500

adsorbed nitroxide causes an increase in the parallel component on the nitroxide 1"N hyperfine
splitting (h.f.s.), AN.. For example, the value of AN (± 0.2 G) of TEMPO in toluene solution

is 34.6 G, in glycerin-water solution is 37.8 G, in H)P04 (10%) is 39.5 G [12] . For TEMPO

complexes with Lewis acids the A iN value varies in the range from 40 to 45 G, depending on

the electron acceptor strength of the site. Thus, the measured value of A !N! serves as a scale

for evaluating of electron-pair acceptor strength of acid sites. When the nitroxide interacts
with the metal nUcleus having a spin and significant magnetic moment, the EPR spectrum
may contain an additional h.f.s. This unambiguously identifies the nature of the coordinatively
unsaturated (cus) cation (for example, AI or Ga).
Zirconia, alumina and gallia surfaces have an appreciable Lewis acidity.

Coordinatively unsaturated Zr4+ (Zr~~s)' Al l + (AI~~s) and Gal + (Ga~~s) are supposed to be
Lewis acid sites.
The EPR spectra of TEMPO adsorbed on studied samples are presented on Fig. 1.

The most intense spectrum is three component one with the A~i = 42.5 ± 0.5 G. This

spectrum corresponded to the TEMPO complex with coordinatively unsaturated Zr4+ ion [10].
In regions A and B (Fig. 1 a-e) the impact of another type of EPR spectrum is also
noticeable. It is difficult to determine h.f.s. on the aluminum (gallium) in this spectrum
because of the dominant impact of three component one. Nevertheless it is possible to
suppose that this spectrum may be assigned to the TEMPO complex with coordinatively
unsaturated Ae+ (or Gal +) ions [12] .
On all alumina containing samples the spectrum with g=1.976 has been registered
(Fig. 1 a, c-e). It is not associated with TEMPO complexes and can be assigned to the Zr3+
ions [13]. The broadening of this spectrum in the presence of air and the restoration of t~e

signal after an evacuation allow us to conclude that these ions are localized on the surface.
The concentration of this species is quite low (less than 2.10 13 m· 2) and detailed investigation

ofZrl + ions is out of the scope of this work.


As a result, with the help of adsorbed TEMPO it is possible to determine
coordinatively unsaturated Zr4+ ions, which are similar to ones on the zirconia surface. The
possibility of the presence of coordinatively unsaturated aluminum and (or) galhumions can
not be confirmed unambiguously, and there are necessary additional experiments to determine
it.
501

EPR spectra of adsorbed anthraquinone.

It was shown previously [8, 9], that an interaction of the anthraquinone with electron-
accepting sites on alumina and gallia surfaces leads to the formation of cation-radical complex
of anthraquinone on the surface. Depending on experimental conditions and sample
pretreatment, the EPR spectrum of complexes exhibits h.f.s. due to either one or two
equivalent neighboring aluminum nuclei (either 6- or II-component spectrum) [8] or gallium
nuclei (either 4- or 7-component spectrum) [9]. On the basis of different experimental data,
these complexes were intetpreted as complexes with coordinatively unsaturated aluminum or
gallium ions, which are Lewis acid sites on alumina or gallia surfaces, respectively. The

formation of the anthraquinone complex with two equivalent cations (AI~~s in the case of 11-

and y-alumina and Ga~~s in the case of y-gallia and partly of ~-gallia) reveals that these sites

are located regularly on the surface of these oxides. The possibility of regular arrangement of

Lewis acid sites is determined by the crystal structure of alumina and gallia, which has
influence on the surface structure.
With the help of this method, the concentration of coordinatively unsaturated
aluminum or gallium ions can be determined via the maximum concentration of the
anthraquinone complex, taking into account the number of Lewis acid sites (one or two)
involved in the complex formation.
The anthraquinone adsorption on systems under study resulted in paramagnetic

complex formation. The observed EPR spectra (Fig. 2 a-e) were the superposition consisted of

single line spectrum (which was previously described as anthraquinone paramagnetic complex
with Zr~~s [7]) as well as multi-component one. The mUlti-component spectrum contained

h.f.s. from only one nucleus 27Al (6 components), g=2.0036, a=9 ± 0.2 G, or 69.71 Ga (4
components), g=2.0036, a=42 ± 0.5 G. Similar spectra were observed previously after the
anthraquinone adsorption on the alumina [8], gallia [9] and M20rZr02 (M=AI, Ga) [7]. EPR
spectrum of anthraquinone adsorbed on the A1203-Ga20rZr02 sample consisted on the 6-

component spectrum (anthraquinone-AI 3+ complex) and single line spectrum (anthraquinone-

Zr4+ complex) (Fig. 2 c). 4-component spectrum (anthraquinone-Ga3+ complex) was not

detected. Evidently, the anthraquinone adsorption data are complementary to the TEMPO

adsorption ones.
Data on the concentration of coordinatively unsaturated aluminum (gallium) ions
502

determined on maximum concentration of 6- (4-) component spectrum are summarized in the


table 2.

Table 2.
Concentration of coordinatively unsaturated aluminum and gallium ioIls
on the surface of studied samples.
Sample Concentration of Lewis acid sites, EPR
I Nx 10- 16 sites/m2 spectrum
[AI'- cus] [GaY cus] shape *
20 % mol. Al 20 r Zr02 4.7 - 6+S
20 % mol. Ga203- Zr02 - 1.9 4+S
10 % mol. Ga203- 2.6 - 6+S
10 % mol. AI 20 r Zr02
IO%mol. Y20r 4.8 - 6+S
10 % mol. AI 20 r Zr02
10 % mol. In203- 3.0 - 6+S
10 % mol. A1 20 3 - Zr02

* Abbreviations of EPR spectra:


6 - six-component spectrum of anthraquinone complex with single AI~~s;

4 - four-component spectrum of anthraquinone complex with single Ga~~s;


S - single line spectrum of anthraquinone complex with Zr~~s'

Obtained results can be explained by the formation of solid solutions of alumina and
gallia in the zirconia. No alumina or gallia phases formation was taken place because there
was no detection of paramagnetic complex of the anthraquinone with two Lewis acid sites
(II-component or 7-component).

According to presented data the concentration of AI~~s was larger than that of Ga~~s'
This fact can be explained by less stability of metastable solid solutions A1203-Zr02 in
comparison with Ga20rZr02' As a result the surface was enriched by aluminum cations. This
process (enrichment of the zirconia surface by aluminum) increased, when the third
component, Y20 3, which is able to form stable solid solutions with zirconia, has been added.
503

The In203 also improved the migration of aluminum cations to the surface but not so effective
as Y20 J .

Acknowledgments

Authors are grateful to RFBR for financial support (grant 98-03-32129a).

A.V.Fionov is grateful to INTAS for financial support (grant YSF 00-252).

References
1. Kaspar 1., Fornasiero P., Graziani M., Catal. Today. 50 (1999) 285.

2. lkryannikova L. N., Aksenov A. A., Markaryan G. L., et. aI., Appl. Catal. 210 (1-2) (2001)
225.
3. Vidmar P., Fornasiero P., Kaspar 1., et. aI., 1. Catal. 171 (1997) 160.
4. Moreau S., Gervais M., Douy A., Solid State lonics. 101-103 (1997) 625.
5. Gao L., Liu Q., Hong 1. S., et. aI., J. Mat. Sci. 33 (1998) 1399.
6. Barret P., Berthet P., 1. Phys. III France. 7 (1997) 483.
7. Pushkar Yu. N., Parenago 0.0., Fionov A.V., Lunina E.V., Coil. Surf. A: Physicochemical
and Engineering Aspects. 158 (1999) 179.
8. Lunina E. V., Markaryan G, L., Parenago 0. 0., Fionov A. V., Colloids and Surfaces A:
Physicochemical and Engineering Aspects. 115 (1996) 195-206.
9. Parenago 0. 0., Pushkar Yu. N., Turakulova A. 0., et. aI., Kinet. Catal. 39 (2) (1998), 268.
10. Lunina E. V., SeJivanovsky A. K., Golubev V. B., et.. ai., Zh. Fiz. Khim. (in Russian). 56
(2) (1982) 411; Lunina E. V., Badina E. Yu., Kuznetsova N. N., Markaryan G. L., Dokl.
Akad. Nauk SSSR 319 (4) (1991) 914.
II. Navarro L. M.; Recio P.; Duran P., 1. Mater. Sci. 30 (8) (1995) 1931.
12. Lunina E. V., Appl. Spect. 50 (1996) 1413.
13. Morterra C., Giamello E., Orio L., Volante M. 111. Phys. Chem. 94 (1990) 3111.
504

T=77K !g=2.0036
B Z r3+
a ___ ~

b __~

e ___

Fig. 1. EPR spectra of 2,2,6,6-tetramethylpiperidine-N-oxyl (TEMPO) adsorbed on the


surface of (a) 20% mol AlP3 - Zr02; (b) 20% mol Ga203 - Zr02; (c) 10% mol Al 20 3 - 10%
mol Ga203 - Zr02; (d) 10% mol AI 20 3 - 10% molln203 - Zr02; (e) 10% mol AI 20 3 - 10%
mol YP3 - Zr02'
505

g=2.0036
T=293K
Q----------------

/"\
b _ _ _- _______

c _____________ ~

r
d _ _ _ _ ___

e ______________ ~

20G

Fig. 2. EPR spectra of anthraquinone adsorbed on the surface of (a) 20% mol AI 20 3 - Zr02;
(b) 20% mol Ga203 - zr0 2; (c) 10% mol AI 20 3 - 10% mol Ga203 - Zr02; (d) 10% mol
AI 20 3- 10% mol In203 - Zr02; (e) 10% mol AI 20 3- 10% mol Y20 3- Zr02'
CHARACTERIZATION OF ORGANIC NANOPARTICLES
SYNTHESIZED IN MICROEMULSIONS BY lH NMR

F. DEBUIGNE, L. JEUNIEAU AND J. B.NAGY


NMR Laboratory, FUNDP
61 rue de Bruxelles, 5000 Namur, Belgium

1. Introduction

Organic nanoparticles of cholesterol or rhodiarome are obtained in microemulsions by


direct precipitation of the active compound in aqueous cores because of their
insolubility in water [1, 2].
Then NMR is used to know how the nanoparticles are stabilized and the location of
the water and the solvents like acetone or chloroform.

2. Materials and methods

The system of microemulsion used is AOT «bis(2-ethylhexyl)sodium


sulfosuccinate)lheptanelwater [3]. For the study of water by NMR, deuterated water
99.9% (Cortec) is used. The organic molecules are cholesterol and Rhodiarome which
are insoluble in water but soluble in some organic solvents like acetone and chloroform.
For NMR studies of deuterated molecules, acetone d6 and chloroform d (99.8%) are
used. The measurements are carried out on a Bruker MSL 400 spectrometer.
The empty microemulsions are first prepared. Then a solution of the organic
compound in an appropriate solvent is injected and treated with ultrasounds. The
nanoparticles are thus formed by direct precipitation reaction of the active compounds
in the aqueous cores because of their insolubility in water.

3. Results and discussion

3.1. HYPOTHESIS OF STABILISATION OF ORGANIC NANOPARTICLES

The hypothesis of the nanoparticles formation was proposed in previous papers [1, 2].
As far as the stabilisation of the nanoparticles is concerned the following models can be
proposed (figure 1).
507
J. Fraissard and O. Lapina (eds.). Magnetic Resonance in Colloid and Interface Science. 507-512.
© 2002 Kluwer Academic Publishers.
508

wate,
0--

~
.~ -- •
~
~~

solve nl

W ater mo lecu le
A 6 C
Figure I , Hypotheses lor the nan opart icle stabilisation,

The first one (A) is that nanoparticles are stabilized in the organic phase formed by
heptane and are surrOlmded by a layer of water. The surfactant tails are in heptane.
This hypothesis has been previously advanced for AgBr nanoparticles [4]. The second
one (B) is that nanoparticles, in the organic phase, are in direct contact with the polar
heads of the surfactant. In hypotheses A and B the solvent acetone or chloroform are
supposed to be dissolved in the organic phase and also in the water layer for hypothesis
A. The third one (C) is that nanoparticles are surrounded by the solvent (acetone or
chloroform) and some water is also adsorbed on the nanoparticles.
The use of deuterated water and solvents allow us to determine which one of the
hypotheses is correct.

3.2.2HNMROFDEUTERATEDWATER

First, the influence of the temperature is studied on the state of the water in the cores of
the empty microemulsions. The study of the water in the microemulsions containing
only the solvent and also in microemulsions containing nanoparticles is carried out.
The R factor (= [Water]/[AOT]) of the studied microemulsions is equal to 4 and 1. This
factor is related to the size of the aqueous core [5]. The reference is pure deuterated
water. The 2H NMR spectra of different microemulsions are represented in figure 2.

18
16

14

12
20
2.2

1.1 O.S

10 1.' 0,7
1,' Q6
1.2
1.0
0.'
o,.t===::::::::/--""~~==!293!SK 0.3
0.4 0.2"----
263K

002~~~=:=~§3
.0
<),2
.,
ppm
.10
o'oE~~~~?s~~
0,1

.0.1
10 o ppm·' ·10
10 ·S ·10
MIm)

A B c
Figure 2, 2H NMR spectra of water in microemulsions (R=l) A) empty; B) with solvent (chloroform); C)
with nanoparticles of cholesterol as a function of temperature.
509
Two types ofNMR lines are observed at room temperature. The first one is located
at -0.86 ppm. This stems from the free molecules of water and from water molecules
bound to the surfactant. Hydrogen bonds characterize these states of water and hence,
the chemical shift occurs at lower fields. For lower R values, the shift corresponds to
the bound water molecules which interact with the surfactant polar group and with the
counterion Na+. The second and third lines are due to the water molecules as monomers
or dimers. These are located at higher fields (-3.65 and -4.04 ppm, respectively),
because only a small amount of hydrogen bonds are present. The water molecules also
interact with the surfactant. These molecules are trapped at the interface.
The microemulsions are stable as a function of temperature. When the temperature
decreases, the intensity of the line decreases and it broadens until it finally vanishes.
The solidification of the water occurs at 253 K in the empty microemulsions and at 243
K in the other two cases while pure water solidifies "at 273 K These valuet cannot be
directly compared with those obtained in microemulsions with low R Inaeed, the
decrease of the line of free water stems from two factors: the solidification of the water
and the increase of the bound water. There is no significant shift of the NMR lines.
The other two lines of the trapped water do not significantly change: their location
(light shift towards low fields) and width stay constant. TI measurements (longitudinal
relaxation time) have also been carried out. TI is greater for these trapped molecules,
indicating a smaller mobility. Indeed, as T I increases when the temperature decreases
for trapped molecules, we can conclude that an increase of TI corresponds to an
increase of correlation time (a smaller mobility). The comparison of the two types of
trapped water molecules shows that monomers are situated at higher field (-4.04 ppm)
and dimers at lower field (-3.65 ppm). Indeed, for monomers, the shift is towards
higher fields because of the decrease of the hydrogen bonds and the relaxation time is
greater because the water molecules are less mobile. The nature of the "free-hydrogen
bound water" seems to be similar in microemulsions containing the solvent or the
nanoparticles if the values R are low: log Vl/2 increases as a function of liT. For the
trapped water molecules, log VII2 (V1/2 stands for the width at half height) does not
significantly change as a function of the temperature. For R=l, the variation oflog TI as
a function of IIT is similar in all cases, i.e. for empty microemulsions, for
microemulsions containing the solvent and finally with nanoparticles. Indeed, the water
molecules are still structured at the interface and are not perturbed by the presence of
the solvent (or the particle).
From these results it can be seen that the presence of the nanoparticles or the
solvent in the microemulsions does not significantly influence the nature of the water.
Hence, the nanoparticles should be located in the organic phase and not in the aqueous
cores. The other reason for localizing the nanoparticles in the organic phase is that the
diameter of the nanoparticles (50A) is greater than the aqueous cores size (R=I:
d=12.6A) and it is thus impossible to locate the nanoparticles inside the aqueous cores.
On the other hand, some water could be adsorbed on the nanoparticles. Another
hypothesis is that the solvent could be in the organic phase or adsorbed on the
nanoparticles, but not in the aqueous cores (hypothesis C). In order to check the
veracity of this hypothesis, deuterated solvents like CDCh and CD3COCD3 are used in
the following experiments.
510

3.3. 2HNMR OF CDCh

The aim is to determine the location of the solvent chloroform. Different spectra have
been compared: pure CDCh, CDCh in heptane and CDCh in heptaneiAOT solution.
The first line stands for the associated CDCh at 2.40 ppm (pure CDCh) and 2.66 ppm
for CDCh in heptane or in heptaneiAOT solution. The two small lines at -3.18 and-
3.58 ppm (higher fields), in the case ofCDCh in heptane and in heptaneiAOT solution,
stand for CDCh dimers or monomers. In the microemulsions, the lines are situated at
the same position (figure.3). The attribution of the lines is also based on TJ for CDCh:
the line situated at 2.7 ppm has a TJ equal, for example, to 251 ms (at T = 293K and for
R = 1) and the lines at higher fields have TJ equal to 682 ms (for low temperatures).
These last lines at higher fields are due to monomers or dimers because they are less
mobile. These molecules can be situated in the heptane phase, or can be adsorbed on
the exterior of AOT molecules, i.e at the interface. The line at lower fields has almost
the same chemical shift as those obtained in heptaneiAOT solution. Thus, the CDCh
molecules should solvate the polar head group of AOT and should be in contact with the
aqueous cores.
The study of the influence of the temperature on the CDCh in microemulsions
(empty or with nanoparticles) is also carried out. Figure 3 shows the different spectra .
• 75
.11>
.0>
.60

.".'"...
.S>

0.5 291K • ,,~_.-'I"'_--/\_~

.".j------'lII'---'L'---=-=
0.4 WK .lO

OJ 263K
0.2 2S3K
~~j----- '.~---'Irr----=""
.10-1--------1 '-----"10''---,",-"
01 243 K
2DK
..~~t:::==-;7
05+---_-..---_ _
10

""" A
ppm

B
Figure 3. COCb spectra ofmicroemulsions (R=l): A) with the nanoparticles and B) without nanoparticles as
a function of temperature.

The great lines (due to associated CDC h) decrease as a function of temperature in


the two cases. In presence of the nanoparticles, the intensity is smaller and the line is
broader at 233 K in comparison with those without nanoparticles. The hypothesis is
that the CDCh would be adsorbed on the nanoparticles. Indeed, the width of the lines is
inversely related to the T2* (transversal relaxation time). Thus, when the width
increases, the relaxation time decreases and the molecules are more frozen. The other
two lines at higher fields (due to monomers or dimers) do not significantly change as a
function of the temperature.
LogT J in the case of the presence of nanoparticles is greater than when the
microemulsion is empty, especially for R=4. The CDCh molecules should be more
frozen with the presence of nanoparticles as a consequence of their adsorption. For
R=I, all molecules are already frozen and there is no significant change for 10gTJ. The
hypothesis C in figure 1 takes into account the location of the solvent and the
stabilisation of nanoparticles.
511

Nanoparticles of rhodiarome have also been synthesized with the same method but the
solvent used is acetone. CD3COCD3 is used to study its location in the same way than
the chloroform. Different experiments have also been carried out
The chemical shift of pure acetone is located atlow fields (2.64 ppm) because of
the interactions between the molecules of acetone. When the acetone is dissolved in
water, monomers dominate the spectrum at -2.62 ppm because of the interactions
between the molecules of acetone and water are greater than between acetone-acetone.
Finally, in the case of acetone in heptane or in heptane/Am solution, there are
monomers and some dimers at -2.35 ppm.
The influence of the temperature on the solvent in microemulsions (empty and with
nanoparticles) is also studied and the spectra are represented in figure 4.

"
"

ppm ppm
A B
Figure 4. C~COC~ spedra of solvent in miaoenwlsions (R=4): A) with nanoparticles and B)
without nanoparticles as a function of temperature.

The positions of the lines are similar to those obtained in the case of acetone in
heptane/Am. The line at -2.4 ppm (monomers) decreases as a function of temperature
and should stand for the acetone molecules in the aqueous cores or in interaction with
the Am molecules. The intensity is greater in the case of the microemulsions with
only the solvent and the line is larger in the case of micromulsions with nanoparticles at
233K. Hence, the solvent (CD3COCD3) should be also adsorbed on the nanoparticles in
the same way as in the case of CDCI3 . The other lines at -3.1 and -3.5 ppm stand for
monomers dimers in the organic phase and do not significantly change as a function of
the temperature. Hence, hypothesis C also explains these observations.

4. Conclusion

First, the structure of water has been studied by 2H NMR of deuterated water. The
chemical shift is independent of the presence of the nanoparticles. Some water is
adsorbed on the nanoparticles, in addition to the surfactants. Secondly, NMR.
measurements of deuterated chloroform and acetone have been carried out. Those seem
to be adsorbed on the nanoparticles. The NMR. spectra allow one to conclude that the
nanoparticles are essentially stabilized in the organic phase with solvent and surfactant
molecules (water, chloroform or acetone and Aar) adsorbed on them.
512

5. References

1. Debuigne, F. (1999) Preparation of organic monodisperse nanoparticles in microemulsion


AOTlheptane/water, DEA Bachelor Thesis, FUNDP.
2. Debuigne, F., Jeunieau, L., Wiame, M., B.Nagy, J. (2000) Synthesis of organic nanoparticles in different
W /O microemulsions, Langmuir 16. 760S-76 I 1.
3. Rouviere, J., Couret, 1M., Lindbeimer, L., Dejardin, J.L., Marrony, R. (1979) Structure des agregats
inverses d'AOT,J. Chern. Phys. 76,289-296.
4. Jeunieau, L. (1999) Adsorption of spectral sensitizers on silver halides nanoparticles, PhD Bachelor
Thesis, FUNDP.
S. Monnoyer, P., Fonseca, A, B.Nagy, J. (199S) Preparation of coUoidal AgBr particles from
microemulsions, Colloids Surf, 100,233-243.
STUDY OF THE ANOMALOUS SOLUBILITY BEHAVIOUR OF SOLUTIONS
SATURATED WITH FULLERENES BY 13C-NMR

A. DEMORTIER, R. DOOME, A. FONSECA, J. B.NAGY


Laboratoire de RMN,
Facultes Universitaires Notre Dame de la Paix, 61 rue de Bruxel/es, 5000
Namur, Belgium

1. Introduction

Since their discovery in 1985 [1], the fullerenes have attracted a great attention,
especially for one of their special properties: their unusual temperature dependence of
solubility in organic solvents [2,3], where a maximum occurs. Some of the systems
fullerene-solvent have been studied by HPLC [2,3]. Recently, an original method based
on I3C-NMR spectroscopy has been introduced in our laboratory [4]. Both methods
show a solubility maximum, which can be explained by the existence of solid solvates
formed by the fullerene and the solvent which decompose at the maximun temperature
[5,6]. Indeed, the solubility maximum corresponds to an incongruent melting point. The
first direct observation of such a phase transition was related in 1997 by A. Talyzin [7]
for C 60 in benzene while Masin et al [8] have carefully studied the IH nuclear spin
lattice relaxation in the C6o-benzene solvate. Talyzin has also characterised some
solvates by Raman spectroscopy [9]. In this contribution, the \3C-NMR results of the
solubility of the system Cwtetralin are presented and compared with HPLC results.

2. Experimental

C60 (99.5%) has been purchased from Dynamic Enterprises Ltd and 1,2,3,4-
tetrahydronaphthalene, 98+% (tetralin) is used as solvent. 11.4 mg of C60 has been
placed in a 5.6 mm outer diameter sealable Pyrex insert (Wilmad) in order to be put in a
7 mm Bruker NMR rotors connected to a vacuum line. The C60 has been dried under
vacuum (10. 6 Torr) at room temperature for 24 hours and evacuated at 473K for 48
hours (10.6 Torr) to remove oxygen as well as nitrogen from the C60 solid. The solvent
placed in a thick-glass container and connected to the same vacuum line has been
degassed down to a pressure of 10.6 Torr by using about ten freeze-pump-thaw cycles.
The solvent has then been transferred directly by distillation to the NMR tube, which
has been sealed on the vacuum line.
513
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 513-518.
© 2002 Kluwer Academic Publishers.
514

NMR measurements have been realized with a Bruker MSL 400 spectrometer operating
at 13C resonance frequency of 100.61 MHz (Bo = 9.31 T). Magic angle spinnp!g (MAS)
experiments have been run with a 7 mm Bruker sample holder rotating in the frequency
range of 1.0 kHz. All chemical shifts have been measured with respect to the I3C
resonance frequency of the ipso carbon of tetralin (136.9 ppm) used as an internal
reference. The temperature is measured by using a BVT-I000 controller. The
calibration of the temperature has been made with ethylene glycol using the correlation
between the separation of the ethylene glycol NMR absorption lines as a function of
temperature established by Kaplan et al [10] and Amman et al [11]. The spectra were
recorded by using a single pulse program with a 90° pulse of 5 J.1S and the repetition
time was set to 301 sec. At least 120 FIDs were accumulated for each spectrum to
ensure a good signal to noise ratio. To optimize the repetition time, I3C spin-lattice
relaxation time (T I ) were measured by using the inversion-recovery pulse sequence.
Between two measurements, the waiting time was of 6 hours to stabilise the system at
the imposed temperature.

3. Results and discussion

3.1 NMR RESULTS

3.1.1 Spectrum obtained at 295 K


Figure 1 shows the full NMR spectrum recorded at 295 K. Seven NMR lines can be
observed. The first ones are two triplets at 23.70 (± 0.01) ppm and 29.72 (± 0.01) ppm
corresponding, respectively, to carbons 2, 3 and 1,4 of the tetralin. At 125.59 (± 0.01)
ppm and 129.21 (± 0.01) ppm, two doublets stem from carbons 6 and 7 and carbons 5
and 8 of the solvent. The line of carbons 9 and 10 of the tetralin is at 136.9 (± 0.02)
ppm. At 143.04 and 143.59 ppm, there are the NMR lines of the C60 . It is possible to
distinguish the solid phase from the liquid phase because the diamagnetic anisotropy of
the aromatic carbons of the solvent induces a high-field chemical shift on the C60 in
solution. The solution phase of C60 has a chemical shift of 143.04 (± 0.03) ppm while
the solid phase is found at 143.59 (± 0.07) ppm.

A
\
I~

ISO 12s 100 7'5 5'0 is


Chemical shift (ppm)
Figure 1. NMR spectrum of the system Cwtetralin at 295 K.
515

3.1.2 Spin-lattice relaxation time (TIJ determination


We have measured the '3C spin-lattice relaxation times (T,) of the C60 in solid solvate
pha~e, the C60 in solution and the aromatic carbons of the solvent (C 9 and 10, Cs and 8 and
C 6 and 7) . These values were deduced from the evolution of the signal intensity as a
function of the recovery time (see Table I).
TABLE I. I3C spin-lattice relaxation times (T 1)

C60 C60 tetralin


solvate solution Cs and 8
Chemical shift (ppm) 143.59 143.04 136.91 129.2 125.6
Relaxation time/T, (s) 35.0 34.2 17.9 7.9 6.3

It is clearly seen that the repetition time fixed at 301 seconds corresponds to quantitative
conditions.

3.1.3 Evolution of the NMR spectra as a function of temperature


The temperature range investigated is between 295 K and 365 K.

IncreaSing temperature
In Figure 2, the evolution of the NMR spectra is shown with increasing temperature.
Below 341 K, two lines are observed, one at around 143.04 ppm which corresponds to
the C60 in solution while the other one, at around 143.59 ppm is due to the solid phase.
More exactly, taking the Smith theory [5] into account, this line corresponds to solvated
C 60 in solid phase. We can see that above 341 K a new line appears at 143.87 (± 0.02)
ppm which corresponds to C 60 with a less amount of solvent in the solid phase. Indeed,
the pure solid phase without solvent has a chemical shift equal to 144.1 ppm.
Considering these results, we can say that the phase transition occurs between 341 K
and 348 K.

365K
362K
355K
348K
341 K
326K

--o
309K
295K
I • I i i I I i I I
160 140 120 100 80 60 40 20
Chemical shift (ppm)
Figure 2. NMR spectra of the system C 60 -tetralin during increasing the temperature.
516

We can also see on figure 2, that the NMR line due to the C60 in solution becomes
narrower with increasing temperature. The narrowing is due to the faster movement of
C60 at higher temperatures.
Decreasing temperature
Figure 3 shows the evolution of the NMR spectra with decreasing temperature. In this
case, the phase transition is observed between 318 K and 334 K. The difference of the
results obtained during temperature increase and temperature decrease can be explained
by the fact that the formation of the solvate needs a much longer time because the
transition from C60 to C60 solvated crystal is rather difficult and slow. For example, the
time to form C 60 solvated crystal from pristine C60 is evaluated to be of about two
months. Hence, the results obtained during the temperature decrease do not stem from
an equilibrium state.

fG
~
V
~
0\,-
JV\
./

.-J

,...... V ~
.\:
146 144 142 140 138
365 K
362 K
355 K
348 K

----- 334 K
318 K
-[! 309K

-~~
, , I I
, I
, I
, I
, I
, I I
295 K

o
I
160 140 120 100 80 60 40 20
Chemical shift (ppm)

Figure 3. NMR spectra of the system C 60 -tetraiin during increasing the temperature.

3.2 DETERMINA nON OF THE SOLUBILITY OF C 60

First, we have determined the spin-lattice relaxation time (T 1) of the system C60 -tetralin
at 295 K. In order to work in quantitative conditions, we have chosen a repetition time
at least five times larger than the longest T(. However, we did not determine the T(
values at other temperatures. Thus, we have normalized the integrals of the C60 in
solution of all the spectra with the respect of the integral value of the solvent found at
each temperature (see equation 1).
517

I
I Coo in solution. T 1)
e60 in solution, nonnalized, T = ;(
(

I solvent,T
I solvent, 295K
In order to determine the solubility of C60 (g/ml) the following transformations have
been made: the intensities have been normalized to one carbon atom, thus dividing IC60
by 60 and Isolvent by 10, the so-obtained ratio has been multiplied by the ratio of the
molecular weight ofC6o (720) and that oftetralin (132); dividing the mass of the solvent
by its density yields the volume of the solvent. Finally, a normalization factor (NF) is
introduced which is equal to the so-determined solubility at 295K dividing by the HPLC
value obtained by Ruoff [12]:

So1ub}'l't Ie . I ' I' <l,T' MWe .10


} Y = 6OIDsoutlon,nonna.ze 60
,dens}'ty, NF (2)
Isolvent. T' MWsolvent' 60
The results are presented in Figure 4, A maximum of solubility is observed, which
means, that a phase transition occurs between a solvate of C60- tetralin and another solid
of C60 with less amount of solvent. The solubility maximum determinated by HPLC is
around 330 K [4] while the NMR value of the maximum is found between 341 K and
348 K, This observation can be explained by the fact that, in our conditions, we have
waited only six hours between the spectra at different temperatures, while in HPLC
analysis, the waiting time between two measurements was two days, Hence, in our
experience the system has probably not reached its equilibrium value after six hours and
the solvate was not completely decomposed.
30 .

25

~ 20
.§,
g
:.g 15
(5
f/J
-+- Solubility (by HPLC - Ooome et al (4))
to ~ Solubility (by NMR • increasing of temperature)

'" Solubility (by HPLC - Ruoff et alit 2()


5~ ______________ ~ ____ ~ __________________ ~

290 300 310 320 330 340 350 360 370


Temperature (K)
Figure 4. Solubility curves for the system C60-tetralin as a function of the temperature determined by HPLC
and by NMR.

4. Conclusion

I3C NMR of saturated solutions of C60 allows one to determine the solubility C60 and to
characterize the various phases. If aromatic solvents are used, the effect of the chemical
shift anisotropy of the solvent induces enough variations on the C60 chemical shift that
518

the various phases are well resolved. This way the C60 in solution (143.04 ppm), C60 in a
solid-solvate (143.59 ppm) and C60 in a solid but less solvated state (143.87 ppm) could
be distinguished. The solubility curve as a function of temperature passes through a
maximum. At low temperatures (up to ca. 341 K) the dissolved C60 is in equilibrium
with the C60 solid-tetralin solvate. This solvate decomposes at the maximum
temperature and could be transformed to pure C60 solid. However, the latter was not
observed, because the equilibrium conditions have not been reached in our experiments.
Note finally, that the solubility curve is not reproduced during decreasing temperature,
suggesting once again that the equilibrium conditions have not be reached.

5. Acknowledgement

This work was supported by the Belgian Program PAl P4/10 on Reduced Dimensionality Systems. A.
Demortier thanks F.R.I.A. for financial help.

6. References

I. Kroto, H.W., O'Brien, l.R., Curl, R.F. and Smalley, R.E. (1985) Buckminsterfullerene, Nature, 318,
162-163
2. Ruoff, R.S., Malhotra, R. , Huestis, D.L. and Lorents, D.C. (1993) Anomalous solubility behaviour of
C60 , Nalure, 362,140-141
3. Doome, R.J., Dermaut, S., Fonseca, A., Hammida, M. and B.Nagy, 1. (1997) New evidences for
anomalous temperature-dependent solubility of Coo and C'O fullerenes in various solvents, Ful/erene
Science and Technology,S, 1593-1606
4. Doome, R.J., Fonseca, A. and B.Nagy, J. (1999) IlC-MAS-NMR investigation of solutions saturated
with fullerenes : study of anomalous solubility behaviour, Col/aids and surfaces A, 158, 137-143
5. Smith, A.L., Walter, E., Korobov, M.V. and Gurvich, O.L. (1996) Some enthalpies of solution of Coo
and C'O. Thermodynamics of the temperature, dependence of fullerene solubility, J. Phys. Chern., 100,
6775-6780
6. Korobov, MY, Mirakyan, A.L., Avramenko, N.V., Olofsson, G., Smith A.L. and Ruoff R.S. (1999)
Calorimetric studies ofsolvates of Coo and C'O with aromatic solvents, 1. Phys. Chern. B, 103, 1339-1346
7. Talyzin, A.V . (1997) Phase transition C60- Coo*4C 6 H6 in liquid benzene, J. Phys. Chern. B, 101 , 9679-
9681
8. Masin, F., Grell, A. - S., Messari, l. and Gusman, G. (1998) Benzene-solvated Coo: IH nuclear spin-lattice
magnetic relaxation, Solid State Cornrnun., 106, 51-62
9. Talyzin, A. and Jansson, U. (2000) Coo and C'O solvates studied by Raman spectroscopy, 1. Phys. Chern.
B, 104,5064-5071
10. Kaplan, M.L., Bovey, F.A. and Cheng H.N. (1975) Simplified method of calibrating temperature
thermometric nuclear magnetic resonance standards, Analytical chernistry, 47, 1703-1705
II. Amman, c., Meier, P. and Merbach, A.E. (1982) A simple multinuclear NMR thermometer, J. Magnetic
Reson., 46, 319-321
12. Ruoff, R.S., Doris, S.T., Malhotra, R. and Lorents, D.C. (1993) Solubility of Coo in a variety of solvents,
1. Phys. Chern, 97,3379-3383
EPR STUDY OF POLYAMINE COPPER COMPLEXES

N.GUSKOS1.2, V.LlKODIMOS 3, J.TYPEK z, H. FUKS l , S. GLENIS i , M. WASIA l , C.


L. LlN 4 , E. GRECH s, T. DZIEMBOWSKA s,
A. PERKOWSKA 6
JSolid State Physics, Department of Physics, University ofAthens. Panepistimiopolis,
J5784 Zografos, Athens, Greece
1Institute of Physics, Technical University of Szczecin, AI. Piastuw J 7, 70-310 Szczecin,

Poland
J Departement of Applied Mathematics and Physics, National
Technical Universily, J5780 Athens, Greece
'Department of Physics, Temple University, Philadelphia,
PA 19122, USA
)Institute of Fundamental Chemistry, Technical Universily of
Szczecin, AI.Piastow 17, 70-310 Szczecin, Poland
6Institute of Bioorganic Chemistry, Polish Academy of Sciences,

ul.Noskowskiego 12114,61-704 Poznan, Poland

1. Introduction

Organometallic complexes of naturally occurring polyamines (biogenic amines) attract particu-


lar interest due to their implication in a number of biological processes. In particular, the centers
of polyamine interactions with other bioligands constitute potential coordination sites of metal
ions present in living cells, while polyamines may be treated as the interfering factor in the for-
mation of complexes between metal ions and nucleic acids [1,2]. Considerable research effort
has been accordingly devoted to crystallographic and spectroscopic studies of crystalline poly-
amine complexes with various inorganic cations and anions, in order to determine the structural
conformation of polyamine cations and explore the relation between protonation and metalation
of biogenic polyamines [3-5] .
Recently, electron paramagnetic resonance (EPR) combined with photoacoustic and optical
spectroscopies have been fruitfully exploited to resolve the electronic properties of Cu 2+ ions in
polycrystalline polyamine complexes with sulfate and nitrate salts of bivalent copper [6,7]. In
these complexes, the coordination sphere of Cu 2+ ions comprises four equatorial nitrogen atoms
and one or two axial oxygen atoms provided by the water molecules or the inorganic anions [3-
5]. The molecular structure of many six and five-coordinated Cu 2+ complexes is known to be
amenable to static and dynamic lahn-Teller effects, while significant complications may arise in
crystalline complexes due to the presence of exchange interactions among the differently ori-
ented metal sites [8-12] . To this aim, a variable temperature EPR study has been carried out in
order to trace the interplay of exchange coupling and possible dynamical effects in the molecu-
lar structure of three polyamine copper sulfate complexes.

2. Experimental

Polycrystalline samples of the three investigated polyamine complexes, aqua spermine copper
sulfate trihydrate [NHz(CHz)3NH(CHz)4NH(CHz)3NHlCUz+H20]SO/ 3H20, norspermine cop-
per sulfate trihydrate [NH2(CHz)3NH(CH2)3NH(CHz)3 NH 2CU 2+'SO/"PH 20, and homospermine
copper sulfate dihydrate [NHz(CH2hNH(CH2)2NH(CHz)3NH2CU2+S0/]2H20, designated
hereafter as spm343, spm333 and spm323, respectively, were prepared as previously described
[4,5].
519
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 519-523.
© 2002 Kluwer Academic Publishers.
520

[ ] SO.' H ,0

Figure I. The schematic structure of the three investigated com-


plexes
Splft tHJJ

,-,
[ C ·:;/·~
..
• J......., -.-/
] 'H ,0
The former two complexes, which were obtained in
s,. () l) I single crystal form, crystallize in the monoclinic sys-
tem (P21Ic) with four symmetry-related Cu 2+ sites in

[ ] 'H ,0
each unit cell (2=4). In both structures, the Cu 2+ ion
is five-coordinated forming a tetragonal pyramid
with four equatorial nitrogen atoms from the amino
s,. OJ) I
groups and one oxygen atom provided by the water
molecule (spm343) or the sulfate anion (spm333), leading to the formation of tricyclic chelated
complexes [4,5].
EPR measurements were carried out on fine powder samples, sealed in quartz tubes, using a
standard X-band spectrometer Bruker E500 with 100kHz magnetic field modulation. The mag-
netic field was scaled with a NMR magnetometer. The measurements were performed in the
temperature range of 4.1 to 250 K using an Oxford flow cryostat.

3. Results and discussion

Figure 2 shows representative EPR spectra for the spm343 complex at various temperatures. At
higher temperatures (T> 100 K), the EPR spectrum is dominated by an anisotropic powder pat-
tern characteristic of Cu 2+ ions in concentrated compounds, where superexchange interactions
between adjacent ions cause motional narrowing of the copper hyperfine structure. Simulation
of the EPR spectrum for S=1/2 with rhombic g-tensor and anisotropic linewidths shows nearly
axial symmetry with a small rhombic distortion (g3=2.214>g2=2.052"'gl=2.040), in accord with
the analysis of the EPR spectrum at room temperature (6).
g,
I
x10 200 K

xS .O 6SK

x2.S 30 K

x1 .2 10K

xl.0 4.9 K

2800 3000 3200 3400 3600

H (Gauss)
Figure 2. Temperature dependence of the X-band EPR powder spectrum for the spm343 complex (1'-=9.4501 GHz).
The arrows indicate the resonance fields corresponding to the g-values.

The experimental g-values comply with the expected anisotropy of the molecular g-tensor
(gz>gr:gx>2.002) for the distorted square pyramidal coordination of Cu 2+ ions in the spm343
complex. They have been accordingly used along with spectroscopic data, to determine the elec-
521

tronic ground state ofCu 2+ ions, which is characterized by a singly occupied orbital with
d(x 2_/)
a small admixture of d(3i-r2) [6]. Howevt:r, additional weak EPR lines, whose relative intensity
increase as the temperature decreases, can be clearly evinced. Two broad, shoulder-like features
are observed at intermediate g-values, g:::2.14 and g",2.09, while a sharp peak appears gradually
at the high field side of the gl transition. The latter EPR line shifts continuously towards higher
magnetic fields upon lowering the temperature from 10K. Similar EPR lines at intermediate g-
values have been reported in the powder EPR spectra of six and five-coordinated copper com-
plexes, which have been assigned to the cooperative g-tensor of exchange coupled 'Ci+ ions at
differently oriented, and thus spectroscopically, non-equivalent sites [8, I 0, 12]. Furthermore, the
high field EPR line, which becomes clearly resolved below 10K, implies that exchange interac-
tions become effective causing the shift of the resonance field with decreasing temperature.

x8 233 K

x4 90K

x2 45K

x1 9.8 K

x1 5.0 K

2800 3000 3200 3400 3600


H (Gauss)

Figure 3. Temperature dependence of the X-band EPR powder spectrum for the spm333 complex (Jr-9.4496 GHz).
The arrows indicate the resonance fields corresponding to the g-values.

Figure 3 shows EPR spectra for the spm333 complex at various temperatures. In the whole
temperature range, the EPR spectrum comprises a dominant Cu 2+ anisotropic powder pattern
described by a rhombic g-tensor with principal g-values (g)=2.206>gl=2.084>gl=2.040), similar
to those previously determined at room temperature [6]. However, at higher temperatures, an
additional resonance line of weak intensity is resolved at g=2.065(3), close to the g2 transition
(Fig. 3). At temperatures below 65 K, this EPR line merge~ with the adjacent (g2) resonance
line, which becomes slightly shifted and most importantly gains drastically intensity causing a
substantial lineshape distortion of the rhombic EPR powder pattern, maintained down to the
lowest temperature (Fig.2). Further distortion of the total EPR lineshape is observed at the in-
termediate region between g) and g2 at lower temperatures, which may be partly due to the pres-
ence of preferential orientation induced by the applied magnetic field [13]. The total EPR inten-
sity decreases below 10K indicating the presence of appreciable anti ferromagnetic interactions
among the underlying Cu 2+ ions. Although higher frequency EPR measurements are clearly
needed to resolve the individual components of the EPR spectrum and the presence of molecular
and cooperative g-values [12], it appears possible that exchange coupling between the differ-
ently oriented Cu 2+ sublattices is operative from relatively high temperatures.
522

x12 ;20K

x8.0 110K

x3 .5 45 K

x1.2 10.2 K

x1.0 4.65 K

2800 3000 3200 3400 3600

H (Gauss)
Figure 4. Temperature dependence of the X-band EPR powder spectrum of spm323 complex (1'=9.4493 GHz). The
arrows indicate the resonance fields corresponding to the g-values.

In the case of the spm323 complex (Fig. 4), the EPR spectrum exhibits an anisotropic powder
pattern with nearly axial g-tensor and a small rhombic distortion
(g)=2.181 >g2=2.061 ~gl=2.052), resulting in a relatively small admixture of the d(3i _r2) orbital
in the d(x 2_/) ground state wavefunction (6). As temperature decreases, an additional compo-
nent at the high field side of the gl transition becomes gradually resolved below 10K. The latter
resonance line, which can be evinced even at 220 K as a shoulder of the gl transition, shifts with
decreasing temperature, while a broad shoulder appears on the low field side of the g2 transition .
Both these effects are qualitatively similar to the low temperature EPR spectrum of the spm343
complex (Fig. I ), indicating that the latter EPR lines may be due to a fraction of exchange cou-
pled Cu 2• ions with stronger interactions than those responsible fo r the rest of the EPR spec-
trum.
According to the crystallographic analysis of the spm343 and spm333 complexes [4,5], the
five-coordinated Cu 2• ions are connected through an extended hydrogen bond network where
water molecules and sulfate anions act as bridges. In particular, for the spm343 complex, which
shows the most complex EPR spectrum (Fig.2), the axially coordinated oxygen atom belongs to
the strongest bonded water molecule that is also hydrogen-bonded to a sulfate anion and another
water molecule [4]. Diverse superexchange interactions among Cu 2• ions may be then realized
depending on the strength of the hydrogen bonds, leading to the simultaneous presence of dif-
ferent fractions of exchange coupled Cu 2• ions. However, further investigations are clearly
needed to disentangle the complex EPR spectra and the nature of exchange interactions between
the Cu 2• ions in the molecular structure of polyamine complexes.

4. Conclusions

A variable temperature EPR study of three polycrystalline polyamine copper complexes indi-
cates the presence of diverse exchange interactions among the underlying Cu 2• ions. Complex
EPR spectra are observed indicating the simultaneous presence of molecular and cooperative g-
tensors, whi le appreciable antiferromagnetic interactions become operative at temperatures be-
low approximately 10K for a part of the Cu 2• sublattices.
523

References

I. Lomozik, L. (1995) in Berton G. (ed.) Handbook oj Metal-Ligand Interactions in Biological Fluids, Marcel
Dekker, New York, vol. I, p. 686.
2. Lomozik, L. and Gasowska, A. (1998) J. Inorg. Biochem. 72,37-47.
3. Surma, K.. Perkowska, A., and Gawron, M. (1994) Polish), Chern. 68,481-488.
4. Maluszytiska, H., Perkowska, A., and Skrzypczak-Jankun, E. (1995»), Chern. Crystal. 25, 19-23.
5. Perkowska, A. and Maluszytiska, H. (1999) J. Mol. Struct. (fHEOCHEM) 508, III- \ \ 7.
6. Guskos, N., Papadopoulos, G.1., Likodimos, V., Mair, G.L.R., Majszczyk, J., Typek, J., Wabia, M., Grech, E.,
Dziembowska, T., and Perkowska, A., (2000) J. Phys. D: Appl. Phys. 33, 2664-2668.
7. Guskos, N., Papadopoulos, G., Likodimos, V., Majszczyk, J., Typek, J., Wabia, M., Grech, E., Dziembowska,
T., Perkowska, A., and Aidinis, K. (2001»), Appl. Phys. 90,1436-1441.
8. Reinen, D. and Friebel, C. (1979) Structure and Bonding 37, 1-60.
9. Ammeter, J.H., Burgi, H.B., Gamp, E., Meyer-Sandrin, V., and Jensen, W.P. (1979) Inorg. Chem. 18,733-750.
10. Henke, W., Kremer, S., and Reinen, D. (1983) Inorg. Chern. 22,2858-2863.
II. Reinen, D. and Friebel, C. (1984) Inorg. Chern. 23,79\-799.
12. Sastry, B. A., Madhu, B., Ponticelli, G., and Massacesi, M. (1988) J. Phys. Soc. Japan 57, 1140-1141.
13. Hulliger, J., Zoller, L., and Ammeter, J.H. (1982).1. Magn. Res. 48, 512-518.
ANISOTROPY OF TRANSVERSE IH MAGNETIZATION RELAXATION IN
STRAINED ELASTOMERS BY THE NMR-MOUSE®

K. HAILU, R. FECHETE, D. E. DEMCO, and B. BLUMICH I


Institut for Technische Chemie und Makromolekulare Chemie, Rheinisch-
Westflilische Technische Hochschule, WorringelWeg 1, D-52056 Aachen,
German/

Abstract

NMR-MOUSE® is a single-sided NMR sensor suitable for measurements of the


segmental anisotropy induced in large elastomer samples by uniaxial forces or local
strain. Proton transverse nuclear magnetic relaxation was measured by Hahn-echo
decays in cross-linked natural rubber bands to analyse its dependence on the angle
between the direction of the uniaxial stretching force and the Z axis defined by the
direction of the polarising field Bo NMR-MOUSE®. A qualitative understanding of this
dependence is gained by an analytical theory of the transverse magnetization relaxation
adapted to the case of stretched elastomers and strongly inhomogeneous magnetic
fields. The Solid-like and liquid-like spin system response was considered for a pair of
spins ~ in a network. Numerical simulations are also shown which include the statistics
and orientation distribution of the end-to-end vector and the dependence on the
elongation ratio.

1. Introduction

Over the last few years, several new NMR applications have been developed for use in
strongly inhomogeneous static and radio-frequency magnetic fields. Applications
include materials testing (1-4), well logging for oilfield applications (5), and stray field
NMR (6). It has been shown that for protons various NMR parameters like TI (3), Tz
(2,7,8), TIp (2,9), and self-diffusion coefficients (10) can be measured under these
conditions. Moreover, the possibility to excite IH double-quantum (DQ) coherences in
strongly inhomogeneous fields was also shown (11). also it was demonstrated that the
NMR-MOUSE® (NMR-MOUSE® is a trademark of RWTH) can be employed to
investigate the anisotropy of the IH transverse relaxation of Aachilles tendon in vivo
[12].
Recently, a mobile NMR surface scanner has been developed for the nondestructive
investigation of arbitrarily large objects (1-4). The NMR-MOUSE® is a palm-size NMR

I To whom correspondence should be adressed


525
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 525-530.
© 2002 Kluwer Academic Publishers.
526

device which is built up from two pennanent magnets. The magnets are mounted on an
iron yoke with anti-parallel polarization to fonn the classical horseshoe geometry. The
main direction of the polarization field Bo is across the gap, and the field strength
decreases rapidly with increasing distance from the surface. Along the gap changes in
Bo are less pronounced than perpendicular to it due to the shape of the magnets. The
radio-frequency (rf) field Bl is generated by a surface coil which is mounted in the gap.
The magnetic field ,Frofiles define the size of the sensitive volume, which is specific to
the NMR-MOUSE geometry as well as the spectrometer carrier frequency (7). The
NMR-MOUSE® is characterized by strong inhomogeneities in the static and radio
frequency magnetic fields. Due to these inhomogeneities the task to implement specific
NMR techniques is not trivial.
The aim of this communication is to show another relevant application of the NMR-
MOUSE ® in measuring the anisotropy of the polymer chains in a stretched cross-linked
elastomer via the lH transverse magnetization decay.

2. Hahn-echo decays

The segmental anisotropy in uniaxialy stretched elastomer network has an influence on


the lH Hahn-echo decay by the NMR-MOUSE. To show this we consider in the
following the simplified case of an isolated pair of spins ~ attached to the polymer
chains in a cross-linked network. The preaveraged dipolar Hamiltonian can be written
as a sum of a static contribution Hd and a fluctuating one L1Hd(t) , which commute at
any moment of time. Hence the magnitude of thetransverse magnetization decay is
given by
Sm (r, t) ~ (S::lid (r, t )s~qUid (r, t)), (1)
where S,:olid(t) and S:.:qUid(f) represent the solid-like and liquid-like contribution to the
NMR signal. The presence of dangling chains and free sole chains was neglected.
Using the density matrix operator evaluation of the spin system response to a Hahn-
echo pulse sequence: 8 x-t-28 x-t-echo for a voxel at position r
characterized by the
pulse angle 8 we obtain for the quadrature Hahn-echo decays
S;Olid (0, r, f) oc sin B cos 2B cos 2 ~M 2 t T~ (2)
and
S;Olid (0, r, t) oc sinB cos 2B sin (M2 t t ~
2 (3)
where the residual second van Vleck moment is given by

(M2) = (% D'" s;,,)' ([ q' H( 2A' +~X~ P,( cose)p,( cos ll(r))P2( COSPF)+~)-±}+ K( A' -+)])' (4)

The physical quantities present in Eqs. (2)-(4) are the folIoing: D(2) is the dipolar
coupling for methylene protons, S~2) is the scaled dynamic order parameter [14],.1 is
the elongation factor, K describes the effective interchain interactions [15], ij' is the
reduced end-to-end vector for a stretched network [14]. The second-order Legendre
polynomial is Pix) where the angles present in the arguments are defined in Figs. 1 a
527

and b. The angle between the uniaxial stretching force F and the i direction (Fig. Ib)
is denoted by 8 and is an experimental variable.

F f-
8. (1) R

~ ~r t

(a) (b)

Fig.l: (a) The relative orientlltion of the end-to-end vector il' , direction of the
uniaxial force F and the local direction of the static magnetic field Bo(r).
(b) The relative orientation of the Z, F , and Bo(r) vectors
for the NMR-MOUSE<il geometry.

The contribution of the fluctuating dipolar couplings to the Hahn-echo decay for a voxel
F in the sample can be evaluated using the results presented in reference [16]. We can
finally write

S:;qu;d (8, F,t)oc iSin B cos 2B(exp{- fJcosh{JJ). (5)


and

S~:qU;d (8, F,t)oc sin Bcos 2B(exp{- f. {iCOSh{J2}- sinh {JJ]), (6)

The function f, and f2 are derived in the approximation of a single correlation time for
the slow motion of the chains and are given by
j, = 2M,r:[exp(- r/rJ-l + r/r,], (7)
f, = M,r,'[l- 2exp(- r/r,)+ exp(- 2r/rJl. (8)

3. Simulation

The residual van Vleck second moment for a stretched network (Eq. 4) is represented in
Fig. 2 versus the orientation angle 8 for Ss(2) = 0.02 and K = 0.05 [14, 15]. This NMR
parameter assumes a minimum near the magic angle 8 = 54.7°. In the same figure the
values of the <M2(8» for a stretched natural rubber band obtained from ~ function [16]
are also shown. Because of I1T2eff oc <M2>1/2 this simulation is in agreement with the
results shown in Fig. 3.
528

";'",
~9 5.0

~
Ii' 4.5
V

4.0

El[")
Fig. 2. Angular dependence of <M2(0» following Eq. (4). The
points correspond to the residual van Vleck moment derived from
the Pfunction [16].

4. Experimental

The investigated system was unfilled and slightly cross-linked natural rubber (1.5
sulfur/ 1.5 TBBS (benzothiazyl-2-tert-butyl-sulfenamide» . After uniaxially stretching to
the desired lengths the rubber bands were clamped between two pieces of high-grade
steel which were fixed at frame built from aluminum to keep the extension constant.
The NMR experiments were performed by using the palm-size NMR-MOUSE® and a
commercial Bruker Minispec. The NMR sensor was tuned to 20 MHz. The volume of
the samples covered than the sensitive volume defined by the geometry of the rf-coil (8
mm diameter) and by the gap of the permanent magnet of 12 mm.
The IH transverse magnetization relaxation measurements were conducted using the
Hahn-echo decay. The echo time was varied in the range from 0.05 ms to 7 ms. All
measurements were carried out at the room temperature 22°C ± 1°C.

5. Results and Discussions

The angular dependence of lIT2efT was measured for different extension ratios A ranging
from I to 11 (Fig.3). In the polar plots the effective transverse relaxation rates for A > I
do not exhibit a circular pattern like for relaxed state A =1. It is evident form Fig. 3 that
the increased segmental anisotropy leads to shorter transverse relaxation times.
Moreover, with increasing strain the angular dependence becomes more pronounced. In
agreement with the simulation a minimum value of I/T2efT was detected between 50° and
60° near the magic angle of 54.7°. This reflects the distribution in the orientation of Bo
field for the NMR-MOUSE® sensor.
529

1.6
,,-,~:--- A• II
1.2
Vk....,-~;--A · S
o.! ~'<-""""'- A - 4
C 0._ ""'~~---4r- A p 2.
•. ;/VT----'9r--+- A - l.
~ o.o m t-it~~~~~7t==~~~A ~ 1
Jo. .
Q'
1.2
1.6

""
Fig. 3: Polar plot represen tation of the angular dependence of I/ T2<lTfor differen lt
values of the ex tension ratio A.

The angular dependence of the Hahn-echo amplitude at different echo times for the
natural rubber band with A = 8 is shown in Fig. 4.
In agreement with the simulation, only in the region dominated by the liquid-like
relaxation the anisotropy is expressed in the measured transverse relaxation rates.

1.36
1.32

~128
124 OO/lS

~:
]1.12
~l)!
!3I.0I
100 1--,......--,~~-,--.--,-...,..-l--,--,-~-,--.--,-....J

0[0]

Fig. 4: Angular dependence of the amplitude of the Hahn echo for


different echo times t for a natural rubber band stretched to A = 8.

6. Conclusions

The anisotropy of the Hahn-echo decay reflectin~ the segmental anisotropy in stretched
elastomers was measured using NMR-MOUSE sensor. The anisotropy is increasing
with the elongation ratio and was described by a semi-quantitative theory based on the
transverse relaxation of polymer networks. These results can be used for understanding
the anisotropy of various NMR parameters measured in ordered tissues.
530

Acknowledgements

This work was supported by grant from the Deutsche Forschungsgemeinschaft (DE
780/1-1 ).

References

I. G. Eidmann, R. Savelsberg, P. Bliimler, and B. Bliimich, The NMR-MOUSE, a mobile


universal surface explorer, (1996) 1. Magn. Reson . A 122, 104-109.
2. G. Guthausen, A. Guthausen, F. Balibanu, R. Eymael, K. Hailu, U. Schmitz, B. BHimich, Soft-
matter analysis by the NMR-MOUSE, (2000) Macromol. Mater. Eng. 276/277,25-37.
3. A. Guthausen, G. Zimmer, P. Bliimler, B. Bliimich, Analysis of polymer materials by surface
NMR via the MOUSE, (1998)1. Magn . Reson. 130, 1-7.
4. B. Bliimich "NMR Imaging of Materials", (2000) Clarendon Press, Oxford.
5. R.L. Kleinberg, Encyclopedia of Nuclear Magnetic Resonance, (1996) volume 8 chapter Well
loging, pages 4960-4969, John Wiley & Sons, Chichester.
6. P. 1. McDonald, Stray field magnetic resonance imaging, (1997) Prog. Nuc!. Magn. Reson.
Spec/. 30,69-99.
7. F. Balibanu, K. Hailu, R. Eymael , D. E. Demco, B. Bliimich, Nuclear magnetic resonance in
inhomogeneous magnetic fields, (2000) 1. Magn. Reson. 145,246-258.
8. M. D. Hiirlimann, D. E. Griffin, Spin Dynamics of Carr-Purcell-Meiboom-Gill-like sequences
in grossly inhomogeneous Bo and BI fields and application to NMR well logging, (2000) 1.
Magn. Reson. 143, 120-135.
9. A. Wiesmath, R. Fechete, H. Kiihn, D. E. Demco, B. Bliimich, Integrated dispersion of spin-
lattice relaxation time in the rotating frame in strongly inhomogeneous magnetic fields, (in
preparation).
10. M. D. Hiirlimann, Diffusion and relaxation effects in general stray field NMR experiments,
(2001)1. Magn .Reson. 148,367-378.
1I. A . Wiesmath, C. Filip, D. E. Demco, B. Bliimich, Double-quantum filtered NMR signals in
inhomogeneous magnetic fields, (2001) 1. Magn. Reson. 149,258-263.
12. R. Haken, B. BHimich, Anisotropy in tendon investigated in vivo by a portable NMR scanner
the NMR-MOUSE,J. Magn. Reson. (2000) 144, 195-199.
13. P. Sotta, C. Fiilber, D.E. Demco, B. Bliimich, H.W. Spiess, (1996) Macromolecules, 29, 6222-
6230.
14. M. Schneider, L. Gasper, D. E. Demco, B. BHimich, Residual dipolar couplings by IH dipolar
encoded longitudinal magnetization, double- and triple-quantum nuclear magnetic resonance
in cross-linked elastomers, (1999) 1. Chern. Phys. 111,402-415.
15. M.G. Brereton, NMR study of molecular anisotropy induced in a strained rubber network,
(1993) Macromolecules, 26, 1152-1157.
16. R.C. Ball, P.T. Callaghan, ET Samulski, A simplified approach to the interpretation of
nuclear spin correlations in entangled polymer liquids, (1997), 1. Chem. Phys., 106, 7352-
7361.
SYNERGY PHENOMENON IN BULK RUTHENIUM- VANADIUM
SULFIDES:
SIV NMR and ESR studies

R Hubaut 1, A. Rives I , O. Lapina2, D. Kbabilulin2, C.E. Scott3


l.Universite des Sciences et Technologies de Lille, Laboratoire de Catalyse
de Lille, ESA CNRS n080tO, Batiment C3, 59655 Villeneuve d'Ascq, France
2.Boreskov Institute of Catalysis, pr.Lavrent'eva, 5, 630090 Novosibirsk,
Russia
3.Universidad Central de Venezuela, Centro de Catalisis Petr6leo y
Petroquimica" Apartado Postal 47102, Los Chaguaramos, Caracas,
Venezuela

1 Introduction

For a better protection of the atmospheric environment, catalytic hydroprocessing


will become increasingly important in the years to come. This is due to the need for
light distillates simultaneously with a stricter environmental legislation. Since the
works ofChianelli and Harris [1], it is well known that mixing two transition metal
sulfides (TMS) allows developing higher activity than the ones exhibited by pure
compounds in hydrotreatment reaction. From this point of view, the CoMo and
NiMo conventional catalysts have been widely used. So far, an explanation for such
a synergism is still matter of debate and understanding of the synergy phenomenon
becomes fundamental. Our previous results [2,3] show that, rather a structural effect,
we are in presence of electronic effect, which changes noticeably the environment of
active site. As an example, all the iron-based mixed sulfides present, in Mossbauer
spectra, only one iron species, which is Fe2+ low spin. Moreover, it is showed that
there is a linear correlation between the hyperfine field distribution and the activity
of vanadyloctaethylporphyrine hydrodeporphyrinisation. The evolution of the
<HPF> at room temperature for different relative concentration of iron is due to the
changes in the density of state of the d electrons at the Fermi level [4]. Magnetic
resonance appears as one of the technique that can give a part of the response. As
ruthenium sulfide is one of the best catalyst for hydrodesulfurisation (HDS) [5] and
V is also a good candidate for the synergy phenomenom [2], bulk RuV mixed
sulfides were chosen to performed a study by solid state NMR and ESR.

2 Experimental

Bulk NiV, FeV and RuV mixed sulfides with different atomic ratios were prepared
by homogeneous solid precipitation method (HSP). Two series of catalysts were
synthesized. Ruthenium sulfide solutions (RuCh, xH20 (Aldrich) in water for the
531
J. Fraissard and O. Lapina (eds.J, Magnetic Resonance in Colloid and Interface Science, 531-536.
© 2002 Kluwer Academic Publishers.
532
first (X=I) preparation and RuCh (Strem chemical) in ethanol for the second (X=2»
were slowly added to a solution of tetrathiovanadate (ATTY, Strem chemical) in
ammonium sulfide (20% in water, Strem chemical) under nitrogen atmosphere. The
amount of the two solutions was worked out to obtain the expected atomic ratio
Ru/(Ru+ Y) in the final mixed sulfides. Pure vanadium sulfide was also prepared
from ATTY. After stirring, the solutions were filtered off and the obtained solids
were dried at room temperature under nitrogen. At the end, the catalysts were
sulfided at 400°C for 12 hours under hydrogen sulfidelhydrogen flow (10/90).
Catalysts are called XRuYY, where X is the preparation mode and Y is the atomic
percentage of ruthenium.
X-ray diffraction patterns were obtained, with a Siemens D5000 diffractometer (Cu
anode with monochromator). The ESR spectra were recorded on a Bruker ER-200D
spectrometer with a microwave frequency of 9.2 GHz (X-band). The samples were
placed into a double cavity, which permitted the measurement of a sample and a
reference under identical conditions, in particular microwave power. The field
modulation frequency was 100 kHz at constant modulation amplitude of 1 G. A
solid solution of Mn2+ in MgO (hyperfine structure with six lines) served as a
reference material. For quantitative calibration of the spectra a CuCh single crystal
standard was used. The accuracy of the quantitative determination of y4+ centres by
ESR was approximately 2 x 10 17 sites/g of catalyst. All spectra were measured at
77K in 3mm ~uartz ampoules. For spectra simulation, the computer program ESRI
[6] was used. IYand IH NMR measurements were performed using a Bruker MSL-
400 spectrometer at 105.25 MHz and 400.13 MHz, respectively, at room
temperature. Sly MAS spectra of powder samples were recorded at rotation
frequencies between 10 kHz and 15 kHz using a 5 mm rotor and the NMR probe
constructed by NMR Rotor Consulting ApS, Denmark. In some experiments spectra
were recorded on a Bruker 2mm MAS probe at rotation frequencies 30 - 35 kHz.
Repetition times from 0.1 s to 2 sand r.f. pulses with I ~s duration were used in the
experiments. All chemical shieldings are referred to YOCI3 as an external standard.
Simulations of Sly static and MAS NMR spectra were performed taking into
account the second-order quadrupole effects and using the general purposes NMRI
program [7]. The NMRI program is based on an effective average Hamiltonian
obtained in a manner similar to that presented earlier [8]. A particular variant of this
program (NMR2) especially adopted for the fast computation of spinning side band
(ssb) intensities in MAS spectra of quadrupolar nuclei -Herzfeld and Berger
approach [9]- was used for simulations of MAS and static spectra and for least-
squares parameter fitting. Computations were performed on a dual Pentium II 450
MHz IBM PC compatible computer running Linux OS.

3 Results and discussion

Pure Vanadium sulfide has X-ray diffraction pattern close to Y3 S4 and pure
ruthenium sulfide corresponds to RUS2' For the Ru-Y mixed sulfides, only RUS2 are
sufficiently crystallised to give diffraction patterns.
533

v( = 15 kHz '" Static


)
Experimental \
_ __ ._._ -. r- ~~_._._ .. __

,
Simulated " . .,
I I I I I
- - - ---.-
,J
I
- --~ - --
I
I I
3000 ppm ·3000 -6000 3000 0 · 3000 · 6000
ppm

Figure I :MAS and static Sly NMR experimental and simulated spectra of(N~h YS,

Spectra of V3S4 and (NHt)3 VS 4 were performed in preliminary studies. The ESR
spectrum ofV 3S4 exhibits weak anisotropy with giso=1.955. There is no a lot of data
devoted to SIV NMR studies of vanadium sulfides. Nevertheless, according to the
literature data [10-12], the wide range of SIV chemical shifts is typical for sulfido-
vanadates (V). In solutions SIV NMR signals of sulfido-vanadates are significantly
shifted to the low field up to 1600 ppm. ([VS 4]3-, [HVS4t, [V 2S7 are observed in t
the range +1400 to +1575 ppm). In the solid state their chemical shifts depend on the
metal (for CU3[VS4] chemical shift is equal to 620 ppm, while for Th[VS 4] it is equal
to 1525 ppm). SIV NMR spectrum of commercial ammonium tetrathiovanadate is
shown in Figure 1. It demonstrates axially anisotropic line with well-resolved
spinning sidebands in MAS spectra. Analysis of the intensities of the spinning
sidebands by SATRAS allows estimate all NMR parameters for this compound
(tablel).

TABLE I. NMR parameters of (NHt)3 VS 4

CQ 8" 0'00 Ojl (11. aPr


(kHz) 'Ie (ppm) 'I" (ppm) (ppm) (ppm) (deg) (deg) (deg)

2140 0.62 838 0.006 1490 1070 2330 -29 59 -41

The isotropic chemical shift for (NHt)3 VS 4 is equal to -1490 (± 30 ppm depending
on rotation frequency, sample temperature), quadrupolar constant of 2.1 MHz is
typical for tetrahedral vanadium coordination. Large value of asymmetry parameter
indicates a tetrahedral distortion of VS 4 in this compound. According to the
structural data (NHt)3 VS4 belongs to A3BC 4 type which possess exceptionally
simple structures. Depending on the nature of A atom, these solids crystallize in
different, though closely related cubic structures. CU3 VS 4 has space group P43m,
with all equal V-S distances of 2.215A. Th VS 4 possess structure with the space
group 143m. In all A3 VS 4 compounds, S tetrahedrally coordinates vanadium and
what is the most important, vanadium atoms occupy positions of high point
534

symmetry, Td. As a consequence, quadrupolar interactions should vanish at the sites


of vanadium nuclei. ~)3VS4 has space group Pmna, vanadium-sulfide distances
varies from 2.137 to 2.174 A (table 2). The latter reflects in the chemical shielding
anisotropy of 838 ppm for this compound.

TABLE II. Sructural and NMR parameters of some tetrathiovanadates M3 VS 4


(M=TI, Cu, NRt)

Type of distance Ojso


Compound Space group
tetrahedron v-S,A (ppm)
ThVS4 Regular 2.173 143m -1575')

CU3VS. Regular 2.215 P43m -620')

~)NS4 Disorder 2.137 Pmna 1490


2.150
2.153
2.174

a) NMR data from [II]

These data clearly indicate that in comparison with vanadium-oxide compounds


vanadium-sulfides compounds are characterised by a large chemical shielding range,
which is very sensitive to the nature of the atoms in the second coordination sphere
as well as to the distortion of vanadium sites. As a consequence, an observation of
SIV NMR signal at -1490 ppm with anisotropy of 838 ppm for solid (NRt)3VS4 is
not surprising. Moreover it is clear, that less symmetric structures, which could be
formed in real catalytic system, will possess larger value of anisotropy.

a) b)

gIl = 1.95

_.--- J
g.L=1.90

v----./~
, , , , ,
3300 3400 3500 3600 3700 , ,
G JXX) G 41D

Figure 2: ESR spectra of FeV25, NiV25, IRuV25 a) before sulfidation and b) after sulfidation
535
SIV NMR spectra of the catalysts FeV25, NiV25, lRuV25 are similar to the ones of
(NH4)3 VS4. There is no interaction between Fe (Ni, Ru) sulfide and (NH4)3 VS4 at
this stage of preparation. The same conclusion follows from ESR data (Figure 2).
Both ESR and NMR spectra change dramatically after catalysts sulfidation. Broad
lines are appeared in ESR spectra (Figure 2). The disappearance of the signal in
NMR spectra could be explained by significant distortion of vanadium sites as well
as by their reduction. .
ESR spectra of the catalysts prepared with the second preparation method show that
for the mixed 2RuV25 and 2RuV75 sulfides, a new line with gx = 1.962, gy = 1.962,
gz=1.91, Ax = 74 G, Ay = 74 G, Az = 197 G from isolated vanadium (IV) sites is
appeared in addition to the line of V3S4 (Figure 3). The intensity of the latter line
increases with the increase of Ru content. Catalysts sulfidation, as in the first set of
samples, changes dramatically ESR and NMR spectra. It seems reasonable to
explain the observed spectral changes by the formation of Ru-V-S complexes with
lower vanadium valence or low symmetry, which may be responsible for synergy
effect demonstrated by Ru-V sulfides mixture.

__ ---2R-UV25~" ,~ ......

....ljJIIj'~II--....
_....I011 -
a) b)

i I
3000 4000

Figure 3: ESR spectra of2RuV25 and 2RuV75 samples a) before sulfidation and b) after sulfidation

In summary, it is demonstrated in this paper that NMR and ESR of vanadium


sulfides are good tools to understand the comprehension of the synergy
phenomenom in the mixed sulfides.
536

4. References
[IJ Harris S and Chianelli R.R (1986) Catalysis by transition metal sulfides: a theoretical and
experimental study of the relation between synergic systems and the binary transition metal sulfides, J.
calal98, 17-31
[2J Scott C.E., Embaid B.P., Gonzalez -Jimenez F., Hubaut R. and Grimblot J.,(1997) Behavior of iron-
vanadium sulfide catalysts for hydrotreating reaction, J Cala/l66, 333-339
[3J Hubaut R., Betancourt P, Rives A and Scott C. (2090) Hydrotreating on mixed vanadium-nickel
sulfides: a studies of the synergetic effect, calal Today 57, 201-207
[4J Luis M.A, Rives A, Hubaut R., Embaid B.P., Gonzalez-Jimenez F., Scott C.E. (1999) HOS of
dibenzothiophene and VOOEP HOP on bulk Fe-Mo mixed sulfides, Stud SurfSc and Cala1127, 203-207
[5J Pecoraro T. A and Chianelli R. R. (1981) Hydrodesulfurization catalysis by transition metal sulfide,
J. calal.67, 430-445
[6J Shubin AA., Zhidomirov G.M., Strukt Khim Z. (1989) in Russian 30, 67
[7J Shubin, A A; Lapina, O. B.; Zbidomirov, G. M. IX AMPERE Summer School, Novosibirsk, USSR,
20-26 Sept, 1987, Abstracts, p. 103.
[8J Buishvili, L. L.; Kobakhidze, G. K..; Menabde, M. G. JETP(I983) in Russian 81 138
[9J Herzfeld, J.; Berger, A E. J. (1980) Sideband intensities in NMR spectra of samples at the magic
angle, J. Chern. Phys73,6021-6030.
[IOJ Toyomi Hayden Y., and Edwards J. O. (1986) Spectroscopic and NMR reinvestigations of hydrolysis
oftetrathiovanadate ion, Inorg.Chim.Acla, 11463-67
[IIJ Becker K.D., and Berlage U. (1983) NMR chemical shifts of vanadium 51, nobium 93 tantalum 181
and thallium 205 in ternary AlBC. semiconductors, J.Magn.Reson. 54 272-284
[12J Harrison AT. and Howarth O.W. (1986) Oxygen exchange and protonation of polyanions : a
multinuclear magnetic resonance study of tetradecavanado phosphate(9-) and decavanadate (6-),
J.Chem.Soc. Dalton. Trans. 1405-1410.
CORRELATION OF SIV NMR PARAMETERS WITH LOCAL
ENVIRONMENT OF VANADIUM SITES.

D.F. KHABIBULIN, A.A. SHUBIN, O.B. LAPINA


Boreskov Institute of Catalysis, pro Lavrentieva, 5, 630090
Novosibirsk, Russia, olga@catalysis.nsk.su

1. Introduction

' \ Vanadium oxide based catalysts are widely used in industry. Generally,
the activ(ty of heterogeneous catalysts is determined by the structure of active
centers. " FO{ characterization of active vanadium species by 51y NMR
spectroscopy the lmowledge of correlation between parameters of NMR spectra
and local structure of vanadium sites is necessary. It may be found successfully for
individual vanadium compounds with lmown structure. Earlier, the correlation
between the anisotropy of the chemical shielding tensor and the local environment
of vanadium sites was demonstrated [1, 2], but for the detailed characterization of
these sites the knowledge of other 51y NMR parameters is also necessary.
Recently, SATRAS - the numeric analysis of spinning side bands intensities in
MAS NMR spectra, was proposed by Skibsted et al.[3] This analysis allows the
determination of magnitudes and relative orientation of 51y chemical shielding
(CS) and quadrupling coupling (QC) tensors. In earlier reports all eight parameters
of Sly NMR spectra were presented with high accuracy for some orthovanadates
and alkali metal metavanadates [5], as well as for series of divalent metal
metavanadates [6] and pyrovanadates [7].
In this study SATRAS approach [8] was used in order to obtain 51y NMR
parameters for a number of vanadates and other vanadium compounds. This
investigation is devoted to a search for more accurate correlations between the
anisotropy of chemical shielding tensor and local environment of vanadium. In
addition, correlations between vanadium local environment and some other 51y
NMR spectrum parameters were also examined.

2. Experimental Section

Solid-state 5ly MAS and static NMR experiments were performed at 105.2 MHz
(9.4 T) on BRUKER MSL-400 spectrometer using the 5-15 kHz MAS probe from
NMR Rotor Consult. ApS. (Denmark) and 5 rom o.d. ShN4 rotors. 51y NMR
537
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 537-544.
© 2002 Kluwer Academic Publishers.
538

spectra were registered at room temperature using spectral width of 2 MHz, single
pulse excitation for MAS, and spin-echo sequence for static NMR measurements.
5 sec delay between subsequent accumulations was used. Isotropic chemical shifts
are reported relative to VOCh.
SIV NMR spectra parameters were determined from MAS rotational sidebands
intensities using SATRAS approach. Simulations of both SIV static and MAS
NMR spectra were performed taking into account the second-order quadrupole
correction using NMR5 program described earlier [8] . All simulations were
performed on dual PII-400 MHz CPU IBM PC compatible computer running
Linux OS.

3. Results and discussion

Generally, othovanadates contain one tetrahedrally coordinated vanadium atom in


isolated vol unit (Qo type). The structure is based on the hexagonal packing of
oxide atoms and regular distribution of cations. The local environment of
vanadium atoms in orthovanadates is very close to regular tetrahedral. Optimized
NMR parameters for orthovanadates determined in this work are summarized in
the Table 1. We obtained parameters for monovalent (M3V04' M = Li, Cs)
divalent (M3(V04)z, M = Sr, Ba ) and trivalent (MV0 4, M = AI, Sc, Y, La, Lu )
metals orthovanadates (nine compounds). In addition NMR, parameters reported
in literature are also presented in the Table 1 (seven compounds).
It can be seen that chemical shielding anisotropy (..10=10;50-0331) varies in
the limited range from 5 ppm to about 95 ppm, while the QC constants
(CQ=e 2qQIh) are in the wide range from 0.9 MHz to 5.5 MHz (Figure 1). At the
same time, the asymmetry parameters for both CS (l'J0' = 0'2~:1I) and QC
(l'JQ = 2r II , Vii -
V V
33
the eigenvalues of the electric field gradient tensor) tensors are
in the range from 0 to 1.0 (Figure 2). These values are in agreement with the
results of earlier investigations [1,9]. Thus, the regular character of local vanadium
environment determines small values of the CS tensor anisotropy, while the small
differences in V -0 distance leads to the values of CS asymmetry in the interval
from 0 to 1.0.

Table I. Parameters of Sly QC (CQand llQ) and CS (6a, 110 and aiso) tensors for orthovanadates.
Compound type Co, MHz 170 Llo; ppm 17" 0).." ppm

UNO. 1,53 0,05 12 0,80 544,3


NaNO.,ef(2) 1,05 545,0

CsNO. 3,50 0,29 30 0,18 572,9

Sr3(VO.)2 0,53 0,01 10 0,50 610,9

Ba3(VO.)2 0,76 0,01 5 0,27 602,9


539

Mg3(VO.)2 refl5] 1,05 0,56 31 0,53 557,3

Z n3(VO')2 ref ]5] 1,04 0,99 41 0,68 522,5

AIVO. Vl 3,50 0,70 90 0,30 662,0

V2 3,30 0,70 90 0,30 743,0

V3 4,50 0,70 90 0,30 776,0

LaVO. 5,21 0,69 72 0,09 612,1

LuVO. 4,28 0,01 5 0,43 663,4

YVO. 4,85 0,01 40 0,10 665,8

SeVO. 3,95 0,01 40 0,83 679,3


BiVO. ,ofl5] 4,94 0,36 94 0,32 421,1
CeVO. 'ofll0] 5,62 0,21 77 0,10 427,0
TaV0 5 ,ef I5] 0,85 0,28 53 0,24 773,2
NbV05 ,ofl5] 1,20 0,39 70 0,17 791,4

Another situation is realized in the case of pyrovanadates. Two tetrahedral


V0 4 units in pyrovanadates are linked together by sharing their comers and
fonning V20 7 unit (QI type), leading to a decrease of the number of cations
surrounding V0 4 units. The tetrahedral units in orthovanadates are isolated and
nearly regular, while in pyrovanadates one V -0 distance in tetrahedron is
significantly different from three others V-O distances. Moreover, some V 20 7
units have nearly located apical oxygen atom (originating from the neighboring
V 20 7 units) that completes coordination of one V20 7 vanadium atom to the
tetragonal pyramid. NMR parameters for pyrovanadates are presented in Table 2.
S]V NMR parameters for three compounds (Sr2V207, Ba2V207 and ZrV 20 7) were
obtained in this work, while for other pyrovanadates (M 2V 20 7, M = Mg, Ca, Zn)
S]V NMR data were taken from [5] . As mentioned above, the local surrounding of
vanadium atoms may differ from four-coordinated. For vanadium atoms in
tetrahedrons the following regions of NMR parameters are typical. The chemical
shielding anisotropy is in the range from 60 ppm to about 250 ppm, QC constant is
in the region of 1.5 - 10.0 MHz (Figure 1), and the values of asymmetry parameter
for both CS and QC tenso[s are in range from 0 to 1.0 (Figure 2). These values of
S]V NMR parameters reflect some structural features: anisotropy of CS tensor is
larger than its typical value for nearly regular tetrahedron due to the fact that one of
the v-o bonds in tetrahedral environment differs from others v-o bonds. Wide
diapason of observable QC constants may be connected with the dispersion in the
lengths of distorted V -0 distance. Therefore, the presence of a distorted V-O
distance influences the SIV NMR parameters significantly.
540

Table 2. Parameters of 51 V QC (CQ and llQ) and CS (dO, 11" and Oi'o) tensors for pyrovanadates.

Compound type Co, MHz 1]0 .:10; ppm 1],. Oiso. ppm

V1 4,82 0,43 103 0,34 603,5

3,29 0,69 57 0,91 549,2

V1 4,80 0,39 113 0,90 639,3

10,10 0,44 262 0,10 494,4

1,58 0,90 71 0,54 574,9

7,33 0,43 530 0,50 534,0

V1 560

2,50 0,85 78 0,40 582,0

585,6

591

V1 2,09 0,74 53 0,8 577

587

2,04 0,95 71 0,5 596


a-ZnN 207 ref (7) 3,86 0,56 119 0,62 616,6
CdN2 0 7 ref(7) 6,00 0,41 173 0,27 562,7

2,44 0,17 75 0,05 775,9

V1 2,57 0,32 100 0,65 581,6

3,20 0,85 99 0,49 598,8

V1 3,81 0,58 143 0,10 650,1

5,91 0,86 252 0,16 608.4


_a unable to determine
There are two types of vanadium environment in metavanadates consisting
of (V0 3)n single chains (Q2 type) or (V 20 S)n double zigzag chains, respectively.
Single chains are present in monovalent metal metavanadates and isostructural
Ba(V03 )2 and a-Sr(V0 3h, while the vanadium atoms are located in the distorted
tetrahedron due to the presence of two bridging oxygen atoms. According to [3]
for monovalent metal metavanadates and for one unequal vanadium atom in
Ba(V03)2 and a-Sr(V0 3)2 (see Table 3) the anisotropy of CS tensor varies from
240 to 400 ppm, asymmetry parameter of CSA tensor is in the range 0.6 - 0.8, QC
constant 3.0 - 6.0 MHz (Figure 1), and the asymmetry parameter of QC tensor
varies from 0 to 1.0 (Figure 2). Other vanadium atoms in Ba(V03)2 and
a-Sr(V0 3)2 have distorted in environment close to those in pyrovanadates. It can
be seen, that the anisotropy of CS tensor increases with increasing structural
distortion of V04 tetrahedron in agreement with Mastikhin et al. [1]. Therefore,
541

the distortions of the same type determinate the relatively narrow regions of the of
QC constant and CS tensor anisotropy values.
The structures of divalent metal metavanadates differ from those for
monovalent metals. There are vanadium atoms both in thrigonal bipyramidal and
octahedral coordination in double (Y20S)n zigzag chains. In the Table 1 Sly NMR
data for divalent metal metavanadates are shown [4]. Axial distortions of
octahedrons influence CS tensor. CS tensor asymmetry parameter 0[0 - OJ and
CS anisotropy parameter of 310 - 470 ppm are typical for the vanadium in
octahedron. QC constants are relatively large because EFG tensor on vanadium
atom is far from zero due to the distortions in equatorial plane of octahedron
(Figure 1). The asymmetry parameter of QC tensor is in the range 0.3 - 0.5
(Figure 2).
Penta-coordinated vanadium in some divalent metal metavanadates has
values of Sly NMR parameters presented in Table 3. The parameter of CSA
anisotropy is in the range from 200 to 400 ppm, and asymmetry 0.3 - 0.65 . QC
constant is in the region of 2.5-4 MHz, and QC tensor asymmetry is from 0 to 1.
These are the broadest regions of parameters, because the distortions of
pentahedrons are of different types and magnitudes (from highly distorted
tetrahedrons to the highly distorted octahedrons).

Table 3. Parameters of 51 V QC (CQand llQ) and CS (.1a, llo and ai'o) tensors for metavanadates.
Compound type Co, MHz '10 .do; ppm '1a O1so, ppm

LiV0 3 relI5) 3,25 0,88 228 0,72 573,2

a-NaV0 3 relI5) 3,77 0,49 249 0,67 572,8

f3-NaV0 3rei 14) 4,20 0,55 557 0,17 510,4

NHN03 relIS) 2,98 0,28 245 0,70 569,5


KV0 3ref(5) 4,15 0,85 307 0,66 547,6
RbV0 3ref 151 4,23 0,64 314 0,69 565,4
CsV0 3ref 151 4,10 0,48 314 0,67 577,4
TIV0 3ref 151 3,76 0,71 267 0,76 529,1
Mg(V03h ref 161 7,50 0,34 310 0,30 533,9
Zn(V0 3)2 relI61 6,86 0,40 333 0,02 493,8

a-Cd(V0 3h ref 161 1,70 1,00 484 0,15 521,5

f3-Cd(V0 3h re1I6) 6,46 0,47 311 0,31 468,2


Ca(V03h ref(6) 3,06 0,51 517 0,18 563,0

Ca(V03h*4H20 relI6) VI 4,18 0,92 297 0,30 580,0

V2 3,73 0,57 492 0,16 529,5

a-Sr(V0 3h re1I6) VI 4,22 0,12 218 0,32 639,1


542

V2 5,65 0,31 244 0,61 585,9


Ba(V0 3 )2,ef(6) Vl 3,68 0,16 190 0,41 658,5

V2 5,56 0,04 265 0,59 590,6


Pb(V0 3 )2,ef(6) Vl 3,75 0,13 465 0,15 529,8

V2 6,98 0,31 428 0,12 479,7

Vanadium pentoxide Y20 S is a classical example of compound with the


octahedral coordination of vanadium. Vanadium in the same coordination is
presented also in K 3Y SO I4 , Rb 3Y sO I4 , Rb 2Y 60 16, CS 2Y 40 Il , UI-YOP04 , and
!3-YOP04 . Their Sly NMR parameters are summarized in Table 4. The anisotropy
of CSA has axial character, that is a consequence of axial character of octahedron
distortion (Figure 1). The constant of QC is in the small region 0,8 - 3 MHz
(Figure 2), with diffe.rent type of asymmetry of QC tensor (Figure 3).

Table 4. Parameters of 51 y QC (C Qand 11Q) and CS (/lcr, 11" and criso) tensors for some vanadates.
Compound Type Co, MHz '70 .10; ppm '7" 0)50, ppm

Oct 0,797 0,22 635 0,09 612

VI Oct 2,45 0,44 405 0,00 548,1

V2 Oct 3,03 0,89 459 0,00 510,0

V 1 0ct 1,75 0,40 480 0,08 517,0

V2 Tetr 0,22 0,20 80 0,60 624,0

V 1 0ct 2,66 0,986 448 0,16 513,0

V2 0ct 2,34 0,30 400 0,05 547,0

(NH 4)NaOla ,ef(12) V 1 0ct 2,75 0,98 470 0,01 509,7

V2 0ct 2,33 0,34 401 0,05 546,0

V 1 0ct 1,25 0,42 930 0,03 513,0

V2 Tetr 1,68 0,69 290 0,05 575,0

Oct 1,55 0,55 820 0,00 691,0

~-VOP04 Oct 1,99 0,59 818 0,00 755,0

UII-VOP04 Oct 0,83 0,52 582 0,68 776,0

VO[SiO(OTBuhh Oct 2,45 0,10 445 0,05 776,0

The correlation between L\cr and 110" for compounds from Tables 1-4 is
presented in Fig. 1. It is possible to distinguish almost unambiguously among the
regions characteristic for the different types of vanadium surrounding. These
regions, marked by smooth lines in Fig. 1, provide a way to draw conclusions
about the local surrounding of vanadium in compounds with defined S)y NMR
parameters but unknown structure. Evident correlation is also observed between
543

CS anisotropy and QC constant (Fig. 2). As the previous one, this correlation
should be useful for the investigation of vanadium environment. At the same time
no evident correlation, suitable for structural conclusions, was found between .10"
and 'lQ (Fig. 3). .
The comparison of SIV NMR parameters of ortho-, meta- and
pyrovanadates (for Mg, Zn, Sr and Ba) is very interesting because QC constant
increases along with the CS anisotropy in the row vol, v2o/", V0 3•• The
monotonic dependence is visible in Fig. 4 where the correlation between .10" and
CQ presented for low-temperature modifications of M 3(V04)2, M2 V 207 and
M(V0 3)2 (M = Mg, Zn, Sr and Ba). This observation is quite useful and may be
directly used in the investigations of similar systems.

1,0 11 .---------------------~
O,g c • O· 10
cO' 9
O.B
0.7
-a' 8

.
· VO~ 7
fl.6
· vo. Ca, 6
I}.~ MHz
0.5
5 •
0.4
0.3 3
0.2 2

a 200 400 600 BOO 1000


0=------------'
o 200 400 600 OOfl 1000
.la . ppm ,w, ppm
Figure I. The cOlTelation between " Y MR Figu re 2, The correlation between Sl y MR
a ymmetry and anisotropy parameters of S for anisotropy of CS and QC constant for compounds
compounds presented in Tables 1-4. presented in Tables 1-4.

-
1,0 8
to
• •
•t.!J. v
7
0,8 6 <> .

.. ..•.
0,6
."
" Co. 4
5 to
<>
<> rI' ~ MHz to • AMg
~c ~ <> Q O 3
<>
0.4 0-:'
<> •
V v • Q'
~
• · 211
• 2
• ... rv - Sr

..
00 A Q'

0,2 1
<> • va,
1
oI
'" • Ba
I
II va, a 100 200 300 400
P
a ~" . ppm
0 200 400 600 800 1000
,\ r; . ppm
Figure J. The correlation between Sl y MR Figure 4. The correlation between Sl V M R an isotropy
an isotropy of CS and asymmetry of QC for compounds of CS and Q constant for compounds M,(YO.h,
pre cnted in Tab les 1-4 M1Y10 , and M(YOl h (M = Mg, Zn, Sr and Ba)

4. Conclusions

SI V NMR parameters are obtained for Li, Cs, Sr, Ba, AI, La, Lu and Y
orthovanadates, Sr, Ba and Zr pyrovanadates, as well as for some more complex
vanadium compounds (K3VS0 14, Rb 2V60 16, CSN4011 and VOP0 4). The relation
between SIV NMR parameters and vanadium local environment is examined using
544

both new and literature Sly NMR data. Correlation between CS anisotropy and QC
constant is presented, and previously known correlation between anisotropy and
asymmetry parameters of CStensor is refined. Definite intervals of parameters
characteristic for different types of vanadium environment may be found in these
correlations. The correlations clearly indicate that for the characterization of
vanadium sites with unknown structure the knowledge of both CS and QC tensors
parameters is required. That the most informative are: CS anisotropy, QC constant
and CS asymmetry parameter.
For one metal, both CS anisotropy and QC constant are growing in the row
M3(Y04h, M2Y20 7 " M(Y0 3)2' At the same time, quadrupole asymmetry
parameter is not characteristic.
The correlations obtained may be used for Sly NMR studies of vanadium
ions in solids, including supported vanadium catalysts.

Acknowledgment
This work was supported by the RFBR (grant no. 01-03-32364), and INTAS
(lR-97 -0059).

5. References
J. Mastikhin V.M., Lapina O.B., Krasilnikov V.N., Ivakin A.A.(1984) 51V NMR Spectra of Vanadates and
Oxosulfato-vanadates of Alkali Metals, React.Kinet.CataI.Lett., 24, 119-125.
2. Lapina O.B, Mastikhin V.M., Shubin A.A., Krasilnikov Y.N., Zamaraev K.I., (1992) 51y Solid State NMR
Studies ofYanadia Based Catalysts, Progress in NMR spectroscopy, 24, .457-525.
3. Skibsted, 1., Nielsen, N.C., Bildsee H., Jakobsen, H.J. (1992) 51y MAS NMR spectroscopy: determination of
quadrupole and anisotropic shielding tensors, including the relative orientation of their principal-axis
systems, Chem.Phys.Lett. 188,405-412.
4. Skibsted, J., Nielsen, N.C., Bildsee H., Jakobsen, H.J. (1993) Magnitudes and Relative Orientation 0 51y
Duadrupole Coupling and Anisotropic Shielding Tensors in Metavanadates and KYlOS from 51y MAS NMR
Spectra., J.Am.Chem.Soc. 115,7351-7362.
5. Skibsted 1., Jakobsen C.J.H., and Jakobsen H.J., (1998), Sly Chemical Shielding and Quadrupole Coupling in
Ortho- and Metavanadates from 51y MAS NMR Spectroscopy, Inorganic Chemistry 37, 3083-3092.
6. Nielsen V.G., Jakobsen H.J., and Skibsted 1., (2000), Characterization of Divalent Metal Metavanadates by
Sly Magic-Angle Spinning NMR Spectroscopy of the Central and Satellite Transition, Inorganic Chemistry
39,2135-2145.
7. Nielsen V.G., Jakobsen H.J., and Skibsted J., (2001), 51y MAS NMR Investigation of SIV Quadrupole
Coupling and Chemical Shift Anisotropy in Divalent Metal Pyrovanadates, The Journal of Physical
Chemistry 105,420-429.
8 AAShubin, O.B.Lapina, D.Courcot (contributors: AAShubin, O.B.Lapina, D.Courcot, A.Aboukais,
B.Revel, M.Rigole, M.GueJton, S.Caldarelli, J.C.Yedrine), (2000), Characterization by solid state Sly NMR
spectroscopy, Catalysis Today 56, 379-387.
9. Hayakawa S., Yoko T., and Sakka S., (1994), 51V NMR Studies of Crystalline Monovalent and Divalent
Metal Metavanadates, Journal of Solid State Chemistry 112, 329-339.
10. Cousin R., Abi-Aad E., Courcot D., Capelle S., Arssi F.C., Amoureux J.P., (1999), 51V MAS NMR
Characterization of Y -Ce-O catalysts, Colloids and Surface A 158, (1-2), 43
II. Shubin A.A., Lapina O.B., Bosch E., Spengler 1., and Knozinger, (I999), Effect of Milling of V20 5 on the
Local Environment of Yanadium as Studied by Solid-State Sly NMR and Complementary Methods, The
Journal of Physical Chemistry B 103 (6), 3138-3144.
12. Shubin A.A., Lapina O.B., Sly SATRAS NMR structure investigation of solid yrri catalysts. In "Structure
and dynamics of molecular systems", Ioshkhar-Ola-Kazan-Moscow, 1998, p. 36-41 (In Russian).
SOLID-STATE NMR STUDIES OF MESOSTRUCTURED ALUMINO-
PHOSPHATES: STRUCTURE AND DYNAMICS OF THE INORGANIC
NETWORK AND OF THE ORGANIC COMPONENT

YAROSLA V Z. KHIMYAK AND JACEK KLINOWSKI


Department of Chemistry, University of Cambridge
Lensfield Road, Cambridge CB2 JEW, United Kingdom

Abstract

An array of solid-state NMR techniques has been used to characterise alumino-


phosphates with hexagonal and lamellar mesostructures. Differences between the
mesocomposites regarding the inorganic component, state of the template and
arrangement of the inorganic/organic interface have been identified. We describe the
complex picture of the structure of the hybrid solids.

1. Introduction

The similarities between silicates and alurninophosphates extend to mesoporous and


mesostructured solids formed by non-covalent interactions between the inorganic
component and organised organic arrays of amphiphific molecules [1-4].
Mesostructured aluminophosphates (AlPO's) were obtained using neutral [5-7], anionic
[8] or cationic [9, 10] templating. We have described the synthesis of mesostructured
AlPO's in ambient [11] and hydrothermal [12, 13] conditions using cationic templating.
This leads to materials with lamellar (Ll, L2 and L3) or hexagonal (Hex-I, Hex-2 and
Hex-3) mesostructures. Calcination of Hex-I and Hex-2 leads to a mesoporous AlPO
capable of incorporating different heteroatoms such as Mg, Co, Si, Ti, Band Zn [14].
Since the inorganic component of the mesostructured AlPO is known to be largely
amorphous, applicability of XRD for accurate structure determination is limited. On the
other hand, solid-state NMR is ideally suited to study materials with low-ordered
structures. We used several different solid-state NMR techniques to examine
connectivities in the inorganic component [11, IS], the specific organisation of the
organic template [16], and the interaction between the inorganic component and the
organic template. In this work we present a complex picture of the structure of AlPO
mesocomposites. The diversity of mesostructured AIPO's offers a unique opportunity
to compare different products, with both the same and different mesostructures, which
will provide a better insight into the supramolecular assembly of the AIPO system in the
presence of the cationic surfactant.
545
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 545-551.
© 2002 Kluwer Academic Publishers.
546

2. Experimental

Hexagonal and meso lamellar aluminophosphates were synthesized from an aqueous


solution containing Al(OC 3H/soh, H 3P04 (85 wt%), hexadecyltrimethylammonium
chloride (C ,6TMACI) and tetramethylammonium hydroxide (TMAOH). Details of the
syntheses have been given elsewhere [11-13].
The inorganic component was characterised using HAl and 31p MAS NMR with the
acquisition parameters given in [11, 13, 15]. The experimental details of the variable
contact time (VCT) IH_3I p CP experiments and data analysis (determination of the
cross-polarization time TpH, and the IH and 31p spin-lattice relaxation times in the
rotating frame) were presented in [15] . The organic template was studied using
variable-temperature I3C MAS NMR, IH - I3C CPIMAS dynamics, T lpc measurements,
IH MAS NMR and T z measurements (acquisition parameters given in [16]). TCH and
T lpH were determined from VCT IH - I3c CPIMAS spectra using the equation for an
abundant-to-rare CP [17]. T 1pc was measured in a separate experiment [18]. Details of
the Wide-line ~aration 2D NMR (2D WISE) experiments [19] have been given
elsewhere [16].

3. Results and Discussion

3.1 . STRUCTURE OF THE INORGANIC COMPONENT

Mesostructured AIPO' s vary in the degree of condensation of the inorganic network and
in the population of Al sites with different coordination (Table 1). The inorganic
component of Hex-l consists of alternating Al(6) (OAI = -5.7 ppm) and tetrahedral P
units. Even though some AI(4) units are present (OAI = 42.4--43.5 ppm), the low-field
31p chemical shifts suggest the absence of extended regions of alternating phosphate
and aluminate tetrahedra. The surfactant head groups at the inorganic-organic interface
form electrostatic bonds with the phosphate tetrahedra, which are probably connected to
two Al(H zO)3(OP)3 or Al(HzOMOP)z units. Further from the interface, the number of
water molecules in the coordination sphere of AI(6) decreases. AI(4) tetrahedra are
accommodated in the bulk of the inorganic component, leading to a small contribution
of the units O-P(OAI03MOAI(H20)nOs.n)3.k (n = 2, 3; k = 1, 2), where P atoms are
connected to both AI(4) and AI(6), with 31p chemical shifts of between -16.9 and -
19.9 ppm.
The inorganic component in Hex-n AlPO's is amorphous [11]. The differences
between Hex-I, Hex-2 and Hex-3 AIPO's are related to the increased number of H 20
molecules the coordination sphere of the AI( 6) units, as reflected in the downfield shift
of the AI(6) peak to -2.9 ppm for Hex-2 and -1.8 ppm for Hex-3. At the same time, the
proportion of the sites where the phosphate tetrahedra are connected to only two or one
Al atoms (op > -4.0 ppm) increases.
Unlike Hex, the AlPO layers in Ll and L2 are much more regular, as shown by the
much narrower lines in 27Al and 31p NMR spectra. 27Al and 31p NMR spectra ofLl are
consistent with the presence of P(OAl03)OR, R = H, CI6H33N(CH3h + units. Phosphate
547

tetrahedra are dominant at the inorganic-organic interface. The structure of the


inorganic component in L2 is more complex. The dominant 31p NMR peak at -
31.2 ppm indicates the presence of extensive regions with alternating phosphate and
aluminate tetrahedra, while the structure of the interface region is similar to that of the
Ll layers. Thus, the inorganic layers in L2 are thicker than in Ll. In contrast, in L3 the
coordination of Al is mostly octahedral and the inorganic network is much less
condensed, with the main 31 P resonances between -0.1 and -13.0 ppm.

TABLE I. Properties of the inorganic component of the mesostructured products

Parameter Mesoiamellar AIPO Hexagonal AIPO


Ll, L2: AI(4)>> AI(6)
AI(4) : Al(6) AI(6) > AI(4)
L3: AI(6) > AI(4)
Ll, L2: from-17.0 to-31.0ppm
31p MAS NMR chemical shift from - 1.0 to - 19.0 ppm
L3: from -1.0 to -19.0 ppm
Degree of condensation L2>Ll>L3 Hex-I> Hex-2 > Hex-3
LI, L2: P: AI > 1.50
P: AI ratio P : AI = 0.70-1.30
L3: P : AI '" 1.30
Content of organic component L2<Ll<L3 Hex-I < Hex-2 < Hex-3
Thermal stability L2 > Ll > L3 Hex-I> Hex-2 > Hex-3

'000 - -..
, ..o,egppm
2000 I ------1
\ ·183 ppm
2 ·8.0 ppm 2 ·213_
800 :) .135ppm 3 ·233_

.,
1500 .. -31 . ppm
, ,
~
~
~ :/
200 f
iE
1000 !

500

0 - • I
0 -
0 '0 '5 20 25 0 15 20 25
'0
I ...../ms

Figure I. 'H_Jlp CP dynamics of Hex-I Figure 2. 'H_ 3I p CP dynamics of L2

A better insight into the structure of the inorganic component was obtained using 'H_
31 p CP/MAS dynamics, which can not only justify the assignment of different lines to
the specific structural units, but also provide important information on the structure of
the inorganic-organic interface. Hexagonal materials have much faster IH _ 31 p
CPIMAS kinetics than lamellar products (Figures 1, 2). In Hex the TPH is quite short
(0.90-2.72 ms). Polarization transfer may thus be thought of as occurring within the P-
OH bonds. This is justified by the substantial contribution of P sites with a low degree
of condensation. Moreover, the inorganic network consists predominantly of highly
hydrated Al( 6) sites. An increase in T PH for sites resonating at higher field is related to
the increased distance between the P atoms and the source of protons. IH - 31p CP
dynamics also identify P atoms connected to the surfactant headgroups (-0.9 ppm for
548

Hex-l and -1.1 ppm for Hex-2}. For these units, in addition to the protons from the
hydroxyl groups or water molecules coordinated to Al atoms, CP may also proceed
from mobile protons of the surfactant headgroups. For L3, the TpH for sites with a low
degree of condensation is comparable to those for Hex, while the T PH for the P sites
resonating at a higher field is similar to those for Ll and L2. The proton source in Ll
and L2 is much more distant from the P atoms than in Hex, as Ll and L2 have
substantially longer T PH. This is consistent with a higher degree of condensation of the
inorganic network in L, and hence a lower content of P-OH groups. The environment
of Al in L is also different, as AI(4} is dominant. Thus CP proceeds mainly from
surfactant species and remote water molecules.
Measurements of T lpH indicate the micro domain structure of Hex-I, Hex-2 and L3.
Two components with different proton relaxation rates have been identified. The fast
relaxing component is attributed to protons within the inorganic network (mainly water
molecules and OH groups coordinated to octahedral Al sites), while the component with
slower relaxation rate represents IH atoms at the inorganic-organic interface and from
the organic component.

3.2. DYNAMICS OF THE TEMPLATE

Although mesostructured AlPO's contain C I6TMA+ cations as the only organic species,
\3C MAS, IH - \3C CP/MAS, 2D WISE and IH MAS spectra show differences in the
chemical status of the occluded template. \3C MAS NMR reveals different
conformations of the alkyl chains. As seen from the intensity ratio of the peaks at
ca. 31.0 and 33.0 ppm, room temperature conformation of the aliphatic chains in Hex-I,
Hex-2 and L2 is predominantly disordered gauche, and in Ll and L3 ordered all-trans.
Variable-tempera~re 13C MAS NMR indicates conformational changes of the aliphatic
chains, with the all-trans <=> gauche transition occurring at different temperatures for
different mesocomposites. Hex- l and Hex-2 have the lowest transition temperature of
-30°C, while in L2, L3 and Ll, the transition takes place at 10, 40 and 70°C,
respectively (Figure 3).

1.00 1.00 I
& & &
0
&
I r •
& Ll
••

& &
0.80 , 0 0.80 L2
& 0 00 !

0
0
i o L3
0.60 ~

r
~ 0.60 "
I
~
-a
+ • o

"§, OAO ~ ~ 0.40 .


0 &
! • •
• oO ilX~
I
r i
&
o Hex-l
0.20
i
0

o
& Hex-2 I 0.20 ~& ~ 0 &

0.00
I
1._,_l......._......_.L ............... ~

-100 -80 -60 -40 -20


. ~.~. ...... .....o.....o... .~~~.J
0 20
:
40 60
0.00 I ~ _~,_~_~~
-100 -80 -60 -40 -20 0 20 40 60 80
rfc rfc
Figure 3. Relalive conformer composition of the aliphatic chains vs. temperature.
549

Generally, the transItlon temperature corresponds to the transformation from the


conformationally ordered into the disordered state. A higher transition temperature
indicates relatively lower mobility of the aliphatic chain, which decreases in the
sequence Hex-I, Hex-2 > L2 > L3 > Ll.
IH - I3C CP/MAS NMR and T 1pc relaxation show that mobilities of different segments
of the occluded surfactant depend on the type of mesoscopic organisation and on the
conformation of the aliphatic chain [16]. The least mobile segment is located at the
inorganic-organic interface, and corresponds to the N-methylene and the adjacent
methylene group. The tail of the aliphatic chain is the most mobile. CP dynamics
indicate that the mobility of these two segments in the structure is determined by the
conformation of the alkyl chain, and is lower in the mesocomposites with
predominantly trans conformation (Ll and L3). On the other hand, CP dynamics of the
inner-chain carbons serve as a probe for the mesoscopic organisation. These carbons
have significantly lower T 1pc, both in the dominant gauche and the minor trans
conformations, in hexagonal AIPO than in L2 with the same conformation of the alkyl
chain, and in Ll and L3 with all-trans conformation. From IH - \3C CP dynamics and
T IpC measurements we conclude that the mobility of the aliphatic chain decreases in the
sequence Hex-l ~ Hex-2 > L2 > Ll ~ L3, as was found from the variable-temperature
spectra.

100
'/I

~) 60 .1.4) ~o
f'PllfnxnlM!t

Figure 4. 2D WISE spectrum of Hex-I . Figure 5. 2D WISE spectrum of L3

Analysis of 2D WISE spectra confirms the much lower mobility of the surfactant
species in Ll and L3 than in the gauche-dominant mesophases, as the IH wide lines in
Ll and L3 are much broader for all corresponding \3C sites. The FWHM of IH lines for
the gauche-dominant mesophases increases in the series L2 < Hex-l < Hex-2. A
different mobility trend, with L2 showing the highest mobility, can thus be derived.
This trend is also confirmed by the IH spin-spin relaxation measurements [16].
The slight disagreement between results from I3C relaxation studies on the one hand and
IH MAS spectra and WISE on the other, indicate the need to consider a more complex
model of surfactant behaviour in the AIPO mesophases. We distinguish local mobility
of the chains from the bulk or cooperative mobility. The local mobility is reflected in
the T2 relaxation times of IH and the IH linewidths in the 2D WISE spectra.
Cooperative mobility is responsible for the trans <::> gauche transition, and governs the
550
IH - I3C CP dynamics and the Tip relaxation of I3C. The local mobility includes
movements of the segments of the individual aliphatic chains and is monitored over
relatively short distances. On the other hand, cooperative motions of the aliphatic
chains are characteristic of the particular arrangement of the surfactant within the
inorganic host and depend on the mesophase.

4. Summary and Conclusions

Different solid-state NMR techniques have enabled us to understand the structure of


complex mesostructured AIPO, which can be presented schematically (Scheme 1). The
organization of mesostructured AIPO is based on ordered arrays of the template
surrounded by the inorganic component with different degrees of condensation.
L 1 and L3 have organic arrays
Htx·1
consisting of tightly packed surfactant
bilayers with all-trans ordered
~ conformation, around which the

~ inorganic component with a low degree


of condensation is arranged. Although
hexagonal AIPO show a very low
u
degree of condensation of the inorganic
l1111lIlIlI network, the template is in the
UIUUUU
disordered gauche conformation
111I1I1I1I1 imposed by the tubular arrangement of
UIUUUU
the inorganic component.
l2
In L2 the template is in gauche
conformation despite a well-defined

Iffli
meso lamellar structure. This results
from the much larger distance between
the surfactant head-groups imposed by
the specific structure of the
inorganic/organic interface. The
Scheme I. The structure of mesostructured
AIPO . Left: mesoscopic organisation; right:
structure of L2 is based on the highly
organisation of the organic component. condensed inorganic component,
enabling the surfactant to be more
mobile than in L1 and L3.
On the other hand, the lamellar structure leads to the lower bulk mobility of the template
in L2 compared with he':(.agonal AIPO's.
The inorganic/organic interface consists of alternating P and Al units. The surfactant
head-groups are electrostatically bound to the phosphate tetrahedra and are most mobile
in hexagonal AIPO.

5. Acknowledgement

We are grateful to the Cambridge Oppenheimer Fund for the Research Fellowship for
Y. K.
551

6. References

1. Beck, J.S ., et al. (1992) A New family of mesoporous molecular sieves prepared with liquid crystal
templates,J. Am. Chem. Soc. 114,10834-10843.
2. Tanev, P.T. and Pinnavaia, T.1. (1995) A neutral templating route to mesoporous molecular sieves,
Science 267, 865-867.
3. Kresge, C.T., et al. (1992) Ordered mesoporous molecular sieves synthesized by a liquid-crystal
template mechanism, Nature 359, 710-712.
4. Inagaki, S., Fukushima, Y. and Kuroda, K. (1993) Synthesis of highly ordered mesoporous materials
from a layered polysilicate, J. Chem. Soc., Chem. Commun., 680-682.
5. Oliver, S., et al. (1995) Lamellar aluminophosphates with surface patterns that mimic diatom and
radiolarian microskeletons, Nature 378, 47-50.
6. Sayari, A., el al. (1996) Synthesis of meso structured lamellar aluminophosphates using supramolecular
templates, Chem. Mater. 8, 2080-2088.
7. Gao, Q.M., et al. (1997) Synthesis and characterization of a family of amine-intercatalated lamellar
aluminophosphates from alcoholic system, Chem. Mater. 9,457-462.
8. Froba, M. and Tiemann, M. (1998) A new role of the surfactant in the synthesis of mesostructured
phases: Dodecyl phosphate as template and reactant for aluminophosphates, Chem. Mater. 10, 3475-
3483.
9. Kimura, T., Sugahara, Y. and Kuroda, K. (1999) Synthesis and characterization of lamellar and
hexagonal mesostructured aluminophosphates using alkyltrimethylammonium cations as structure-
directing agents, Chem. Mater. 11,508-518.
10. Luan, l., et al. (1998) Characterization of aluminophosphate-based tubular mesoporous molecular
sieves,J. Phys. Chem. B 102, 1250-1259.
II. Khimyak, y.z. and Klinowski, 1. (2000) Synthesis of mesostructured aluminophosphates using
cationic templating, Phys. Chem. Chem. Phys. 2,5275-5285.
12. Khimyak, Y.l. and Klinowski, 1. (1998) Synthesis and characterisation of two novel mesolamellar
aluminophosphates, Chem. Mater. 10,2258-2265.
13. Khimyak, Y.l. and Klinowski, 1. (1998) Synthesis of new mesostructured aluminophosphates, J.
Chem. Soc., Faraday Trans. 94,2241-2247.
14. Khimyak, Y.l. and Klinowski, 1. (2001) Incorporation of magnesium in mesostructured and
mesoporous aluminophosphates, Phys. Chem. Chem. Phys. 3, 1544-1551 .
15. Khimyak, Y .l. and Klinowski, 1. (2001) IH_ 3I p CP/MAS NMR studies of mesostructured
aluminophosphates, Phys. Chem. Chem. Phys. 3, 2544-2551.
16. Khimyak, Y.l. and Klinowski, 1. (2001) Solid-state NMR studies of the organic template in the
mesostructured aluminophosphates, Phys. Chem. Chem. Phys. 3,616-626.
17. Voelkel, R. (1988) High-resolution solid-state \3C NMR spectroscopy of polymers, Angew. Chem. Int.
Ed. 27, 1468-1483.
18. Schaefer, J., Stejskal, E.O. and Buchdahl, R. (1977) Magic-angle \3C NMR analysis of motion in solid
glassy polymers, Macromolecules 10, 384-405.
19. Clauss, 1., et al. (1992) Stiff macromolecules with aliphatic side chains: Side chain mobility,
conformation, and organization from 2D solid-state NMR spectroscopy, Macromolecules 25, 5208-
5214.
DISPERSION IN POROUS BEAD PACKS STUDIED BY VELOCITY
EXCHANGE SPECTROSCOPY

ALEXANDRE A. KHRAPITCHEV', SIEGFRIED ST APF# and


PAUL T. CALLAGHAN'
, Institute of Fundamental Science. Massey University.
Palmers ton North. New Zealand
# Lehrstuhl for Makromolekulare Chemie. ITMC, R WTH Aachen
Worringerweg 1. D-52074 Aachen. Germany

The use of Pulsed Gradient Spin Echo (PGSE) NMR to investigate flow in porous
media is well established [1,2] . In particular, the Fourier Transformation of the echo
attenuation with respect to the encoding wave vector yields the probability density of
displacements, or the propagator. The shape of the propagator depends on the encoding
time L'1, during which dispersion processes allow a sampling of different portions of the
flow field by a given fluid particle. Using two pairs of gradient pulses in the PGSE
experiment, separated by a mixing time 'em, allows the examination of velocity
fluctuations by comparing displacements during the two encoding intervals. In
particular, Velocity Exchange Spectroscopy (VEXSY) [3-5] experiments can reveal the
gradual transition to asymptotic behavior as the ratio of the exchange time to the
correlation time is increased. Furthermore, the VEXSY results can be employed to
calculate the conditional probabilities of velocities, a representation which is suitable to
visualize the loss of correlations with increasing 'em.

In this work, a detailed investigation of displacement propagators for flow in a random


packing of monodisperse beads is presented. The Peelet number scaling effects are
investigated by varying the fluid, flow rate and bead diameter. Measurements
employing the VEXSY technique are presented and compared with results of recent
numerical simulations.

l. Introduction

Brownian motion is a stochastic process in which molecular migration is a random walk


and the asymptotic (i.e., at times much greater than the molecular collision time)
ensemble-averaged moments of the displacement distribution are <Z>=O and
<Z2>=2Dt, where t is the evolution time. For a stationary fluid, self-diffusion provides
the only mechanism for molecular migration. In the presence of flow this stochastic
migration of molecules across the flow direction is known as Taylor dispersion. In
porous media, further mechanisms for dispersion are observed such as streamline
spreading, which introduces stochastic variations in both the velocity and the
553
J. Fraissard and O. Lapina (ells.), Magnetic Resonance in Colloid and Interface Science. 553-558.
© 2002 Kluwer Academic Publishers.
554

corresponding molecular displacements, an effect known as mechanical dispersion. In


addition, changes of velocities are also brought about by holdup of fluids in the dead
ends of pores and slowing down of particles in boundary layers.

A crucial parameter in defining the temporal structure of the velocity field is the
correlation time, 'tc , corresponding to the duration of flow across the characteristic
length scale. For a medium with pore size or pore spacing given by dimension d, the
correlation time may be written

d
r,. =<v>
-- (I)

where is <V> is the average velocity of the fluid. In this work we shall describe flow in
a packed bed of spherical beads, for which we shall take d to be given by the bead
diameter, d p

The second important quantity describing the relative weight of convective and
diffusion contributions to the fluid motion is the Peclet number (Pe). This quantity
expresses the ratio of the time taken to diffuse across a pore to the time taken to flow
across a pore, and is given by

(2)

where Do is the molecular self-diffusion coefficient. In the case of bead packs it is usual
practice to take the characteristic dimension I as being given by the effective pore
diameter, defined by l=</ldpf(l-<I», where <I> is factor of porosity.

2. Pulsed Gradient Methods

PGSE NMR [1,2] is used to measure translational displacements of spin-bearing


molecules over the duration between two gradient pulses. Each of these pulses produces
a local phase shift in the transverse magnetization, which depends on the nuclear spin
position. Because the two pulses are arranged to produce opposite shifts, the cumulative
effect of the spin pair is to produce a phase difference proportional to the change in
position, i.e. displacement.

a)
acq
c=:::=

Figure I u. Schematic rf and gradent pulse sequence for a single PGSE NMR experiment in which the
gradient pulse area (GIi) is stepped.
Figure I b. rf and gradient pulse sequence for a general Double PGSE NMR experiment in which the
gradient pulse areas (GIi) of both pairs are stepped indq>endently.
555
The basic PGSE pulse sequence is shown in figure la. A gradient pulse of duration 0
and amplitude G causes a local dephasing for a spin isochromat at position r by an
angle $=,¥oG.r, leading to a phase factor exp(i$). We will assume that the magnetic field
gradient is applied along some particular axis of the sample (say the z-axis) and that the
amplitudes of the two gradient pulses are given by G 1 and G2 . Assuming the narrow
gradient pulse approximation (O«L\), and labelling the position of a particular spin at
the time of the two gradient pulses as Zl and Z2 respectively, the net result of the spin
echo sequence (neglecting relaxation effects) is to induce a phase shift L\<J>=yO[-G1Z 1+
G2Z 2 ]. The subtraction of phases is essential to the echo formation and leads to
cancellation of precession differences due to chemical shift effects and field
inhomogeneities. In the standard PGSE experiment G1 and G2 are set to be identical in
value (G), so that there will also be a cancellation of phase shifts due to the mean spin
position. Only the change in position, Z= - ZI + Z2, over the duration L\ leads to a phase
term and this may be expressed by exp(iyOGZ).

It is conventional to regard the effect of the two pulses as defining a scattering


wave vector [1], q= (21t)" lyGO. The normalized echo attenuation is related to the spin
displacements ZI (parallel to the gradient direction) which take place over the duration
L\ between the gradient pulses as

E(q) =< exp(i2nqZ) > (3)

The ensemble average < .. > is precisely the integral of the phase factor exp(i2nqZ)
weighted by the probability density for displacement over the time L\, namely the
- -
average propagator P'(Z,L\). Thus E is the Fourier Transform of P'(Z,L\) and this
propagator can in principle be reconstructed by measuring E over a range of q values
and performing an inverse FT.

In Double PGSE [4,6], a second pair of PGSE gradient pulses is applied at a later time
(along the same encoding axis) to the same transverse magnetization. The result is to
add a second phase difference arising from the later displacement. The corresponding q
and Z values associated with these differences are ql and q2 and ZI and Z2. The basic
Double PGSE pulse sequence is shown in figure 1b. The resultant echo has the
normalized amplitude

(4)

In the case that the two pulse pairs have their q values varied independently, data is
acquired in a two-dimensional manner. This method is known as Velocity Exchange
Spectroscopy (VEXSY)[3] and generates a two-dimensional data set E(qJ,q2) as given
by equation 4. Double Fourier transformation with respect to ql and q2 returns the two
dimensional Fourier spectrum P,(Z"L\) P(Z),'C mIZ2,L\) where P(ZI,'Cm IZ2,L\) gives the
conditional probability that a molecule which moved by ZI over the time delay L\, will
556
displace by Z2 over ~ when this latter ,.
water 2. 1 10·< 01 'ls)
measurement is made after the delay t m . ( D~

VEXSY may be used to examine how the


V .. 2 . ·~ mmls
be.lds Oll ll<
distribution of velocities during flow changes t, 130 01s
Pc 3S(J
over a well-defined timescale. In that sense
a)
tm plays the part of an exchange or mixing
time, in the manner of a classical two-
1:J t, = 0.0 1!
dimensional NMR experiment [7]. ;>.,
.'::::
'"s::
3. Experiments and results ""'
~y -a
I!.l

V .q
All experiments were performed on an AMX E
-t:, .D
300 MHz Bruker spectrometer. As an I!.l .D
«:
--1
-.'-- -.. --i
0
example for porous media, polystyrene beads C/l
c<:: ....
0-
<...) 0.0 mOl OA
were used with variety of sizes from 100 Jlm U
.;; .. . ... ,
to 500 Jlm. Bead packs were inserted inside a
.....
0
....;>.,
0,4
't~ /-co.:.. 0.0 I'
cylindrical sample with a diameter of 2 mm c.
for the smallest beads (100 Jlm) and JO mm S
;:.., .D
~

for all others. Maximum available gradients '"


c<::
I
.D
....
0 e
E
for these systems were 80 O/mm and I!.l
..... 0-
16 O/mm respectively and the duration of the 0....
'"s::
gradient pulses was set 250 JlS. The choice of ....0
~ depends on the investigated flow rate and -a
s:: 0 .0 mm 0.4
the maximum available gradient, but in all 0
u
-
cases the relation 8« ~ < T, was kept.
Two liquids with different diffusion '
fI.4 ! 1:,../ 1,~
- 6.5 1
coefficients were measured, pure water
0
(Do = 2.1 10-9 m2/s) and 70% solution of
glycerol (Do = 2.3 10- 10 m 2 /s) . The liquids '"s::
<1J
-a
were pumped through the systems with flow ""' C
~y
rates ranging from 100 mllh to 10 Ilh which
1\ .D
was equivalent to average velocities from E ~
I:'" .D
2 mmls to 50 mmls and corresponded to '-" 0
....
correlation time in the interval between I!.l c. 0.0 mm OA
10 ms and 150 ms. For each set of system '"c<::
<...)
()
parameters the experiments were performed .;; 0 0.4 '
' t IC I t (: 6.5 i
with different mixing times. .....
0
:.0
c.
In figures 2 bod exemplary experimental
S
;>.,
.D
0
'" E
.... E:
«
C/l
c.
results for three different Peclet numbers are
presented; for comparison, one set of '"0s::
computer simulations is shown in figure 2a. :-a
.e (J.o r
,
Figure 2 o. Numerical simulations [8] of the VEXSY s::
0 0.0
- .-mm- _..... _- IJA
experiments with parameters shown in top of chart. U
Figure 2 bod. Experimental data for the system and .I

system parameters (shown on top of each chart).


557

water (D, ~ 2.1 10' m'/s) W'dtcr(D. - 2.1 10· m'fs) 70% glyc.:roJ (D. = 2.3 10"" m'fs)
v . =2.1 mmls Vm = IS mmls V.~ 50 ",mi.
G

beads 100 I!S beads ~ 400 I!S beads 400 ~IS


t. ~ 50ms t , = 30 ms t,= IOms
Pe ~ 100 Po - 3000 Pe 100000
b) c) t/)

0. 1 t., 1 't, = 0.Q3 0.4

E
E

0.0

mm OA

0. 1 't",/ 't, = 0.02 0.4 'tm 1 't, = 0.03 0.4

E
E ~

o.o HIWI/IIf1'.?-""'=--------1 0.0

0.0 mm 0. 1 0.0 mm 004

't",/ 't, = 6.5 0.4

0.0 0.0

mm 0.1 0.0 mm 0.4 0.0 mm U.4

0. 1 't m / 1, = 5.0 0.4 't",1 't, = 6.5 0.4 1 m I 1, = 5.0

8 E
E s !J

0+' lc:!,!!!!!I!I~Ii~ ")


0.0

I_ U,o mm 0. 1 0.0 mm 0.4 (J.U mm 0.4


558
Results for joint probability densities as well as conditional probabilities are presented
for mixing times much shorter and much longer than the correlation time (see eq. 1),
respectively. For very short times a strong correlation between initial and final
velocities can be observed. In the VEXSY plots this can be seen from an alignment of
the joint probability density along the main diagonal of the coordinate system.
However, the conditional probabilities show these dependencies in a much clearer way,
as well as providing additional information from the exact shape of the plot of these
two-dimensional functions . One particular detail is the "S"-shape of P(ZI;tmIZ2'~)'
which is observed in both experimental and simulated data (see, e.g., the second row
from the top of fig. 2).

For mixing times 'tm well exceeding the correlation times the VEXSY plots tum into
triangular shapes. P(ZI,'tmIZ2'~)' on the other hand, consists mainly of lines parallel to
the abscissa axis, this being equivalent to conditional probabilities, which become
independent of the displacement during the first encoding interval.

4. Conclusions

The evolution of the flow profile in random packings of spherical particles has been
investigated by two-dimensional Velocity Exchange Spectroscopy over a wide range of
Peelet numbers by varying the bead size, fluid viscosity and flow rate. The direct plot
of the joint probability density was supplemented by representations of the conditional
probability density, which provides easier visualisation of the loss of the correlation'
between initial and final displacement as a function of the mixing time.

References
I. Callaghan P.T., (1991 ) Principles ofNMR Microscopy, Oxford: Clarendon Press.
2. Bliimich B. , (2000)NMR Imaging for Materials, Oxford Science Publications.
3. Callaghan P.T., Manz B., (1994) "Velocity Exchange Spectroscopy",Iournal of Magnetic Resonanc~
AI06, 260·265
4. Callaghan PT, Khrapitchev A.A., Magnetic Resonance Imaging(in press).
5. Bliimich B.. Callaghan P.T., Damion R.A., Han S., Khrapitchev A.A., Packer KJ., StapfS.,Iournal of
Magnetic Resol/ance(in press).
6. Callaghan P.T., Khrapitchev A.A., Submitted ta/oumal ofMa[7letic Resonance.
7. Ernst R.R, Bodenhausen G., Wokaun A., (1 987)principles ojNuclear Magnetic Resonance in One and
Two Dimensions, Oxford: Clarendon Press.
8. Stapf S., Bliimich B., Magnetic Resonance Imaging(in press).
ALUMINA AND ZEOLITES AS CATALYSTS FOR DECOMPOSITION AND
TRANSFORMATION OF CHLOROFLUOROCARBONS STUDIED BY
MUL TINUCLEAR NMR METHODS

Z. KONYN, I. HANNUS', P. LENTZb, 1. B.NAGy b and I. KIRICSI'


aDepartment of Applied and Environmental Chemistry. University of Szeged.
H-6720 Szeged. Rerrich Bela ter I (Hungary)
bLaboratoire de RMN. Facultes Universitaires Notre-Dame de la Paix.
B-5000 Namur. 61 Rue de Bruxelles (Belgium)

1. Introduction

Heterogeneous catalytic reactions of chlorofluorocarbons (CFCs), like CCl2F2 and their


substituents, hydrochlorofluorocarbons (HCFCs, for example CHCIF 2, are allowed to be
used temporarily, until 2030 [1]), are environmentally important problems. Their
decomposition or transformation to other valuable compounds could lead to novel
technological procedure. Alumina and different zeolites are potential catalysts or supports
in these reactions. Solid state NMR techniques (I3C, 19F, 29Si, 27Al) proved to be powerful
tools to investigate the mechanism of both the surface reaction and the deactivation caused
by the reaction between the reactant and the potential catalysts [2,3].

2. Experimental

HZSM-5 (home-made) and y-alumina (Degussa product) were used for adsorption and
reaction. The reactants CCl 2F 2 and CHCIF 2 were Fluka products.
Surface acidity was tested by the adsorption of pyridine [2]. For HZSM-5, both
Bronsted and Lewis, while on the y-A1 20 3 only Lewis acid sites were measured.
In situ MAS 13C and 19F NMR measurements were performed on an MSL-400
BRUKER spectrometer operating at 100.6 MHz (4.0 Jls (0=1t/2) pulse) and 376.4 MHz (5
Jls (0=1t/2) pulse, 5 s repetition time), respectively. The samples were packed into the
NMR tubes and evacuated at 723 K. The activated zeolite or alumina samples were loaded
with CCl 2F2 or CHCIF 2, then the tubes were carefully sealed to achieve proper balance and
high spinning rate (3.8 kHz). Spectra were recorded after heating the tube at various pre-
selected temperatures.
In separate experiments for 29Si and 27 Al NMR investigations, the solids were
reacted with chlorofluorocarbons. For 29Si (79.4 MHz), a 4 JlS (0=1t/6) pulse was used with
a repetition time of 6.0 s; for 27AI (104 .3 MHz), a 1.0 JlS (0=1t/12) pulse, 0.1 s repetition
time. All chemical shifts are referred to external standards: aqueous AICl 3 solutions for
27AI, tetramethylsilane for 13C and 29Si and liquid CCl3F for 19F NMR.
559
J. Fraissard and O. Lapina (eds.). Magnetic Resonance in Colloid and Interface Science, 559-564.
© 2002 Kluwer Academic Publishers.
560

3. Results and discussion

3.1. MAS NMR SPECTRA OF ZEOLITES

It has been shown that CFC and HCFC molecules are easily transfonned on zeolites. The
earlier IR results [4] and 13C NMR measurements proved, that the final product is CO 2 and
CO in the case ofCChF 2 and CHClF 2, respectively (Figure I).
CC112

l
CHCIF2

RT l_
473K
1 h

CO
473K
4h

I I I I I I I I I I
175 150 125 100 75 175 150 125 100 75
PPM
Figure 1. I3 C NMR spectra of CCI 2F2 and CHClF 2 adsorbed on HZSM-5 zeolite at
different temperature.

The adsorption of CChF 2 on the zeolite


HZSM5-MFI is not followed by any
reaction either at room temperature or at
373 K, only the expected triplet of CChF 2
(8=126.1 ppm, 1cF=325 Hz) was
observed. However, the I3C NMR spectra
taken after treatment at 473 K reveal a
very clear sequence of surface reactions
occurring between CChF 2 and the zeolite
HZSM-5. The clearly visible signal of ____ ~rr ________--__ !
i~ ____ •
COCh is seen to appear after the
-----+~------------------d
fonnation of CCI 4 • This suggests that ------~\------------------(
-------w~----------------b
COCh does not fonn directly from
I I·' 'I' I ' I ' I ' I '
CCI 2 F2• 8() 48 40 -80 ·120 ·1611 -200
PPM
As Figure I shows for the surface
reaction of CHCIF 2 the intennediate Figure 2. 19F NMR spectra of CChF 2
character of CHCh is not so characteristic adsorbed on HZSM-5 zeolite at room
than the CCI 4 in the case ofCChF 2• temperature (a), at 473 K 0.5 (b), I (c),2
(d) and at 573 K 0.5 hours (e).
561

The intensity of the 19F NMR line of CChF 2


adsorbed on HZSM-5 (8= -6.1 ppm) did not a
change at 373 K, but it decreased at 473 K with
increasing time. This signal disappeared at 573 K
meaning that the reaction between CChF 2 and the
zeolite was complete. Simultaneously, a new
signal appeared at -163.4 ppm due to aluminium b
fluoride, as the reaction product of fluorine (see
Figure 2). This is in good agreement with the DC
NMR results, because at 573 K only CO 2 signal
was observed, as the final product of the carbon .
Liquid CHClF 2 has a doublet signal (8= - c
73.5 ppm, l HF =64 Hz). During the adsorption on ...............~
zeolites only a singlet was observed, on NaY at -
76.6 ppm, on HZSM-5 at -78 ppm. No surface
reaction was observed on increasing the
temperature until 573 K.

3.2 . MAS NMR SPECTRA OF Ah03

The adsorption and reaction steps of CChF 2 on y-


Ah03 are somewhat different from those found for
CChF 2 on zeolites. It is well-known that so-called
dismutation (disproportionation) reactions take
place on the surface of alumina [7] . Spectra taken
in the early stages of the reaction are presented in
Figure 3. Decomposition of CChF 2 already occurs
at room temperature. Besides the expected triplet
of CChF 2 a quadruplet attributable to CCIF 3
(&= 125 .2 ppm, JcF =325 Hz) and a weak singlet
characteristic of CO 2 (&=125.5 ppm) complicate
the observed pattern. CC1 4 is responsible for the
appearance of the line at 96.6 ppm.
At 373 K, the reaction proceeds slowly
towards the total conversion of CChF 2 into h
products, but the amount of CO 2 remains small.
After 14 hours, the reaction was continued at 473 I I I I I I I I I I I I I
K, because no significant change was detected at 140 130 120 110 100 90
ppm
373 K within reasonable reaction times.
CChF 2 is totally consumed after reaction at Figure 3. DC MAS NMR spectra
473 K and the intensity of the CO 2 line increased
ofCChF 2 adsorbed on y-Ah03 at
at the expense of the CClF3 signal, while the RT (a), at 373 (b, c) for 0.5, 1, at
relative amount of CCl4 remained constant. 473 (d, e) for 0.5,3 and at 573 K
(f, g, h) for 0.5, 1,3 hours.
562

Finally, at 573 K, the intensities of both the CCIF) and the CCI4 lines decreased and
no product other than CO 2 was observed after 8 hours of reaction. This proves the
intennediate character of CCIF) and CCI 4 in the decomposition. COCh is scarcely detected
in the course of the reaction.
On alumina dismutation reactions take place as primary steps in the case of CHClF2,
too. CHCb and CHF) are the products of these reactions. CO is the final gaseous product of
the carbon content of the freon molecule similarly to the surface reaction on zeolites.
Figure 4 shows the 19F spectra of the surface reaction products of CChF 2 on y- AI 20)
recorded at the same temperature and time sequences as the I3C spectra. These spectra
support the results of I3C NMR
measurements, namely, that on alumina
dismutation reactions take place as
primary steps at room temperature.
Besides the signal of the starting
material CChF 2 at -12.1 ppm another
signal was observed at -33.9 ppm
caused by one of the product of
dismutation reaction CCIF), in good
agreement with the I3C NMR results,
while the other product is CCI 4 •
As the spectra show the intensity
of the CCIF) passes through a maximum. --HH+th-'-------"----f g
This proves the intennediate character of --H+rH~-------------,
---4+~-'---------------d
this compound. -----+~-<--------------c
-----~~-------------b
The final product of the fluorine I I iii I iii I i
content is AIF). Its signal was observed • .. -10 ·111 ·IIt .2tO

at -172.8 ppm, when the temperature of PPM

the sample was increased to 573 K.


Figure 4. 19F MAS NMR spectra ofCChF 2
In the case of CHCIF 2 the
dismutation reaction could be observed adsorbed on y-AhO) at room temperature
by the 19 F NMR, too. Besides the (a), at 373 K (b, c) for 0.5,1, at 473 K (d, e)
doublet of the starting material at -76.76 for 0.5, 3 and at 573 K (f, g, h) for 0.5, 1, 3
and 76.91 ppm the signal of CHF) hours, respectively.
appeared at -83.57 and -83.75 ppm.

3.3. 27 AI AND 29Si MAS NMR RESULTS AND PROPOSED MECHANISMS

For zeolites, MAS NMR spectroscopy is a technique widely used to check the crystallinity
of materials C7 AI and 29Si nuclei). The zeolite samples used in this study were likewise
investigated by these experimental techniques [5]. The results obtained can be summarized
as follows:
Dealumination of the zeolitic framework was identified in both the 27 Al and the 29Si
NMR spectra. The signal due to octahedrally coordinated Al appeared at around 0 ppm,
revealing the removal of Al from the tetrahedral framework position. This fact was also
563

reflected in the 29Si spectra. Since the signal attributed to Si co-ordinated to 1 Al neighbour
decreased, while the resonances due to SiOH defect groups increased. These results
indicated that substantial changes occurred in the crystal structures of the zeolites [6].
On the basis of the experimental findings, a reaction scheme is suggested for the
transformations that occur. The decomposition of CCl2F2 is of second order with respect to
CCI2F2·
On A1 20 3, the following dismutation reactions may occur:
2 CCl2F 2 ~ CCl3F + CCIF3
2 CCl3F ~ CCl4 + CCl2F2
Overall: 3 CCl2F2 ~ 2 CCIF3 + CCl4
In these reactions, twice as much CCIF3 is formed as CCI 4. The partial
decomposition of CCl 4 also being taken into account:.
[AI0 2lW + CCl4 ~ AICl 3 + CO2 + HCl
The CCIF 3 is also transformed into CO 2:
[AI0 2lW + CCIF3 ~ AIF3 + CO 2+ HCl
An interesting feature was observed for the reaction carried out first at 373 K. After
some time, the reaction stopped. Heating the sample at 473 K did not initiate further
reaction. Finally, at 573 K, the reaction went to completion, yielding CO 2 as the final
product. A reasonable explanation is the importance of surface defects necessary for the
transformation of both CCl 4 and CCIF3. At 373 K, the amount of surface defects decreased
to zero after some reaction time. Heating the sample at 473 K did not generate new defect
groups. Finally, when the temperature was high enough to generate new defect groups, the
reaction could continue and proceeded to completion.
The reaction of CCl2F2 on the zeolite HZSM-5 involves quite different behaviour.
The kinetic step of its decomposition is of first order with respect to the CCl2F2
concentration. Our NMR measurements - confirmed by our earlier IR results - allow the
following reaction scheme to be proposed:
1. CCl 2F2 + Z ~ CCI2-Z + Z{2F}
2/a. CCI 2-Z + CCl 2F 2 ~ CCI. + {C} + Z{4F}
21b. CCI 2-Z + Z{2CI} ~ CCl. + Z
3. CCl. + Z ~ COCl 2 + Z{2Cl}
4. COCl 2 + Z ~ CO 2 + Z{2Cl}
where Z means zeolite and the species in brackets are unidentified surface species, the
nature of which is unknown yet.
At the beginning of the transformation, reaction 2/a occurs, leading to unidentified
carbonaceous species. Indeed, some 34% of the initial carbon is rapidly "lost", and the
working catalyst sample becomes black. These carbonaceous species could not be
identified under our reaction conditions. During the transformation of CCl4 into COCl 2 and
the decomposition of the latter, Cl-containing residues are produced, and are used for the
generation of CCl. according to reaction step 21b. This step therefore takes over after the
first phase of the reaction. No "carbon loss" due to the formation of carbonaceous species
occurs subsequently, as confirmed by the experimental data. This is depicted in the scheme.
As the results discussed above demonstrate, different reaction steps are involved in
the mechanism of the reactions that occur over A120 3 and zeolite catalysts.
564

4. Conclusion

The reactions of various chlorofluorocarbons were studied with solid reactants: zeolites and
y-AI 20 3. CCl 2F2 led to the fmal product CO2, while CO was the fmal product starting from
CHCIF 2, proved by 13C NMR technique. The fate of the solid reactants - aluminium
extraction, AIF3 fonnation, etc., - was determined using 29Si, 27AI and I~ NMR.
On zeolites both Bronsted and Lewis acid sites were detected, while no dismutation
reaction products were found. It means that in the case of zeolites immediate attack of the
framework takes place. On alumina dismutation reaction take place as primary steps caused
by the Lewis acid sites, exclusively identified by pyridine adsorption.
For zeolites and alumina 27Al and 19F NMR results proved that the CI content of
chlorofluorocarbons is released as AICl 3 and HCI (or NaCI), while the F content remains
deposited as AIF3 in the solid, because the sublimation point of AlF3 is 1500 K [8]. 19F
resonances are different for the AI-F with zeolitic (-163.4 ppm) and alumina (-172.8 ppm)
origin.

Acknowledgments
This work was performed with the help of grants OTKA T 025248, OTKA F025245,
FKFP-040012000, Hungary, and with financial help from CGRI, Belgium.

5. References

1. Montreal Protocol, Coppenhagen amendment, 1992.


2. Hannus, 1., Konya, Z., Lentz, P., B.Nagy, 1. and Kiricsi, 1. (1999) Multinuclear MAS
NMR investigation of zeolites reacted with chlorofluorocarbons, Mol. Struct., 482-483,
359-364.
3. Grey, c.P. and Corbin, D.R. (1995) 19F and 27AI MAS NMR Study of the
Dehydrofluorination Reaction of Hydrofluorocarbon-134 over Basic Faujasite Zeolites,
J. Phys. Chern., 99,16821-16823.
4. Konya, Z., Hannus, I. and Kiricsi, 1. (1997) Zeolites in the Environmental Protection-
Decomposition of Chlorofluorocarbons over Zeolite Catalysts, Stud. Surf. Sci. Catal.,
105,1509-1516.
5. Hannus, 1., Konya, Z., B.Nagy, 1., Lentz, P. and Kiricsi, 1. (1998) Solid state MAS NMR
investigation of Y-type zeolites reacted with chlorofluorocarbons, Appl. Catal. B:
Environmental, 17, 157-166.
6. Hannus, 1., Konya, Z., Kollar, T., Kiyozumi, Y., Mizukami, F., Lentz, P., B.Nagy, 1. and
Kiricsi, 1. (1999) Spectroscopic investigations of the decomposition of CCl 2F 2 on three
different types of zeolites, Stud. Surf. Sci. Catal., 125, 245-252.
7. Hess, A. and Kemnitz, E. (1994) Characterization of Catlytically Active Sites on
Aluminum Oxides, Hydroxyfluorides and Fluorides in Correlation with Their Catalytic
Behavior,1. Catal., 149,449-457.
8. CRC Handbook of Chemistry and Physics (Ed.: R.C. Weast) 68 th edition, Boca Raton,
Florida, 1988, B-68.
SOLID STATE NMR STUDIES AND REACTIVITY OF SILICA-
SUPPORTED 12-TUNGSTOPHOSPHORIC ACID

Wenxing Kuang, Alain Rives, Michel Fournier, and Robert Hubaut


Laboratoire de Catalyse de Lille" UPRESA8010, Batiment C3, Universite
des Sciences et Technologies de Lille, 59655 Villeneuve d'Ascq Cedex, France

1. Abstract

NMR study of 12-tungstophosphoric acid (HPW) deposited on silica support was


performed. The activity of these solids for n-hexane isomerization was also
measured. The yield of the isomerization products increases with the HPW content
in all cases. The NMR spectra show that, without care, the silica supported HPW
balance between a bulk-like state and a quasi-liquid state without interaction with
the support. The sample with a low HPW content and kept under inert atmosphere is
well dispersed and interacts with silanol groups.

2. Introduction

Heteropolyacids, such as 12-tungstophosphoric acid (HPW) or 12-tungstosilisic acid


(HSiW) are well known to possess purely Bronsted acidity. In the past decades, they
have been widely used in acid catalysed reactions [1,2] but, owing to their poor
surface area, pure heteropolyacids generally show very low catalytic activity. Thus,
supported heteropolyacids, more particularly on silica, were used [3]; as an example,
H 3PW 12040, nH 20 and its acidic salts have been used for the isomerization of n-
alcanes [4,5]. For a better understanding of the role of HPA in the reaction
mechanism, studies on the structure of supported HPA by 31 p MAS NMR has
received considerable attention [6-9] because solid NMR is one of the most useful
probe to study the state of heteropolyacid. This is the reason why we have tried to
correlate the NMR data of tungstophosphoric acid (HPW) with various loading on
silica and the activity of these solids in the n-hexane isomerization.

3. Experimental

Heteropolyacids (HPAs) were prepared in a classical way as described in the


literature [10]. Silica supported heteropolyacids were prepared by the impregnation
method. The slurry of silica and impregnation solutions was constantly stirred at
323 K until water was completely removed, and then dried at 383 K under vacuum.
One part of the HPW (10 wt%) was kept under inert atmosphere to avoid the
moisture before NMR experiment in order to know the influence of trace of water.
565
1. Fraissard and o. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science. 565-569.
© 2002 Kluwer Academic Publishers.
566

31 p MAS NMR spectra were recorded at room temperature on a Bruker ASX400


spectrometer. Concerning the isomerization of n-hexane, the samples were plac~d
into a V-type glass fixed bed micro reactor, first reduced at 498 K for 4 h in
hydrogen with a flowing rate of 22 mL·min- 1 under atmospheric pressure, and then
their catalytic properties were evaluated under the same conditions with hydrogen
passed through a saturator containing n-hexane (Scharlau company). The WWH
(weight ofn-hexane per weight of catalyst and per hour) was 0.36 h-I. The products
were analyzed by gas chromatography with a SE-30 capillary column and a flame
ionisation detector.

4. Results and discussion

The surface areas of silica-supported HPW with different loading are listed in table
1. The surface area of pure silica support is 320 m2 .g- l • Noticeably, with the increase
of HPW loading from 10 wt % to 60 wt %, the surface decrease drastically from 267
m2.g- 1 to 105 m2.g- l • The decrease mainly results from the blocking silica pores due
to the formation of large crystalline particles of HPW samples as exhibited from
XPS measurements [II].

Table I surface area of the pure silica and HPW/Si0 2 catalysts

HPW loading (wt%) 0 10 20 30 40 50 60


Surface area (m2.g"I) 360 267 240 192 149 124 105

The effect of the various loading on the reactivity for n-hexane isomerization of
silica supported HPW samples is in table 2. The conversion of n-hexane increases
remarkably with the increase of HPW loading while the selectivity for the
isomerization decreases. Simultaneously, the selectivity for isobutene and
isopentane, which are the main cracking products, increases. However, the yield of
isomerization products increases with the HPW content.

Table 2: Effect of different loading on the reactivity of silica-supported HPW after 5


minutes reaction.

HPW Conversion of Selectivity for the Isomerization Selectivity for the


loading n-hexane (%) isomerization (%) yield (%) isobutene and
isopentane (%)
10 wt% 3.7 79 2.9 13
20wt% 8.8 71 6.2 22
30wt% 16.6 58 9.6 34

Reaction conditions: 498 K, hydrogen flowing rate = 22 mL·min- l , WWH (weight of


n-hexane per weight of catalyst and per hour) = 0.36 h-I.
567

The figure 1 displays the NMR spectra of, respectively, HPW in aqueous solution,
solid HPW and various silica supported HPW. Aqueous solution of HPW presents
an unique signal at -13.8 ppm while pure solid HPW spectrum has a mean peak
located at -15.34 ppm with a shoulder at -15.27 ppm. For silica supported samples,
only one broad signal, shifted around -15.3 ppm is visible corresponding to mainly
crystalline HPW on Si02 . This is in agreement with the previous results [12-13].
However, some authors find also a shoulder at -11.7 ppm and claim that is due to
partially dehydrated HPW [12,14]. We never observe this peak except with the 10
(wt%) sample which is kept under inert atmosphere, but, XRD and FTIR data of this
sample, show that HPW structure is retained on the support without neither any
other species nor partial reduction. So, we can assume that the 31 p NMR signal
observed on the particular sample at -11.8 ppm is rather due to Keggin PW 12 0 40 3•
ion dispersed onto the support with an important interaction with Si0 2.Thus, the line
at -11.8 ppm could be assigned to the interacting species such as
(Si(OHh)\H 2PW 12040)". This signal represents roughly half of the total HPW
deposit on the support but it appears as a weak signal because of its more important
half width (125Hz vs 50Hz).
Concerning the broad signal around -15 ppm for silica supported HPW, we claim
that we have samples with a water shell of variable number of molecules thickness
and, this leads to a more or less good mobility for HPW species which decreases the
interaction with the support. Ultimately, when the water amount is larger, we obtain
a quasi-liquid state and the NMR lines of HPW shift up to -13.8 ppm, which is
characteristic of HPW in solution. Besides, previous report [9] claims that there are
two lines around -15 ppm and -14 ppm in the spectra of silica supported
tungstophosphoric acid up to a loading of 25 wt% while at higher loading, (30-50
wt%) the former line is only present.
The whole results are in good agreement with the expected effect of silica
interacting HPW.
-First, the protonated silanol groups in interaction with bridged oxygen
atom in HPW structure is expected to cause a deshielding of the 31 p resonance,
because of the decreasing of the diamagnetic screen induced by oxygen atoms
(probably the bridging oxygen atoms Oc and Ob). This is observed with the low
HPW content sample kept under inert atmosphere.
-Secondly, a strong interaction through dipolar coupling is expected to
increase the spin-spin relaxation (T2) leading to a larger line. This is observed with
the "dry" sample, which has a signal at -11.8 ppm, which indicates a strong
interaction of HPW with the support
-Thirdly, the "wet" samples present a broad signal, which overlaps the
hydrated bulk HPW and the moving species through the water layer on silica. This
signal varies between -15.3 ppm and -13.8 ppm, depending on the water amount.
HPW balances between a bulk-like state and a quasi-liquid state.

Concerning the n-hexane isomerization activity, there is no difference between the


two 10(wt%) samples. This probably is due to the high hygroscopic nature of silica
supported HPW, which in few time gives rise to a rehydrated sample into the
reactor, and we could passed a hot stream of inert flow before test. Works are in
progress to perform such experiments and also the T1 and T2 measurements to
568
obtain more information about the various HPW species on silica and more
particularly the evolution ofT! value with the amount of water molecules.

, 1"" ",I~
a)

b)
" i
-12.0

' ------ -
-13 .5

I
-15.0 - 16.5

: f)
-12.0 -13 .5

;
-15.0

-- --&
-- -
-16.5

~-

I'" "'!I im' !i' l'i~,mt'1 l' ' ' ' ' ' "' 'i,m,~~m'I'I'T'l'm'i"I'ml'i, L~ ~
-12.0
,,,,,,,,,,,,
, 11,,,,,'
-13.5 - 15.0 -16.5
I -12.0 -13.5 -15.0 -16.5

- 12.0 -13 .5 -15 .0 -16.5 -12.0 -13.5 -15.0 - 16.5

-12 .0 -13.5 -15 .0 -16.5 -12.0 -13 .5 -15.0 -1 6. 5


(ppm)

Figure I: NMR spectra of a) HPW in solution, b) solid HPW, c) 5% HPW/Si02, d) 10% HPW/Si02 dry ,
e) 10% HPW/Si02 wet, f) 20% HPW/Si02, g) 30% HPW/Si02, h) 50% HPW/Si02.

In summary, we can assume that the observed shift at about -15.3 ppm of the "wet"
silica supported HPW is due to the heteropolyacid. This band can be decomposed in
two lines. The former being due to the bulk HPA while the latter is due to the more
or less hydrated species without interaction with the support. On the contrary, the
sometimes-observed barid at -11.8 ppm probably can be attributed to dry dispersed
species with strong interaction between HPW and silica.

5. Acknowledgements

W.K is grateful to the financial support of CNRS and the authors thank the RMN
service for the spectra.
569
6. References

[I) Kozhevnikov, LV. (1988) Catalysis by heteropolyacids and multicomponent polyoxometallates in


liquid phase reactions, Chern. Rev. 98,171-198.
[2) Mizuno, N., Misono, M. (1998) Heterogeneous catalysis, M., Chern. Rev. 98, 199-218.
[3) Izumi, Y., Hasebe, R., and Urabe, K.(1983) Catalysis by heterogeneous supported heteropolyacids,J.
Catal. 84, 402-409.
[4) Na, K., Okuhara, T., and Misono, M. (1993) Skeletal isomerization of n-butane catalysed by an acidic
caesium salt of 12-tungstophosphoric acid; J chern. Soc Far. Trans. 91,367-373.
[5) Guisnet, M., Bichon, P., Gnep, N.S., Essayem, N. , (2000) Transformation of propane, n-butane and
n-hexane over HlPW 12040 and cesium salts. Comparison to sulfated zirconia and mordenite catalysts
Topics Catal. 11-12,247-284.
[6) Mohana Rao, K.; Gobetto, R.; Iannibello, A., Zecchina,A. (1989) Solid state NMR and IR studies of
phospho molybdenum and phosphotungsten heteropolyacids supported on Si02 ,AhOl and Si02-AhOl, J.
Catal. 119, 512-516.
[7) Mastikhin, V.M.; Kulikov, S.M.; Nosov, A.V.; Kozhevnikov, LV.; Mudrakovsky, I.L.; Timofeeva,
M.N, (1990) Proton and phosphorus-31 MAS NMR studies of solid heteropolyacids and HlPWI2040
supported on silicon dioxide. J. Mol. Catal. 60,65-71.
[8) Lefebvre, F. (1992) Phosphorus-31 MAS NMR studies of HlPWI2040 supported on silica J. Chern.
Soc., Chern. Cornrnun., 756-757.
[9) Kozhevnikov, LV.; Kloetstra, K.R.; Sinnema, A.; Zandbergen, H.W.; van Bekkum, H. (1996) Study
of catalysts comprising heteropolyacid HlPW 12040 supported on MCM-41 molecular sieve and
amorphous silicaJ. Mol. Catal., A: Chernical114, 287-298.
[10) Rocchiccioli-Deltcheff, C., Fournier, M., Franck, R., and Thouvenot, R.,(1983) Vibrationnal
investigations ofpolyoxometallates: evidence for anion-anion interactions and molybdenum(VI) and
tungsten (VI) compounds related to the Keggin structure, Inorg. Chern. 22, 207-210.
[II) Kuang,W .; Rives,A.; Foumier,M.; and Hubaut,R. Structure and reactivity of silica-supported 12-
tungstophosphoric acid, J. Catal., submitted.
[12) Mizuno N. and. Misono M (1987) Reaction of hydrogen-deuterium and hydrogen-oxygen over
HlPW12040, Chern. Lett. 1195-1197
[13) Chang T.H. (1995) NMR characterization of the supported 12-heteropolyacids , J. Chern. Soc. Far.
Trans. 91 ,375-379
[14] Kanda Y.,.Lee K.Y, Nakata S.I., Asaoka S.and Misono M. (1988) Solid-state NMR of
polytungstophosphoric acid ,Chern.Lett.,139-140.
MOLECULAR MOTION IN CROSSLINKED POLYMERS AS STUDIED BY
STIMULATED SPIN ECHO

T. P. KULAGINA 1, G. E. KARNAUKHI, G. T. AVANESYAN 1, and


F. GRlNBERG 2
lInstitute of Problems of Chemical Physics, RAS
J University Ulm, Germany

The methods of stimulated spin echo are being widely used in studies on the molecular
dynamics in polymers, liquid crystals, and porous and other media [1 , 2]. These methods are
based on measuring the amplitudes of primary and stimulated echoes after sequences of two or
three pulses in a gradient magnetic field.

(TCl2)_x- (t 1)- (1t /2)x- (t 1)- (Primary echo)- (t 2- t 1)- (1tI2)y- (t 0- (Stimulated echo) ( I)

Let us consider the dynamics of three equivalent spins 112, Sk, Sl and Sn. The local rotating-

frame Hamiltonian contains the RF term HrJ. the local field gradient offset Hg , and the secular

part of dipolar coupling (DDI) H/O) :

(2)

where Hg = -n ng(r) (Skz + Slz +S nz), nng(r) = /).

571
J. Fraissard and O. Lapina (eds.J. Magnetic Resonance in Colloid and Interface Science, 571-576.
© 2002 Kluwer Academic Publishers.
572

Ii nd =b, b=bkl- (3co/ 6kl- 1)/ r k/ is the dipolar coupling constant, which is the same for
all spin pairs in the system of tree spins. The same situation may appear if three- spin fragment

rotates fast about a certain axis. In this case, constant b cOlTesponds to the residual coupling

constants averaged due to fast rotation similar, for instance, to the rotation of CH3 fragment.

We restrict ourselves to the high-temperature approximation, P(r)-1- parr), and consider the
evolution of the polarization operator a(r) .
It is convenient to divide the space R of the spin states into two subspaces, RI and Ro.

(4 )

where PI.O are the projection operators on R), Ro, respectively. There are no dipole-dipole
interaction terms in it (6), but only double degenerate spin Yz, which interacts with the external
magnetic stationery and rf fields. The evolution of the ~stem is described by the known
expressions [2, 3). On the subspace RI the effective spin 3/2 states are affected by pulses, Eq.
(1), the magnetic field, and the effective quadrupole interaction with the Hamiltonian
o
P1Hd =(bI2)P,(Sz -5/4). (5)

Thereby the problem is reduced to the calculation of evolution within each of the subspaces, Ro
and RI.

Analytical expressions are obtained from the density matrix evolution equation which takes into
account the transfer and precession of coherence during and between the pulses [2) .

(6)

(7)

To analyze the effect of molecular motion on the echo signals Eqs. (7,8), it is necessary to take

into account [I} the fluctuations of the dipole-dipole coupling, and also the fluctuation of phase

shifts between pulses, by means of correlation functions over all spins of the system:
573

In addition, the fluctuations of phase shifts between RF pulses

by means of con'elation function k("t) are:

(ob(t')&(t") = G(,) = (o2b)k(,) < a 2 qy >= 2 < alb>


,
f ('I -t)k(t)dt (8)

(9)

(10)

At present to get a quantitative description, two- and three-spin theoretical model are being
used [1, 2]. Until now, the influence of fluctuations of dipole interactions due to molecular
motion in polymers have been taken into account by the exponential correlation function.
However, it did not give quantitative description of experimental data.
In this communication, we present a theory for analysis of the influence of three-pulse
sequences on the dynamic of spin system and spin-echo signals [2 , 3]in crosslinked polymers
with a quasi-network of entanglements.
The dynamic model of polymer chain was suggested in the theory [4] with tree models
of polymer motion observed experimentally within three temperatures ranges : Tg+ 30° < T< Tg
+ 60°, T >Tg + 100° and interval between with the following correlation function of molecular
motion:

(II)

where k, (x)-the function of Bloembergen-Purcell-Pound,

k1() "
r = ~Cje
-r / r~j) ' c
r =m).z'nr(;)c ' (12)
i
k:(r) is the function of Kargin-Slonimskii-Rouse of small-scale and long-scale motions,
574

N
k2(r)=No-'Ie-p2R2T'TCN02, (13 )
p=1
k.](8, N) is the function of motion of polymer chain with N statistical segments over the high-
temperature range, is independent of correlation time and is determined by chain conformation
[4]:

(14)

The signal of free induction decay (FID) was taken into account in the Anderson-
Wiess model with correlation function (11). The expression for FID was obtained after
isotropic averaging (11) over all direction of chains and over chain lengths N with distribution
function peN) :

ff f(t - r)k(r,e, N)P(N) sin BdBdNdr


/<IORI2

In G(t) = -02b
oI 0 (15 )

The simulations of FID, primary and stimulated echoes were made in the theory [I]
and [2] with correlation function (11) and showed the coincidence of both theories for the
parameters which have been considered (Fig. I ). The analysis of calculations allowed
determining the time of molecular motion correlation over wide temperature range. The
calculations were made for time in units <,fb>14 with the Gaussian function peN) and

parameters: <b> = 0, '2 = 10, No = 100.


575

o 0.04 0.08
1 ....11------- ----- --- - ----- - -- _. - 1 - . -- - -- - - - -- . - - -- - -- - --

''\~,II~
\ ~-n -OOO
\
'.
~--..:,O• ........
0 8 0
00 0 -- 0 _

.....-'.
0.1 \ -~-.~.~_
,..- . - . .-- O· g--B- O 0

" . '-.--"'---..
'\
'\
Prim. Echo (2)
0.01 -_. _- StIrn . Echo (2)
• Prim. Echo (3)
o Stirn . Echo (3)
_ .. _. FlO
0.001

Figure J. Primary and stimulated echoes and FID as calculated/rom expressions (6)-(15) for
Tc =10--4.

o 0.004 0.008

0.1

0.01 ·
Stirn. Echo Theor.
• Prim. Echo Subtr. 0 00. 0
o Stirn. Echo Subtr. - Q

0.001

Figure 2. Comparison o/the theoretical and experimental echoes in PEG.

In our opinion, the coincidence of simulations in theories [1, 2) is due to using the correlation
function (11), which was derived in (4) for a multispin system. The difference between the
576

theories is also observed for other correlation functions (e.g., exponential [2])\and depends on
the type of compound under study.
For comparison with experimental data obtained for polymer PEO, we used the
expressions (6) and (7) taking into account the correlation functions of molecular motion (II).
The experiment was performed in PEO with M,;=186 000, at the temperature 70C. It is well
known that polymers with Mn > 10 5 form a quasi-network of entanglements. Because of
heterogeneity of the polymer at this temperatures, the observed primary and stimulated echoes
were represented by a sum of signals. After subtraction of the exponential component of the
echo signals, the reduced signals we attribute to the existence of a quasi-network of
entanglements. The theoretical and experimental curves are compared on Fig. 2. The
calculations were performed for time in units <clb>14, with the Gaussian function P(N), and
parameters: <b> = 0, '2 = 20, No = 180, 'c = 10 4 (Fig. f) and 'c = f.66 f02 (Fig. 2). The
obtained expressions are supposed to be used in quantitative analysis on polymer networks, to
determine the correlation functions, and correlation times of molecular motion, on the basis of
measured NMR spectra.
The work was supported by the Russian Foundation for Basic Research (project no. 00-03-
33034 ).

References

I. Grinberg, F. and Kimmich, R. (1995) Characterization of order fluctuations in liquid crystals by the dipolar-
correlation etlect of the stimulated echo, J. Chem. Phys. 103,365.
2. Kulagina T. P., Karnaukh G. E., Kimmich R., Grinberg F., Avanesyan G. T. (1999), In Russian. Signali
pervichnogo i stimuliravannogo ekha v trekhspinovoi sisteme. "Struktura i dinamika molckul. sistcm". vip. VI,
Kazan, Unipress, ,p. 327 .
3. Kulagina T.P., Kall1aukh G.E., Avanesyan G.T., In Russian:Analiz dinamiki sshilykh polimerov s videlennoi
Irekhspinovoi gruppoi me/odom s/imulirovallllOgo ekha. II Vserossiiskii Karginskii Symposium,
Chell1ogolovka, 2000, p.2-85.
4. Kulagina T.P., Marchenkov V.V., Provotorov B.N., In Russian:Teoria YaMR spek/I'ov v po/imernikh
selkakh. Vysoko11lol. Soedin ., 1989, v.31 p. 381.
STRUCTURAL CHANGES IN ZrO z CATALYST DOPED WITH Fe AND Cu.
EPRSTUDY.

M. LABAKI, J.-F. LAMONIER, S. SIFFERT, E.A. ZHILINSKAYA,


A. ABOUKAIs
Laboratoire de Catalyse et Environnement, EA 2598, ULCO,
145 avenue Maurice Schmann, 59140 Dunkerque
e-mail :jilinska@univ-littoral.fr. Fax: 03-28-65-82-39

1. Introduction

The catalytic performances of solids depend particularly on the crystalline structure of


solids and on the quantity, valence state and local structure of the active chemical
elements. Zirconia shows polymorphism [1] and the three most common phases are :
monoclinic, tetragonal and cubic. Special attention has been paid to the retention of the
metastable tetragonal phase at low temperatures. Some structural and composition
parameters have been established which influence the retention of the metastable
tetragonal phase, including grain size [2,3], grain morphology [4], type and amount of
stabilisers [5,6] and the matrix constraint [7].
In this paper, MxZr oxides (M = 'Cu, Fe) were synthesised by coprecipitation or by
impregnation of M nitrate on zr02 or ZrO(OH)2' The aim of this work is to evidence
the phase transformation and the morphology changes due to the type, the quantity of
doping metals and the thermal treatment conditions. The samples were characterised by
Thermal Analysis, Electron Paramagnetic Resonance (EPR) and Raman Spectroscopy.

2. Experimental

Different MxZr samples (M = Cu, Fe; x(atomic ratio) = 0; 0.01; 0.1; 0.5) were
synthesised by three different ways : coprecipitation (Mx-Zr), wet impregnation on the
zirconium hydroxide ZrO(OH)2 (M/ZrOH) and wet impregnation on the zirconium
oxide (zirconia) previously calcined at 873K (M/Zr). The precursors were
ZrOCI 2.8H20, CU(N03)2,3H20 and Fe(N03)3.9H20. The precipitating agent was
NH3.H20 0.7M. These samples were dried at 373K overnight and then calcined 4h
under air flow (2L.h") (dynamic conditions) at 723K and 873K (2K.min"). Some of
them were heated for a whole night in static conditions (without a continuous flow of
air) at temperatures from 373K to 1473K.
Thermal analysis measurements were performed using a Netzsch STA 409C equipped
with a microbalance, differential thermal analysis (DTA) and a flow of air
577
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 577-583.
© 2002 Kluwer Academic Publishers.
578

(75 rnL.min· I). The temperature was raised at a rate of SK.min· 1 from room temperature
to desired temperature. EPR spectra were recorded at 77K and at room temperature with
a Bruker EMX spectrometer working at the X-band microwave frequencies (9.3 GHz).
The magnetic field is modulated at 100KHz. The g values were determined from the
simultaneous precise measurement of the magnetic field and the microwave frequency.

3. Results and Discussion

3.1. THERMAL ANALYSIS

r
~--

893K
I CU/Zr
exo
lOO8K
.... I
0.5 38K
~
C 1038K
~
:e< 738K I 228K

653 753 853 953 1053 1153 1253


Temperature (K)
"Figure J. " DT A curves of Cu-Zr coprecipitates during dynamic calcination.

These DTA curves (figure 1 and figure 2) presented two exothermic peaks not
accompanied by any mass loss. The first peak is due to the crystallisation of the
metastable tetragonal phase of Zt02[8]. The tetragonal phase is slowly converted to a
monoclinic phase with the changing of the calcination temperature [9]. The more the Fe
or Cu content, the more the crystallisation peak is delayed.
For Fe/Zr ratios lower than 0.13, the second exothermic peak is not visible that means
that we need a higher Fe/Zr than CulZr ratios to observe this peak before 1273K. This
peak could be assigned to either the transformation of Zt02 from tetragonal form into
cubic form either to the formation of Fe-Zr or Cu-Zr compounds and/or to the
crystallisation of the copper or iron oxides [10].
For the coprecipitated samples, in the case of copper, the 6.T/6.C value is 315
(6.T=difference between the temperatures of zirconia crystallisation (first peak),
6.C=difference between the MlZr atomic ratios for the maximal and the minimal
concentrations) while in the case of the iron the 6.T / 6.C value is 236, that indicates that
the copper has a more important effect of delaying the crystallisation temperature.
579

FelZr i exo
0,50~_------ _ _--,

0,13

0,06

0,01

572 772 972 1172


Temperature (1<)

"Figure 2." DTA curves of Fe-Zr coprecipitates during dynamic calcination.

During the calcination of the impregnated CulZrO(OH)2 or Fe/ZrO(OH)2 samples, the


same trends are observed. However, for the same Fe/Zr or CulZr ratio, the
crystallisation peak is less delayed than in the case of the coprecipitates samples. The
crystallisation temperature is 724K for Feo)ZrO(OH)2 and 710K for Feoo .!ZrO(OH)2,
716K for CUo.oI/ZrO(OH)2 and 840K for CUo)ZrO(OH)2.
The delays of the crystallisation temperature of zirconia in the presence of Cu or Fe
induce the stabilisation of tetragonal zirconia phase at lower temperature. This result
has been evidenced by XRD and Raman spectroscopy.
For Cu or Fe impregnated over zirconia oxide no crystallisation peak is detected in
thermal analysis because zirconia was previously calcined and then crystallised before
impregnation. In addition to this fact, the second exothermic peak is not observed.

3.2. EPR

3.2.1. Influence of the Kind of Thermal Treatments

3.2 .1.1. Static conditions. In fig. 3, all the samples treated at T~973K exhibit the EPR
spectra consisting of two signals. The first one is attributed to isolated Cu2+ ions (I), the
second one to Fel + impurity ions (II). At Tc=1093K, the first signal is composed of two
types of Cu2+ spectra with different EPR parameters. Parameters of these sites 1 and 2
are present in table I. The resolution of the EPR spectra for the samples treated under
static conditions is worse than that under dynamic conditions (see below). In order to
determine the exact values of EPR parameters, we have done simulations of the spectra
using the SimFonia Bruker program. Moreover, at Tc=1093K, the formation of Cu2+
agglomerates starts and progresses up to Tc=1223K (signal III). The treatment at
1223K<Tc~1373K leads to the disappearance of all Cu2+ signals from EPR spectrum.
580

o 1000 2000 3000 4000 5000


Magnetic fiekl (Gauss)
"Figure 3." EPR spectra recorded at 17K ofCuo.OI"Zr coprecipitate calcined at different temperatures in
static conditions.

Besides, a new phenomenon was observed at this temperature treatment : formation of


Z~+-CU2+ pairs (zoom A in fig. 3). These pairs give rise to EPR signal at g=4.3 (dM=2
forbidden transition) with Hyper Fine Structure (HFS) constant A=8.3G due to 9tZr
isotope. The quantity of paramagnetic species increases with Tc up to 1223K,
whereupon it suddenly decreases. In these static conditions where the oxygen is not
enough, we can guess the formation of Cu+Zr4+0x compound without any paramagnetic
species. The observed Z~+_Cu2+ pairs can be supposed as defects in this compound and
can be explained as a result of the migration of the electron from Cu(I) to Zr(IV). The
formation of a new compound is supported by the apparition of a new phase (XRD) and
of not identified addition Raman bands.

3.2.1.2. Dynamic conditions. In this case, we have the same three EPR signals I, II and
III. One type of isolated Cu2+ ions are observed up to Tc=873K. The parameters are
given in table I. In the dynamic conditions, the continuous flow of air creates an
atmosphere rich in oxygen.
The difference between the dynamic and the static conditions is remarkable. Firstly, the
one type of isolated Cu2+ species in dynamic treatment samples is different compared
with the two types ofCu2+ sites in the static treated samples for all the Tc. Secondly, the
most important difference is observed at Tc=1473K (Figs 3 and 4). Indeed, in the
dynamic case, EPR signal III is still present in contrast to the static one where a new
signal at g=4.3 is observed. It can be supposed the formation of different Cu-Zr
compounds according to the both treatments, for the static condition Cu+Zr4+0x and for
the dynamic one Cu2+Zr4+0z (z>x) and/or CUO. The formation of such solids could be
581

responsible for the second exothermic peak seen in the thermal analysis (Fig. 1), which
is shifted with copper content.

Not calcined GAIN


1
i
i 2
723K

'c::s I
! 873K 2
~
g
:e
~ 1273 K
C-
.;;;
c 0.4
~
..5 i 1473K

1000 1500 2000 2500 3000 3500 4000

Magnetic field (Gauss)


"Figure 4." EPR spectra recorded at room temperature ofCuo.ol-Zr coprecipitate calcined at different
temperatures under a flow of air.

3.2.2. Influence of the Preparation conditions

3.2.2.l.Cu-Zr. The EPR parameters (Table I) are nearly the same in case of
CU/ZrOz(873) and CU/ZrO(OH)2' In the coprecipitates samples, the copper is located in
another type of sites since the EPR parameters are different · (Table I). This is in
convenience with the fact that in the coprecipitated samples the copper is more
dispersed and mainly present in the bulk whereas in the impregnated samples it is
essentially present on the surface.
For Tc:2:1273K, we observe the formation of the Cu2+ agglomerates and/or of the Cu-Zr
compounds which give rise to a large signal with a g;so value (table I).

3.2. 2. 2. Fe-Zr. In Fig. 5, two signals, with gt=8-9; g2=4.8; go=4.28 values due to the
isolated Fe3+ ions located in an octahedral or tetrahedral sites with a near rhombic
distortion, are observed for coprecipitate Fe-Zr samples. The signal with g3=2.01-2.14 is
attributed to the Fe3+ agglomerates and disappears completely for Tc=1273K.
It can be supposed that at Tc21273K a new compound Fe-Zr is formed which leads to
the disappearance of EPR signal III. The formation of a such compound could be
related to the presence of the second exothermic peak observed in DTA experiments
(fig. 2). In contrast, the EPR signal III is absent for the two types of impregnated Zr-Fe
samples up to Tc=1273K. In this case, the signal can be attributed to FeP3 phase which
has been detected by Raman and X-ray diffraction. The formation of FeP3 phase for
the impregnated samples is may be due to the higher concentration of Fe on the surface.
582

TABLE I : EPR parameters of the different preparations of Cu-Zr compounds.

Samples Treatment T(K) A/~G) gil g.l giso


ClIo.oJ/Zr02 (873) 723 and 873 134 2.34 2.067
(imp.) dynamic 1273 2.177
ClIoo/ZrO(OH)2 723 and 873 134 2.34 2.073
(imp.) dynamic 1273 2.294
ClIo.oJ - Zr 723 and 873 125 2.35 2.069
(cop.) dynamic 1273 2.285
1473 2.200
ClIo.OJ - Zr (cop.) 723 133 2.34 2.073 2.078
static 973 112 2.36 2.065 2.4
1093 1)115 2.36 2.059 2.54
2)162 2.25
1223 2.23
1473 2.24

GAIN
& zr+
~r02 873K ~/ :;r],
. \ -:f \.:--------- if v---~---i
-

Jf i ./'l.- ImK'
-~--------___i
2

:e$ I 873K
I
~
.;;;
1,25
B 723K
.5

550 1050 1550 2050 2550 3050 3550 4050


Magnetic field (G1uss)
"Figure 5." EPR spectra recorded at 77K of the Feo.ol-Zr coprecipitates calcined at different temperatures
and of zr02 calcined at 873K.
583

4. Conclusion

1) The presence of Cu or Fe stabilises the tetragonal phase of zirconia with a delay of


the crystallisation temperature. This delay increases with Cu and Fe content. The effect
of Cu is more important.
2) The static and dynamic conditions of treatments for Cu coprecipitated samples lead
to the formation of different isolated Cu2+ sites in zr02 matrix and to supposed different
Cu-Zr compounds at Tc~1373K. The Cu2+-Zr+ pairs for the static treatment are
observed.
3) In dynamic conditions, it is demonstrated that the synthesis methods, impregnation
and coprecipitation, give rise to :
a) formation of different types of Cu2+ sites (in
volume and at surface).
b) different influence ofCu and Fe on the
crystallisation phenomena.
c) supposed formation of Cu-Zr and Fe-Zr
compounds in coprecipitated case and of FeP3 in
impregnated one.
4) The Raman and XRD experiments partially confirm the results obtained by DTA and
EPR studies.

Acknowledgements

The conseil general du Nord, the region Nord-Pas-de-Calais and the European
community are gratefully acknowledged for fmancial support for the purchase of EPR
spectrometer and thermal analysis.

References

\. M. Yoshimura, Ceram. Bull. 67 (1998) 1950.


2. Y. Murase, E. Kato, 1. Am. Ceram. Soc. 62 (1979) 527.
3. R.C. Garvie, J. Phys. Chern. 69 (1965) 1238.
4. R. Srinivasan, Chern. Mater. 5 (1993) 27.
5. FJ. Berry, M.H. Loretto, M.R. Smith, Solid State Chern. 83 (1989) 91.
6. S. Popovic, B. Grzeta, G. Stefanic, J. Czako-Nagy, S. Music, J. Alloys Compo 241 (1996) 10.
7. L.S. Pan, S.Horibe, 1. Mater. Sci. 31 (1996) 6523 .
8. T. Yamaguchi, Catalysis Today, 20 (1994) 199.
9. J. Matta, 1.-F. Lamonier, E. Abi-Aad, E.A. Zhilinskaya, A. Aboukais, Phys. Chern. Chern. Phys., I
(1999)4975.
10. D. R. Lide, handbook of chemistry and physics, 77 1h Edition 1996-1997, ISBN 0-8493-0477 -6.
IH-NMR-INVESTIGATION OF THE PHASE TRANSITION OF THERMO-
REVERSIBLE POLYMERS IN SOLUTION AND AT INTERFACES

A. LARSSONI.3, D. KUCKLING 2, M. SCHONHOFF 1

J Max-Planck-Institute for
Colloids and Interfaces.
D-14424 PotsdamlGolm. Germany
J Institute for Macromolecular and Textile Chemistry. TV Dresden.
D-O 1062 Dresden. Germany
3 current address: Institute for Surface Chemistry.
p. 0. Box 5607. SE- 11486 Stockholm. Sweden

Abstract

The thermoreversible polymer Poly-N-isopropylacrylamide (PNIPAM) shows a coil to


globule transition at 32°C in solution. We address the question of whether charged groups
in the polymer are affecting the phase transition, and study their influence in solution and
at the interface.
IH spectra under liquids conditions and PFG-NMR diffusion coefficients are measured at
various concentrations in solution. A decrease of the IH liquid signal at the phase
transition temperature monitors the formation of solid globules. The transition is sharp and
for the copolymer no significant influence of the charges on the transition is found. PFG-
NMR diffusion coefficients prove that the size of the polymer coils is maintained until
close to the transition.
Both polymers are adsorbed to colloidal silica (Cab-O-Sil) to monitor the phase transition
behaviour in the restricted geometry of an adsorption layer. A liquid IH signal from loops
and tails is observed, the intensities are decreasing with temperature, which is interpreted
as a phase transition of the loops and tails. The phase transition is substantially broader
than in solution, especially at low surface coverage. For the copolymer, the transition is
further broadened, and mobile segments remain even above the phase transition. We
attribute this to a comparatively mobile arrangement of the copolymer layer: Due to the
electrostatic repulsion from the surface and between polymer segments, the copolymer
layer is probably confined in a configuration extending further from the interface.
Thus, while the copolymer charges do not affect the phase transition in solution, they
determine the phase transition properties at a charged interface.

1. Introduction

The reversible conformational changes occurring in thermoreversible polymers in solution


make them interesting compounds for applications with a demand to switch material
585
1. Fraissard and O. Lapina (eds.). Magnetic Resonance in Colloid and Interface Science. 585-590.
© 2002 Kluwer Academic Publishers.
586

properties by external conditions. Particularly the phase transItIon of poly-N-


isopropylacrylamide (PNIPAM) in aqueous solutions has been studied in great detail [1].
PNIPAM forms random coils at low temperatures, which collapse to solid globules in a
sharp transition at the lower critical solution temperature (LCST) around 32°C.
Our interest lies in the phase transition properties in ultrathin organic films, where
electrostatic interactions are a relevant coupling mechanism for layer formation. For
applications of thermoreversible polymers in layered materials, thus the effect of charged
segments on the phase transition, and furthermore the transition in the restricted geometry
of adsorption layers is of interest, which is the subject of our study.
In studies of the phase transition most commonly light scattering is employed, where the
particle size, or the total thickness of an adsorption layer can be determined. Here, we
apply liquid state 1H NMR, where the selective detection of liquid spins offers a simple
possibility to probe the mobility of polymer segments. The task of this work is to
investigate the phase transition of PNIPAM by IH NMR under the influence of charged
groups and in adsorption layers as compared with the transition of free coils in solution.

2. Materials and Methods

Poly(N-isopropylacrylamide} homopolymers (1) (Mw = 10 5 g/mol, and 3.5 ' 105 g/mol) are
used as obtained from Polymer Source Inc. A copolymer (l) ofN-isopropylacrylamide and
a carboxylic acid containing acrylarnide co-monomer, see Fig. 1, is synthesised by radical
polymerisation as described previously [2]. The co-monomer content is 10 mol-%, in
agreement with the feed composition, and the molecular weight (Mw) is 1.3 '10 5 g/mol.

~J\-­
Fig. I: Chemical structure of
~A.nNH... a~A'mNH
a
1~ 5· ~6.
PNIPAM (!) and charged
statistical copolymer ill.

eOOH

PNIPAM (1) Co-PNIPAM (~)

Colloidal silica 'Cab-O-Sil' (Fluka), surface area 200 m2/g, is used as obtained. For
adsorption samples, silica and polymer solutions are prepared as described previously [3].
A Bruker 400 MHz A vance spectrometer is used for all NMR investigations. 1If spectra
are measured by acquisition after a 90° pulse. The prescan delay is lOs, and the dead time
before acquisition is 40 Jls. These experiments allow the detection of 'liquid' spins only, i.
e. spins with a sufficiently slow T2 relaxation time (T2 > dead time). Diffusion experiments
are performed in a pulsed field gradient probe head (Bruker DIFF 30) using a stimulated
echo with gradient pulse lengths of 8 = 1 ms and variable gradient strength, separated by tl
= 100 ms.
587
3. Results and Discussion

3.1., COIL STRUCTURE REARRANGEMENTS BELOW THE LCST

The structure of the polymer coils in solution (D 20) is investigated at temperatures below
the LCST by IH-NMR self- diffusion measurements. Echo decay data are fitted with
stretched exponential functions. The width of the distribution of diffusion coefficients does
not depend on concentration or temperature and thus can be attributed to the polydispersity
of the polymers. The mean diffusion coefficients increase with temperature, which is
mainly attributed to the decrease of the viscosity of water. As compensation for such
indirect effects of the temperature, the diffusion coefficients shown in Fig. 2 are
normalised on the D 20 viscosity and temperature, since it is D = kTI(61rTJRHJ.

1.6

1.4
fin
<> Fig. 2: Diffusion
1.2 <>
~
't coefficients normalised on water

*.•
.
1.0
66
b viscosity and temperature for (!)
~ 6 6 6 6 0 0 (open symbols) and ill (solid
•• ~~
0.8

o ~ •• i ••
symbols) in 0 20 at different

0.6

<> <> <> polymer concentrations: ~ A: 3


• •••
<>
• • 0:
0.4
mg/ml; D, e : 10 mg/ml; U,
0.2
30 mg/m!.
0.0
18 20 22 24 26 28 30 32 34
T('C]

A decrease of the diffusion coefficient with increasing concentration is observed for both
polymers. This decrease is expected due to the increasing hindrance by surrounding coils.
An estimate of the hydrodynamic radius from the diffusion coefficient data at 3 mg/mI,
where the effect of restricted diffusion is negligible, results in RH for (1) and (~ of 9 nm
and 11 nm, and further in an overlap concentration of c* ::::: 50 mg/ml and 35 mg/ml,
respectively. No significant change of the normalised diffusion coefficient D· 'lIT is
observed with increasing temperature below the phase transition, only a slight increase
occurrs around 30°C. There is thus no evidence of a substantial rearrangement or
aggregation of the coils in solution below the phase transition.
The data points at the phase transition, i.e. at 32°C, which show drastically increased
values of D'l1/T are difficult to interpret in terms of a structural model, since at the phase
transition solid particles are formed, and the signal arises from the liquid fraction of the
spins only. The diffusion coefficients at this temperature might be dominated by mobile
segments attached to already partly collapsed polymer chains, or might reflect the
polydispersity, assuming that shorter chains would collapse at a slightly higher
temperature.
The concentration effect of the coil-to-globule transition has been discussed in previous
work: At extremely dilute polymer concentrations, individual chains undergo a transition
to a globule [4], while in the region ofc::::: mg/ml the globule size depends on the PNIPAM
588

concentration [5], and at high concentrations above the overlap it is discussed that the
transition occurs via chain aggregation and subsequent collapse [2].
Our diffusion results prove that at concentrations c < c*, neither drastic changes of the coil
conformation nor aggregation of the chains occur below the phase transition temperature.
Only a slight increase of the diffusion coefficient is found about 2°C below the LCST,
which points at a minor rearrangement of the chains, probably similar to the formation of
'crumpled coils ', as discussed by Wu and Wang [4] for the extremely dilute regime. This
holds for the PNIPAM homopolymer as well as for the copolymer, though the increase of
the diffusion coefficient is less pronounced for the copolymer, possibly the formation of
'crumpled coils' is slightly hindered by electrostatic repulsion.

3.2. PHASE TRANSITION OF PNIP AM POLYMERS IN SOLUTION

Liquid IH spectra in D 20 solution are measured in dependence of concentration and


temperature. Evaluating the integrated area of the strongest resonance, the methyl protons
of the NIPAM group, the amount of liquid spins is followed with increasing temperature.
The NMR signal is constant at low T, with a sharp decrease to zero at 32° C, due to the
formation of solid globular particles, where spin relaxation is too fast to allow detection in
the liquid type experiment. The sharp transition of segmental mobility is consistent with
previous structural investigations of PNIPAM [I] . NMR spectra of the copolymer show
exactly the same properties, demonstrating that the presence of the charged co-monomer at
this amount does not influence the LCST behaviour. This is consistent with turbidity and
DSC measurements of this polymer at low pH values [2]. Above the overlap
concentration, however, a decrease of the IH liquid signal occurs already below the phase
transition, and the transition is broadened. Again, the phase transition properties of the
copolymer at these conditions are identical to those of the homopolymer. Intensity data are
shown and discussed in detail elsewhere [3].

3.3. PHASE TRANSITION IN ADSORPTION LAYERS

In adsorption layers, the methyl resonance is substantially broadened as compared with the
polymer in solution. The increased spectral width can be attributed to homogeneous
broadening arising from an increased relaxation rate and thus a reduced mobility of the
segments observed.
The polymer signal decreases in magnitude with increasing temperature. Fig. 3 shows the
signal of the compounds (1) and (~) in dependence of temperature for samples of different
mixing ratio of polymer and silica. Similar to the solution data, the signal decreases with
T, indicating a phase transition. The transition is significantly broadened compared with
that of free polymers in solution, where the width was < 2K. The broadening is more
strongly pronounced at a lower surface coverage. It can be concluded that the broader
transition is arising from polymer segments which are located closeer to the silica surface,
and exhibit a very low mobility. At higher initial concentrations of polymer the transition
is dominated by a fraction of polymer segments further away from the interface, which can
be assumed to be more mobile. These segments exhibit a sharper transition. The data thus
support a picture of the width of the phase transition depending on the distance of the
segments from the surface.
589

Fig. 3: Liquid state 1H signal


(integral of the methyl proton
resonance) of polymers adsorbed
to colloidal silica in dependence
of temperature: open symbols:
ill, and solid symbols @ at
different polymer/silica initial
weight ratios: 0, .: 0.85; 0,
.A.: 0.25. The lines are fits to the
data as described in (6).

At a polymer/silica ratio below 0.2 no liquid signal is observed, and it can be concluded
that the transition at low coverage in Fig. 3 reflects the behaviour of tails and loops in the
adsorbed layer, while polymer segments directly interacting with the surface ('trains'), are
too rigidly bound to show a liquid signal. In T 2 experiments performed with a solid echo,
indeed a biexponential decay with a fast component of T2 - 40-50 J.ls is found, which is
decaying too fast to be detected in liquid spectra. This corresponds to the general picture of
a 'solid' type dynamics of trains, and 'liquid' type dynamics of tails and loops in swollen
adsorption layers of polymers [7].
Though the phase transition properties of the copolymer in solution are the same as those
of (1), distinct differences are observed in adsorption layers, since the dependence on
adsorbed amount is far more pronounced for the copolymer layers: At low polymer
coverage the phase transition is not only further broadened as compared with (1), but the
liquid signal seems to reach a plateau value at high temperatures, which is larger than zero,
indicating that not all polymer segments are taking part in the phase transition. The fit
results of the curves in Fig. 3 confirm the finding of a decrease of the width of the phase
transition with increasing polymer coverage for both polymers, and a nonzero plateau at
high T for (l) only, which is largest at low coverage [6].
This behaviour of the copolymer is interesting, since on the one hand the broadened phase
transition is an indication for strongly immobilised polymer chains close to the surface. On
the other hand, the presence of a liquid signal proves the existence of rather mobile
segments at high temperatures. This apparent contradiction can however be understood by
considering the electrostatic repulsion of the copolymer from the negatively charged
surface, and the implications for the mobility on different length scales: The copolymer is
probably directly adsorbed with NIPAM monomers only, while the carboxylic groups are
repelled from the surface. This repulsion might immobilise the polymer configuration on
the scale of several segments and thus hinder the rearrangements necessary for a collapse,
so that a certain mobility is preserved on the segmental scale, resulting in a liquid signal.
This effect occurs predominantly at low coverage. At high coverage, obviously segments,
which are repelled from the surface, can collapse, since they have access to a sufficient
number of polymer segments in the vicinity, due to the presence of excess polymer. Only
590

at low coverage, these rearrangements can not take place, so that a larger absolute number
of segments stays mobile above the phase transition.

4. Conclusions

IH spectra qualitatively monitor the mobility of polymer segments and thus the coil-to-
globule transition. The charged copolymer shows a sharp phase transition in dilute solution
at about 32°C, identical to the homopolymer of PNIPAM. Results from IH spectra and
PFG-NMR diffusion agree in the finding that there is no conformational change of either
polymer below the phase transition in solutions below and around the overlap
concentration. A slight increase of the diffusion coefficient is evident some degrees below
the LCST.
Both polymers bind to the silica particles. With lH NMR, segments not directly bound to
the surface (tails, loops) are monitored. The phase transitions at the interface are much
broader than in solution, the width probably reflecting the polymer mobility gradient in
dependence of the distance from the surface: Polymers close to the surface show a broader
phase transition than polymers at a larger distance from the surface. While PNIPAM
undergoes a complete transition to a solid state, the phase transition of the copolymer is
hindered by electrostatic repulsion from the surface. This effect is most pronounced at low
coverage.

5. Acknowledgements

A. L. was supported by a research grant, partly from the Royal Society of Arts and
Sciences (Goteborg) and partly by the Max Planck Society.

6. References
I. Schild, H.G. (1992) Poly(N-isopropylacrylamide) : Experiment, Theory and Application. Prog. Polym. Sci.
17, 163-249.
2. Kuckling, D., Adler, H.-J., Arndt, K.-F., Ling, L. and Habicher, W. D. (2000) Temperature and pH
dependent solubility of novel poly(N-isopropylacrylamide) copolymers, Macromol. Chem. Phys. 201,273-
280.
3. Larsson, A., Kuckling, D., Schonhoff, M. (2001) IH NMR of Thermoreversible Polymers in Solution and at
Interfaces: The Influence of Charged Groups on the Phase Transition, Colloids and Surfaces A :
Physicochemical and Engineering Aspects 190(1-2), 175-182.
4. Wu, C. and Wang, X. (1998) Globule-to-Coil Transition ofa Single Homopolymer Chain in Solution, Phys.
Rev. Lett. 80, 4092-4094.
5. Chan, K., Pelton, R. and Zhang, J. (1999) On the Formation of Colloidally Dispersed Phase-Separated
Poly(N-isopropylacrylamide) Langmuir 154018-4020.
6. Schon hoff, M., Larsson, A., Welzel P., Kuckling, D. The Phase Transition of Thermoreversible Polymers
Adsorbed to Colloidal Silica: A I H NMR and DSC Study, 1. Phys. Chem. B, submitted.
7. Barnett, K. G., Cosgrove, T., Vincent, 8., Sissons, D. S. and Cohen-Stuart, M. (1981) Measurement of the
Polymer-bound Fraction at the solid-liquid Interface by pulsed NMR, Macromolecules 14, 1018-1020.
THE DILUTE TO SEMI-DILUTE TRANSITION IN PETROLEUM COLLOIDS,
AS STUDIED BY lH NMR RELAXOMETRY AND VISCOSIMETRY

D. MASTROFINI, L. BARRE, S. GAUTIER, E. RICHARD


Institut Fram;ais du Petrole,
I et 4, Avenue de Bois-Preau, 92852 Rueil-Malmaison Cedex, France

The present paper investigates Safaniya asphaltenes in perdeuterated toluene solutions


at different concentrations (in between 1 and 13% w/w), over a range of temperatures,
by means of 'H NMR relaxometry (measure of proton spin-spin relaxation time T2)'
The transition from the dilute to the semi-dilute region is studied through the
observations of proton spin-spin relaxation time (T2)' The experimental T2 decay curves
are either mono or hi-exponential; this behaviour is a function of asphaltene
concentration and temperature. The resolution of the T 2 decay into two components is
here attributed to the occurrence of the dilute to semi-dilute domain transition. Hence,
as for polymeric solutions, an overlap concentration c* is determined.
Observations performed with NMR relaxometry are compared and related to
viscosimetry and small angle neutron scattering (SANS) measurements.
It appears that T2 'H-NMR measurements complement the mechanical results giving
more precise information on the temperature dependence of c*.

1. Introduction

Crude oils are colloidal systems whose disperse phase is mainly composed of
asphaltenes [1-2]. Asphaltenes are defined as a solubility class, typically the n-pentane
or n-heptane insoluble fraction of a crude oil. At molecular level, they consist of highly
condensed polyaromatic rings bearing aliphatic and naphthenic chains with hetero-
atoms (0, Nand S) as part of the aromatic system; aromatic sheets stack together by
1t-1t bonds [3] .
The behaviour of asphaltenes, both in solution and in their natural medium, is governed
by association phenomena. Small angle X-ray and neutron scattering measurements
show that asphaltenes self-associate forming highly polydisperse aggregates, with a
very open, mass fractal-like internal structure.
More precisely, an apparent fractal exponent d equal to 2.0 ± 0.2 was found for
asphaltenes in toluenelheptane mixtures [4], as typical for aggregates formed by
reaction-limited colloidal aggregation (RLCA) or percolation process.
As for the strong polydispersity of asphaltene aggregates in solution, the mass
distribution of asphaltene aggregates in toluene, at room temperature, was recently
measured by Fenistein et al [5] through ultracentrifugation. The mass distribution N(M)
591
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 591-596.
© 2002 Kluwer Academic Publishers.
592
follows a power-law equation N(M)ocM" with -c=1.66. The polydispersity index,
defined as the ratio of weight average (Mw) to number average (Mn) of the system, is
calculated to be ~ 3.87.
Changes in the average size of asphaltene aggregates take place in response to a change
of solvent and temperature (1,2,6,7]: the structure gradually evolves from dispersed
micelles into clustered micelles. Asphaltenes behave as individual colloidal aggregates
dispersed in a solvent in the 'dilute' domain. As asphaltene content in solution
augments, i.e. the 'semi-dilute' domain, inter-aggregate interactions get more and more
important and the colloidal aggregates overlap each other. The critical concentration, at
which the overlapping process begins to occur, is called, by analogy with polymeric
solutions, the 'overlap concentration' c*.
The modifications occurring in the colloidal structure of asphaltene aggregates, above
c*, strongly affect the rheological and physical properties of petroleum colloids, such
phenomenon having a significant impact on the production, transportation and refining
operations [1,2]. As far as asphaltene aggregates are concerned, such modifications take
place at relatively low asphaltene volume fraction, as typical for systems with large,
open aggregates. For smaller aggregates, i.e. hard spherical ones, changes are observed
at much higher concentration.
The possibility to determine the overlap concentration threshold and to establish how it
varies as a function of changes in temperature conditions is, therefore, an important key
in our understanding of the processability of refinery products.
The main goal of the present work is to highlight the potential use of 'H NMR
relaxometry, through the measurements of the spin-spin relaxation time T z, to study the
transition from the dilute to the semi-dilute domain in petroleum colloids and the
influence of temperature on the c* onset.
During the last decades, NMR relaxometry has been extensively applied to a variety of
dispersed systems, i.e. polymeric gels, microemulsions, colloidal suspensions, in order
to derive useful information about structural parameters [8-10]. Greater attention was
devoted to the spin-lattice relaxation time (T,) than to the spin-spin relaxation time (T z),
T z measurements being far more complex to interpret. A number of papers, recently
appeared, deal with the measurements of the spin-spin relaxation time T z of crude oils
and asphalts in relation with their viscosities [11-12].
In the present research, we have elucidated the onset of the overlap concentration c*, as
a function of temperature, for Safaniya asphaltene in perdeuterated toluene solutions at
different concentration through the observations of proton spin-spin relaxation time
(T z). A comparison among small angle neutron scattering (SANS), viscosimetry and
NMR relaxation observations is also given.

2. Experimental

Asphaltenes were obtained by preclpltation in an excess of n-heptane (IP 143/81


method) from the Safaniya vacuum residue. Asphaltene solutions were made by
dissolving the appropriate amount of asphaltene in perdeuterated toluene (Euriso-top)
and allowing the solutions to stand overnight.
593

Proton T2 experiments were performed on a Bruker Minispec PC-120 spectrometer at a


proton resonance frequency of 20 MHz. This spectrometer was equipped with a system
for varying the temperature from -20 to 100°C. Thirty minutes were allowed for the
temperature of the samples to equilibrate. The Carr-Purcell-Meiboom-Gill (CPMG)
sequence (RD-90ox-T-(l800 y -T-acq-T)n) was used to eliminate the effect of
inhomogeneity in the static magnetic field and reduce the contribution of diffusion to
relaxation signal. The experimental parameters used in the data collection, previously
optimised, were: Pulse Interval Time T = SOilS, Recovery Delay RD = 5s, Number of
Scans ENH2 = 100.
Viscosities were measured using a shear-rate controll~d Low-Shear Contraves 30
viscosimeter, at different temperatures (8, 20 and 40°C). In the range of shear rates
accessible, we observed a Newtonian behaviour for all the asphaltene solutions under
investigation.

3. Results and Discussion

3.1. IH NMR RELAXATION MEASUREMENTS

Figure 1 shows the spin-spin relaxation curves of several asphaltene in perdeuterated


toluene solutions over the range of concentrations (in between I and 13% w/w) and
temperatures (8, 20, 40 and 75°C) investigated. For each temperature, the ratio between
the magnetisation amplitude at a time t, M(t), and that at time zero, Mo, on a
logarithmic scale, is reported as a function of time.

:I :~:~~:::::""""""" '"
'"I. . .. : . . .
~
•••. '... . .
.••....
. '. : I:::

. ",

.J3",,,
.

!:~ rn
~~\6 f ",' .... .::::: ••. ".. !.~..! !
,x> ; • '" •••••• :::::::.: ••• ,
.... I

'.'
•• •

'" .1 .......-...... ::: .. . :.:::::."'~


" . ',.
i 'f*3P(
.. .
~~~~_ _
_:_._~~~.~~_..J
.~()

'L2 1
'-.-_ _ __ _ _

'" ~--" · r··~'"


::1:::~:~::: ..... ~:: . ~".
.:«~::~
R
i ." I
J .......::::::.: ........ :: ..... 1::::11
- I
I"l ... :"-::::::;".:." ..
'u
: '''' 1
I; ' ~ -!

" • • •• •• • • • • • • • , '1.l!_ 1 I :; "" .... :. .... . .


• • • • • • • • • .. • ' . L--, I : I' T-,.,( " • ."

.I·{liI T-4O'C

.,_
,L • ••••••• . .~I'--- _ _ _ __ - - -- - - --"

Figure 1. IH Tl relaxation decay for Safaniya asphaltene in toluene-D solutions over


a range of concentrations at 8°C, 20°C, 40°C and 75°C.
594

For each temperature, the relaxation signal can be resolved into one or two
exponentially decaying functions. For the sake of simplicity, only the fits relative to
mono-exponential decaying magnetisation curves are shown.
The resolution of the T2 signal into two components (a short spin-spin relaxation time,
T2s , and a long spin-spin relaxation time, T2d is here attributed to the occurrence of the
dilute to semi-dilute domain transition and to the modifications involving the colloidal
structure of the system at the overlap concentration, c*.
In the dilute domain, asphaltenes behave as individual, not interacting colloidal
aggregates dispersed in the solvent (perdeuterated toluene); the single exponential decay
reflecting the relaxation rate of the individual colloids. In this case, the transverse
relaxation time T 2, being related to the mobility of asphaltene aggregates, is therefore
closely related to the viscosity of the system.
Beyond the overlap concentration c*, the whole system may be drawn as follows . The
bigger asphaltene aggregates, which partially overlap each other, give rise to a network-
like structure; the smaller ones, 'isolated and non-interacting', are trapped in the pores
of that network.
As a consequence, the short spin-spin relaxation time T2s corresponds to the relaxation
of the bigger, overlapped aggregates, that is the fraction of the system more 'solid-like'
and therefore the fast-relaxing one. The long relaxation time T2L is, on the contrary, due
to the relaxation of the non-overlapped fraction of aggregates, the smallest ones, and
relates to the mobility of them in the pores of the network.
As typical for large, highly solvated polymeric aggregates, the transition from dilute to
semi-dilute regime occurs at relatively low concentration (in this case, lower than 4%
w/w) . The critical asphaltene concentration, at which the appearance of a bi-exponential
relaxation process occurs, increases with temperature. c* of 1.5-1.5-2.5 and 4.0% w/w
are found at 8, 20, 40 and 7SoC, respectively.

3.2. VISCOSITY MEASUREMENTS. COMPARISON WITH SANS OBSERVATIONS.

In Figure 2 the relative viscosity (l1R) against asphaltene concentration (CA) trend is
reported for Safaniya asphaltene in perdeuterated toluene solutions, at 8, 20 and 40°C.
In the dilute region, 11R varies linearly with concentration, according to Einstein law.
The lower the temperature, the higher the intrinsic viscosities [11], i.e. the higher the
hydrodynamic volume of aggregates in solution. [11] of 14.6, 12.1 and 8.60 gig are
found at 8, 20 and 40°C, respectively. By increasing concentration (semi-dilute region),
viscosity rises dramatically, providing evidence of a high degree of interactions among
aggregates. The asphaltene concentration at which strong deviations from linearity
firstly appear represents the overlap concentration [4, 13].
To distinguish that value determined by NMR relaxometry, the overlap concentration
measured by viscosimetry, is named c~ *. The values of c~ *, calculated at each
temperature, do not seem to depend on temperature. A value of 4-5% w/w is found
whatever the temperature under investigation is.
SANS measurements carried on by Roux et al [14] on Safaniya asphaltene in
perdeuterated toluene solutions at different temperatures are in line with viscosimetry
evidences. On the one hand, the decrement of experimental radius of gyration Rg,
observed for increasing temperature, reflects the increment of the hydrodynamic volume
595

20c

18
I ZT=8'C I
16 I· T=20·C ,
14 ~.T=4Q:C I •
12

,! 10
8
6

4

2 i
0
0 2 4 6 Ca. %~/w 10 12 14

Figure 2 Relative viscosity (1]0 as a function of asphaltene concentration (Ca) for Safaniya asphaltene in
toluene-D solutions at different temperatures.

of aggregates. On the other hand, the critical concentration obtained by SANS


experiments (c*sans) [13] is, as well as cll *, in between 4-5 %w/w, independently on
temperature.

4. Conclusions

This work proposes 'H NMR relaxometry as an alternative technique to viscosimetry


and scattering methods, to study the transition from the dilute to semi-dilute domain in
petroleum colloids, i.e. the overlap concentration c* and its dependence on temperature.
The variation of asphaltene average size following to a change of temperature, as radius
of gyration and intrinsic viscosity vs. temperature trends underline, validates the
hypothesis that a shift of c* towards lower concentrations has to occur as temperature
decreases.
However, as far as the determination of the overlap concentration is concerned, NMR
and viscosimetry (and SANS) measurements provide different results. c* augments as
temperature increases, according to NMR relaxation spectra, whereas c~ * and c* sans are
constant as a function of temperature.
The above-mentioned deviations in behaviour of the overlap threshold may be
explained when the different time scale, which NMR relaxometry and viscosimetry
study, is taken into account. Spin-spin relaxation time (T2) measurements involve much
shorter time scale than that observed by viscosity and, as such, when aggregates in
dilute solutions interact, few aggregates suffice to account for the appearance of a multi-
exponential 'H NMR relaxation decay. On the contrary, because steady state viscosity
(110) depends on weighted averages [15], the sharp increase of viscosity indicative of the
dilute to semi-dilute transition is observed at higher concentrations as the aggregates
begin to overlap appreciably.
The overlap concentrations measured by RMN will result, hence, far smaller compared
with those found by viscosimetry.
In conclusion, in our opinion, NMR measurements are more representative of the dilute
to semi-dilute transition in petroleum colloids than viscosimetry and scattering ones.
596
5. Acknowledgement

The authors thank Mr. J-F. Argillier and Ms. F. Brucy for the helpful discussions.

6. References

I. E.Y. Sheu, a.c. Mullins Editors (1995), Asphaltenes. Fundamentals and Applications, Plenum Press,
New York.
2. E.Y. Sheu, a.c. Mullins Editors (1998), Structures and Dynamics of Asphaltenes, Plenum Press, New
York.
3. T.F. Yen, J.G. Erdman, S.S. Pollack (1961) Investigation of the structure of petroleum asphaltenes by X-
ray diffraction. Anal. Chem. 33,1587-1594.
4. D. Fenistein, L. Barre, D. Broseta, D. Espinat, A. Livet, J-N. Roux, M. Scarsella (1998) Viscosimetric
and neutron scattering study of asphaltene aggregates in mixed tolueneiheptane solvents. Langmuir
14(5), 1013-1020.
5. D. Fenistein, L. Barre (2001) Experimental measurements of the mass distribution of petroleum
asphaltene aggregates using ultracentrifugation and small-angle X-ray scatterins, Fuel 80, 283-287.
6. P. Thiyagarajan, 1. Hunt, R. Winans, K.B. Anderson, J.T. Miller (1995) Temperature-dependant
structural changes of asphaltenes in I-methylnaphthalene. Energy & Fuels 9(5), 829-833.
7. E.B. Sirota, (1998) Swelling of Asphaltenes, Petroleum Sci. & Technology 16(3&4),415-431.
8. F. Duval, P. Porion, H. Van Damme (1999) Microscale and macroscale diffusion of water in colloidal
gels. A pulsed field gradient and NMR imaging investigation. J. Phys. Chem. B 103,5730-5735.
9. A. Dekmezian, D.E. Axelson, J.J. Dechter, B.Borah, L. Mandelkem (1985) Carbon-13 NMR relaxation
and transitions in polymers, J Polym. Sci. 23,367-385.
10. A. Charlesby, E. Jaroszkiewicz (1985) Entanglement and network formation in polystyrene. Viscoelatsic
behaviour from pulsed NMR, Eur.Polym. J. 21(1),55-64.
II . 1. Jacob, L.A. Davis (1999) The NMR T2 response of crude oils at elevated temperatures, SPE 56797.
12. R.S . Kashaev (2000) A Pulse nuclear magnetic resonance study of the structure of resinous-asphaltenic
components of crude oils, Petroleum Chemistry 40(4),236-240.
13. I. Henaut, L. Barre, J.F. Argiller, F. Brucy, R. Bouchard (2000) Rheological and structural properties of
heavy crude oils in relation with their asphaltene content, SPE 65020. .
14. J-N. Roux, D. Broseta, B. Deme, SANS Study of asphaltene aggregation in toluene: concentration and
solvent quality effect, submitted to publication.
15 . J.D. Ferry (1980) Viscoelastic properties ofpolymers,J. Wiley & Sons Inc. (3,d Ed.), New York.
SIV MAS NMR STUDIES OF INORGANIC VANADATES WITH VERY SMALL
AND LARGE CHEMICAL SHIFT ANISOTROPIES

ULLA G. NIELSEN, HANS 1. JAKOBSEN, AND J0RGEN SKIBSTED


Instrument Centre for Solid-State NMR Spectroscopy
Department of Chemistry, University of Aarhus
DK-8000 Aarhus C, Denmark

Abstract Magnitudes and relative orientation of SIV quadrupole coupling and very
smalIlIarge chemical shift anisotropy (CSA) tensors can be obtained with high precision
from high-field SIV MAS NMR spectra of the central and satellite transitions.
Employing SIV MAS and triple-quantum MAS NMR this is demonstrated for YV04 and
LaV04, which possess very small SIV CSAs, while tris(triphenylsilyl) vanadate
illustrates the method for a SIV site with a large SIV CSA. Relationships between
anisotropic SIV NMR parameters and structural data are discussed.

1. Introduction

Vanadium-based heterogeneous catalysts are used in a number of industrially important


chemically reactions including selective oxidation of hydrocarbons, selective reduction
of NO x, and oxidation of S02 in the production of sulphuric acid. Information about the
local structure of the vanadium species on the surface of these catalysts is crucial for
understanding and improving the catalytic properties. Solid-state SIV NMR is a useful
technique in studies of such materials because this method specifically probes the local
structure of the vanadium species on the surface [1]. Structural information is mainly
derived from the SIV parameters characterizing the quadrupole coupling interaction and
the chemical shift anisotropy (CSA). Although SIV MAS NMR spectra are generally
dominated by the quadrupole interaction, valuable information about the local geometry
of the VOx polyhedra can also be achieved from the SIV CSA [1-8]. Thus, methods for a
reliable determination of very small and large CSAs are important for the application of
SIV NMR in studies of different vanadium species. In this work we demonstrate that the
combination of multiple-quantum MAS NMR and single-pulse MAS NMR of the
central and satellite transitions [3,4] is particularly useful for determination of very
small SIV CSAs. This is illustrated by 5lV MAS and triple-quantum (3Q) MAS NMR
spectra of the orthovanadates YV04 and LaV04 • Furthermore, a SIV MAS NMR study
of tris(triphenylsilyl) vanadate, a model compound for the chemical binding of
vanadium to a Si02 surface [9], is presented and serves as the illustrative example of the
high precision in 5lV NMR data that can be achieved from MAS NMR of the central
and satellite transitions for vanadium species possessing large CSAs.
597
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 597-602.
© 2002 Kluwer Academic Publishers.
598

2. SIV MAS and triple-quantum MAS NMR ofYV04 and LaV04

The Sly MAS NMR spectrum of the central and satellite transitions for YY04 (Fig. 1)
exhibits a highly symmetric manifold of spinning sidebands (ssbs), demonstrating that
this spectrum is overwhelmingly dominated by the Sly quadrupole coupling interaction.
The strong quadrupole coupling also results in partly splittings of the individual ssbs
into separate resonances as a result of the difference in second-order quadrupolar shift
for the individual satellite transitions [3,4]. Least-squares fitting of simulated to
experimental ssb intensities, employing the same theoretical approach as recently used
in studies of other inorganic vanadates [6-8], gives the parameters CQ = 4.76 ± 0.02
MHz and 11Q = 0.02 ± 0.05, when only the quadrupole interaction is considered in the
optimization. The axially symmetric quadrupole coupling tensor (i.e., 11Q = 0) is in
accord with the four-fold inversion symmetry for the SlY site in the crystal structure of
YY0 4 [10]. This also implies that 110 = 0 and thus coincidence of the quadrupole
coupling and CSA tensors (i.e., \jI = X = ~ = 0°). Incorporation of the CSA in the
optimizations only gives a rough estimate of the numerical value for the shift anisotropy
parameter 00 I I : :;
12 ppm, indicating a very small CSA for Sly in YY04. To obtain a
more reliable determination of the sign and magnitude of the 00 parameter, YY0 4 is
investigated by Sly 3QMAS NMR (Fig. 2a), since this experiment magnifies the CSA
interaction by a factor of three in the isotropic dimension of the 2D spectrum. A high-
resolution spectrum in the isotropic dimension is reconstructed from the 3QMAS
spectrum, employing the approach by Wang et at. [11]. The reconstructed spectrum
(Fig. 2b) is clearly dominated by the SlY CSA interaction and least-squares analysis of
the ssbs in this spectrum gives the shift anisotropy 00 = -15 ± 3 ppm and the optimized
simulation shown in Fig. 2c. To our knowledge this is the smallest Sly CSA reported so
far for a vanadate. Employing the 00 value from 3QMAS NMR as a fixed parameter in
the optimization to the single-pUlse MAS spectrum (Fig. la) gives the values CQ = 4.76

,
92 88 (kHz) ·620 -660 (ppm) 92 -620 -660 (ppm)

(a) (xIO) (b) (xIO)

' i' "1''' i I j'" iii i i Pi I


50 0 300 100 -1 00 - 300 - 500 (kHz) 500
Figure 1. (a) Sly MAS NMR spectrum of YY04 recorded at 14.1 T using a spinning speed v, = 6.0 kHz.
(b) Optimized simulation corresponding to the Sly parameters given in the text. The left-hand insets illustrate
the line shapes of the individual ssbs. The spectral region for the central transition is shown in the right-hand
insets on a ppm scale relative to neat YOeI 3.
599

F2
(ppm) (a)

-690
-680
(b)
-670
~ ~,~, ~
-660
-650 *
-640
-630 ~
(c)
-620-
I I I I
I I
I I I
I I

120 140 160 Fl (ppm) 135 145 155 165 (ppm)


Figure 2. (a) Contour plot of the Sly 3QMAS NMR spectrum of YY04 (14.1 T, v, = 5.0 kHz) recorded using
the Z-filter three-pulse sequence, a 80 kHz spectral width in both dimensions, 256 II-increments, an rf field
strength of rB/2rr. =90 kHz,'a relaxation delay of 38 s, and 24 scans. (b) Reconstructed summation of the
isotropic dimension. (c) Optimized simulation of the spectrum in (b), considering the scaled CSA interaction
for a 3QMAS experiment and corresponding to the parameters 00 =-IS ppm and 110 =0.00.

± 0.02 MHz and 0iSO =-662.1 ± 1.0 ppm for Ocr = -15 ± 3 ppm, 11Q = 11cr =0 and 'I' =X =
S=0° for YV04 , which should be considered the optimum data for YV04 resulting from
the combined analysis of the 3QMAS and MAS NMR spectra. These parameters are
employed in the optimized simulation shown in Fig. lb. The CQ, 11Q' and 0iSO parameters
reported earlier for YV04 [1] are in good agreement with the 5lV NMR data determined
in this work.
The 5lV MAS NMR spectrum of LaV04 (Fig. 3a) exhibits the characteristic features
for orthovanadates, i.e., a fairly strong quadrupole coupling and a small CSA. Least-
squares fitting of simulated to experimental ssbs for the spectrum in Fig. 3a gives a
straightforward determination of the quadrupole coupling parameters (CQ = 5.20 MHz
and 11Q = 0.68). However, incorporation of the CSA and the relative orientation of the
two tensors results in an ambiguous determination of the Ocr parameter, since two
minima for the rms function are observed which correspond to Ocr ~ ±45 ppm. As
illustrated above for YV04 , this ambiguity can be solved employing 5lV 3QMAS NMR
which gives the parameters Ocr = -50 ± 5 ppm and 11cr = 0.71 ± 0.05 [12]. These
parameters are subsequently used as fixed parameters in an optimization to the MAS
spectrum (Fig. 3a), which gives the data CQ= 5.20 ± 0.02 MHz, 11Q = 0.68 ± 0.02, 'I' =
165° ± 23°, X =77° ± 6°, S = 36° ± 8°, and 0iSO = -604.7 ± 1.0 ppm for LaV04 and the
optimized simulation shown in Fig. 3b. Lapina et al. [1] have earlier determined the
CSA parameters for LaV04 (OiSO =-609 ± 10 ppm, Ocr =54 ± 10 ppm, 11cr =0.80 ± 0.1)
from analysis of the central transition at 9.4 T and reported the quadrupole coupling
parameters (CQ = 5.13 MHz, 11Q = 0.69) from another study. The CSA data are in
favourable agreement with the data determined in this work when 0iSO is corrected for
the second-order quadrupolar shift of the central transition (7 ppm). However, the data
obtained from the spectrum in Fig. 3 are expected to be of higher precision since both
interactions are considered in the least-squares analysis of the experimental spectrum.
600

o -200 -400 -600 (k Hz) 800 600 400 200 -200 -400 -600 (kHz)

Figure 3. (a) Experimental 'Iy MAS NMR spectrum of LaY04 (14.1 T, vr = 9.0 kHz). (b) Optimized
simulation corresponding to the Sly parameters given in the text. The left-hand insets illustrate that the
individual ssbs are partly split into separate resonances as a result of the difference in second-order
quadrupolar shifts for the satellite transitions whereas the spectral region for the central transition is shown in
the right -hand insets. The centerband is indicated by an asterisk.

3. 51V MAS NMR of Tris(triphenylsilyl) Vanadate

The coordination of vanadium in tris(triphenylsilyl) vanadate «<P3SiOhVO) includes


one V=O and three V-(OSi) bonds [9]. Thus, (<i>3SiO)3VO may be a suitable model
compound for the chemical binding of vanadium to the surface of a Si02 support.
(<i>3SiO)3VO has been synthesized by transesterification of triphenylacetoxysilane
(prepared from triphenylchlorosilane and acetic anhydride) and tris(isopropyl) vanadate
using the method described by Sekiguchi et al. [13].
The highly asymmetric ssb manifold, observed in the 51V MAS NMR spectrum of the
central and satellite transitions for (<i>3SiO)3 VO (Fig. 4a), is a characteristic feature for a
vanadium nucleus influenced by a large CSA. An estimate of the quadrupole interaction
CQ ", 3 MHz is obtained from the width of the ssb manifold (1.3 MHz). Optimization of
simulated to integrated ssb intensities and subsequent error analysis result in the
parameters CQ = 3.19 ± O.oI MHz, llQ.= 0.13 ± 0.02,0;50 = -731.4 ± 0.5 ppm, O(f =
420 ± 2 ppm, 11(f= 0.08 ± 0.08, X = 6 ± 2 °, and; = 16° ± 16° for 51V in (<i>3SiOhVO.
The error analysis reveals that the ssb intensities are insensitive to variations in 'II which
is ascribed to the fact that llQ '" 0, i.e., theory predicts that", is undefined for llQ = 0 [3] .
The 51V chemical shift parameters have previously been determined from 51V static and
MAS NMR of the central transition at 7.1 T by N. Das et al. [9] (o;so = -736 ± 1 ppm,
O(f = 414 ± 10 and, 11(f = 0.04). These data agree well with the parameters determined
from Fig. 4, when the value for o;so in their work is corrected for the second-order
quadrupole shift of the central transition (4 ppm at 7.1 T). The results indicate that
vanadyl species bonded to the surface of a Si02 support with three V-O bonds possess
large CSAs (O(ftV) '" 400 ppm) and approximately axial symmetry of the 51V
environments (11(f' llQ '" 0).
601

lJ
-680
jIII" I
iii iii i
·730 (ppm)
I
·680
If 111 I I Iii i i i i I
·730 (ppm)

(a) (b)

1""1 " "1""1""1""1""1""1""1""1" " 11""l""l""l""l""l""j""j""j""j""j


400 200 0 ·200 (k Hz) 400 200 0 ·200 (ki lL)

Figure 4. (a) Experimental Sly MAS NMR spectrum of the complete manifold of ssbs from the central and
satellite transitions for (<P3SiOhYO (l4.l T, v, = 8.0 kHz). (b) Simulation of the spectrum in (a) using the
optimized parameters given in the text. The kHz scale is relative to the isotropic peak whereas the ppm scale
is referenced to neat YOCI 3.

4. Inorganic Vanadates

During the past decade a large number of crystalline inorganic vanadates have been
characterized using either SIV static-powder and/or MAS NMR of the central transition
[1,2,5,9] or by MAS NMR of the complete manifold of spinning sidebands (ssbs)
observed from the central and satellite transitions [3,4,6-8]. The latter method has the
advantage that the SIV parameters for the quadrupole coupling, the CSA, and the
relative orientation of the two tensorial interactions (the Euler angles '1', X, and ~) may
be taken into account in the analysis of the experimental spectra, thereby providing
parameters of higher precision as compared to the results from analysis of the central
transition alone. From the reported SIV NMR parameters a number of relationships
between these data and structural parameters have been observed. For example,
isostructural vanadates are observed to possess very similar SIV CSA parameters and
Euler angles as well as a systematic variation in the magnitude of the quadrupole
interaction. For inorganic vanadates with vanadium in tetrahedral coordination,
including isolated (QO), dimers (QI) or chains (Q2) of V04 polyhedra, the SIV NMR CSA
parameters are observed to reflect the condensation of the V0 43. anions in the following
manner: 0 $ Iocr I $ 100 ppm for QO sites, 60 $ Iocr I $ 200 ppm for QI sites and 190 $
Ocr $ 320 ppm for Q2 sites. Moreover, the eclipsed and staggered conformation of the
pyrovanadate anion can be distinguished by the sign of the SIV CSA, since the eclipsed
conformation (LVOV < 140°) have 00 > 0 whereas pyrovanadates with the staggered
V 20l conformation (LVOV > 140°) have Ocr < 0 [8]. For the metavanadates with 5lV in
octahedral coordination and the pyrovanadates with pentacoordinated SIV sites one of
the V-0 bonds are significantly longer (> 2 A) than the remaining V-0 bonds, which
strongly affect the appearance of the SIV MAS NMR spectra. Generally, these
vanadates possess some of the largest quadrupole coupling and CSA values observed
for inorganic vanadates, i.e., CQ ~ 6.5 MHz and Ocr ~ 260 ppm. The pentacoordinated
metavanadates have similar CSAs (ocr ~ 300 ppm) but weaker quadrupole couplings
(1 $ CQ $ 4 MHz).
602

5. Conclusion

Magnitudes and relative orientation of 51V quadrupole coupling and CSA tensors can be
obtained with high precision from 51V MAS NMR spectra of the manifold of spinning
sidebands observed for the central and satellite transitions. For very small CSAs the
sign and magnitude of the shift anisotropy parameter (00 ) can unambiguously be
determined from 51V 3QMAS NMR at a high magnetic field. Thus. the combination of
the two experiments may lead to improved precision for the anisotropic 51V NMR
parameters as demonstrated in the present work for YV04 and LaV04 •

6. References

I. Lapina, O. B. Mastikhin, V. M. Shubin, A. A. Krasilnikov, V. N. and Zamaraev, K. I. (1992) SIV Solid


State NMR Studies of Vanadia Based Catalysts, Progr. NMR Spectr. 24,457-525.
2. Eckert, H. and Wachs, I. E. (1989) Solid-State 51V NMR Structural Studies on Supported Vanadium(V)
Oxide Catalysts: Vanadium Oxide Surface Layers on Alumina and Titania Supports, l. Phys. Chern. 93,
6796-6805.
3. Skibsted, 1. Nielsen, N. C. Bilds0e, H. and Jakobsen, H. J. (1992) 51V MAS NMR spectroscopy:
determination of quadrupole and anisotropic shielding tensors, including the relative orientation of their
principal-axis systems, Chem Phys. Lett. 188,405-412.
4. Skibsted,1. Nielsen, N. C. Bilds¢e, H. and Jakobsen, H. 1. (1993) Magnitudes and Relative Orientation
of slV Quadrupole Coupling and Anisotropic Shielding Tensors in Metavanadates and KVPs from 51V
MAS NMR Spectra. 23Na Quadrupole Coupling Parameters for u- and ~-NaV03' l. Arn. Chern. Soc.
115,7351-7362.
5. Hayakawa, S. Yoko, T. and Sakka, S. (1994) slV NMR Studies of Crystalline Monovalent and Divalent
Metal Metavanadates, l. Solid State Chern. 112,329-339.
6. Skibsted, 1. Jacobsen, C. J. H. and Jakobsen; H. J. (1998) SIV Chemical Shielding and Quadrupole
Coupling in Ortho- and Metavanadates from slV MAS NMR Spectroscopy, [norg. Chern. 37, 3083-
3092.
7. Nielsen, V.G . Jakobsen, H. J. and Skibsted, J. (2000) Characterization of Divalent Metal Metavanadates
by 51V Magic-Angle Spinning NMR Spectroscopy of the Central and Satellite Transitions, [norg. Chern.
39,2135-2145.
8. Nielsen, V . G. Jakobsen, H. 1. and Skibsted, J. (2001) SIV MAS NMR Investigation of Sly Quadrupole
Coupling and Chemical Shift Anisotropy in Divalent Metal Pyrovanadates l. Phys. Chern. BIOS, 420-
429.
9. Das, N. Eckert, H. Hu, H. Wachs, I. E. Walzer, 1. F. and Feher, F. 1. (1993) Bonding States of Surface
Yanadium(Y) Oxide Phases on Silica: Structural Characterization by SIV NMR and Raman Spectro-
scopy, l. Phys. Chern. 97,8240-8243.
10. Lohmiiller, Von G. Schmidt, G. Deppisch, B. Gramlich, V. and Scheringer, C. (1973) Die
Kristallstrukturen von Yttrium-Vanadat, Lutetium-Phosphat und Lutetium-Arsenat, Acta Cryst. B29,
141-142.
II. Wang, S. H. Xu, Z. Baltisberger, J. H. Bull, L. M. Stebbins, 1. F. and Pines A. (1997) MuItiple-quantum
magic-angle spinning and dynamic-angle spinning NMR spectroscopy of quadrupolar nuclei. Solid State
Nucl. Magn. Reson. 8,1-16.
12. Nielsen, V. G. Jakobsen, H. J. and Skibsted, 1. (2001) Small SIV Chemical Shift Anisotropy for LaV04
from MQMAS and MAS NMR Spectroscopy (submittedfor publication).
13. Sekiguchi, S. and Kurihara, A. (1969) A New Synthesis of Tris(tripheny1sily1) Vanadate, Bull. Chern.
Soc. lpn. 42,1453-1454.
IH, 13C, and J29Xe NMR STUDY OF CHANGING PORE SIZE AND
TORTUOSITY DURING DEACTIVATION AND DECO KING OF A NAPHTHA
REFORMING CATALYST

X. -H. Ren, M. Bertmer, H. Kuhn, S. Stapf, D. E. Demeo,


B. Blumich, C. Kern, A. Jess
Institut fUr Technische Chemie und Makromolekulare Chemie, Rheinisch-
Westflilische Technische Hochschule, Worringerweg }, D-52056 Aachen,
Germany

Abstract

Proton, J3C, and J29Xe NMR were applied for characterizing the change of the
tortuosity, the chemical structure of the coke, as well as the pore size during the
deactivation and decoking of a commercial naphtha reforming catalyst (PtlRe-AhOJ).
All experimental evidence indicates that a full recovery of the activity of the clean
catalyst is not achieved by the regeneration process, and that the quality of regeneration
depends on the coke content reached during the deactivation/regeneration cycle.

1. Introduction

Catalytic hydrocarbon conversion processes are of great industrial relevance in the


refinery and petrochemical industry, e.g. the catalytic reforming of naphtha to convert
low-octane gasoline to high-octane gasoline, and to produce benzene, toluene and
xylenes. Although the deactivation of these reforming catalysts has been already
intensively studied and different approaches have been developed to model the
deactivation of the catalyst by coke formation, there is still no generally accepted model
for coke formation. Moreover, the regeneration of the catalysts has been less studied
than deactivation, and the understanding of several aspects of fundamental mechanisms
remains poor or inconsistent, which has hindered the accurate simulation and
optimization for both procedures. Furthermore, the reaction is so complex that the
results from one reaction system (e.g. naphtha reforming) are not directly applicable to
another. Thus it is essential to characterize the structure of these catalysts, and to follow
the dynamics of diffusion and chemical reactions during catalytic processes.
Among the techniques that can be used for the characterization of catalysts, NMR
is particularly useful because it provides information about the physical structure and the
chemical nature of the coke formed without the need to extract the coke or dissolve the
framework of catalysts [1-3]. The main aim of this work is to investigate pore space
morphology or tortuosity changes and pore size distributions in the processes of
603
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 603-608.
© 2002 Kluwer Academic Publishers.
604

deactivation and regeneration using multinuclear NMR techniques, thereby taking a


commercial naphtha catalyst (PtlRe-Ah03) as a model system.

2. Experimental

2.1. CATALYST AND CHEMICAL REACTIONS

The commercial refonning naphtha catalyst was purchased from Engelhard Company
(USA). The catalyst has a cylindrical geometry (1.5-2.0 mm x 5.0-8.0 mm). It consists
of the catalytic active species (for hydrogenation and dehydrogenation) Pt and Re,
supported on A120 3, which also acts as an acid component, catalysing isomerization and
cyclization. The catalyst is mesoporous, with pore diameters ranging from 4 to 10 run.
Coke was deposited in the catalyst by passing toluene or n-heptane (as a model
hydrocarbon) over a fixed bed at a total pressure of 6 or lObar and a constant
temperature of 540°C. Decoking (regeneration) experiments of the catalyst were carried
out in a home-built stainless steel tubular flow reactor at a total pressure of 2 bar and
temperatures up to 435°C. The products were analyzed by means of on-line analyzers
(CO, CO 2, O 2), the coke content was calculated by planimetry from the gas analysis
(mainly CO2; CO < 5 ppm) and the total volume rate of the off-gas.

2.2. NMR EXPERIMENTS

Liquid n-heptane was selected as a test substance for NMR diffusion measurements.
Prior to the experiments, the samples were saturated with n-heptane for at least 12 hours,
and after removing the excess liquid from the external surface, the cylindrical catalyst
pellets, completely filling with liquid n-heptane, were placed with their long axes
parallel to the axis of the 5 mm NMR tube aligned with the direction of magnetic field
Bo for detennination of self-diffusion properties in dependence of the spatial orientation
of the pellets. The experiments were performed at 7.05 T on a Bruker DMX
spectrometer, working at a proton resonance frequency of 300 MHz.
The IH longitudinal (TI ) and transverse (T2) relaxation times of n-heptane in
catalysts were detennined employing inversion recovery and CPMG pulse sequences [4-
6], respectively. The self-diffusion coefficients of bulk n-heptane molecules and n-
heptane confined in coked, decoked as well as fresh catalysts were measured on IH
using a pulsed gradient stimulated echo (PGSTE) sequence with a crusher gradient
inserting after the second 900 pulse in order to destroy residual coherence in the
transverse plane. The diffusion encoding time Ll was fixed at typically 266 rns, and 128
signals were accumulated for each step. The self-diffusion was shown to be isotropic in
catalyst pellets. In addition, a 13-interval alternating APGSTE sequence [7] with bipolar
gradients was used in order to compensate the effect of the internal gradients induced by
the heterogeneity of the medium.
IH and I3C solid state NMR spectra of the same catalyst samples were obtained on
a Bruker DSX 500 spectrometer with a field strength of 11.7 T using magic-angle
spinning (MAS). For the IH spectra a spinning frequency of 20 KHz was used with a 2.5
mm (ID) rotor to eliminate sidebands, and 256 accumulations with a recycle delay of 5 s
were required. Spectra of I3C at natural abundance were measured using cross-
605
polarization (CP) from protons at a spinning frequency of 5 kHz employing a 7 mm (ID)
MAS rotor in order to increase the signal-to-noise ratio, resulting in a strong spinning
sideband pattern. Typically, 15000 scans were accumulated.
129Xe NMR-measurements under a Xe pressure of 20 bars were carried out in a
home-build high-pressure 10 mm NMR tube. The tube consisting of white sapphire was
sealed by a non-magnetic Ti alloy valve suitable for introducing gases under pressure
(max. 30 bars). The spectra were obtained on a Bruker AC 300 spectrometer using a
pulse width of 5 J,lS and accumulating 4000 scans with a recycle delay of lOs.

3. Results and discussion

3.1. THE EFFECT OF COKE FORMATION ON CATALYST TORTUOSITY

The decrease in measured IH TI and T2 versus coke content for both processes shows
that the reorientation of adsorbed n-heptane molecules is severely restricted by the
deposition of the coke, but both relaxation times of n-heptane in the fully regenerated
samples are different to those in the fresh catalyst, which implicates a different structure
of the catalyst in both states. The difference in TI between fresh catalyst and fully
regenerated catalyst is considerately greater than that in T2• One possible explanation for
this fact could be in the different contributions of relaxation processes to the total
relaxation rate. TI is most sensitive to molecular reorientation mechanisms on the order
of the Larmor frequency ~. The transverse relaxation rate h on the other hand, is
dominated by slow processes and is strongly affected by the presence of paramagnetic
impurities such as metal ions.
The self-diffusion coefficient during the observation time Ll is given by the
attenuation of the signal intensity as a function of the wave vector q [4-6]:
D =_-_1_ 8InE(q) , (I)
Lf- 0/3 8(q2)
where q=;go , and E is the echo amplitude. This relation is exact if the probability
distribution of displacements is purely Gaussian, as in the case of unrestricted self-
diffusion. For deviation from a Gaussian shape, the low-q behavior, corresponding to
the slope of the initial part of the decay, is proportional to an effective diffusion
coefficient connected with the averaged rrns displacements.
For a long diffusion encoding time L1, the displacement of spins becomes
independent of L1 and depends only on the pore morphology of the matrix, and the self-
diffusion coefficient D(Ll) reaches an asymptotic limit defined by the Einstein relation:
lim D(t1) =.!. . (2)
A...... Do r
where r is the tortuosity of the porous medium, which is a characteristic of the pore
structure and describes the long range connectivity of the medium. The tortuosity is
often used to estimate the fluid permeability in pores media and is an important
parameter for the simulation of reactions in the catalysts because it determines the
transport of reactants to and of the products from the catalytic surface.
In order to estimate the effect of internal magnetic gradient fields, the I3-internal
APGSTE sequence [7], which can cancel or reduce the cross terms, was compared to the
606

conventional PGSTE sequence in measuring the diffusivity of the n-heptane molecules


in the fresh catalyst. As shown in Fig. 1, the difference between both echo decay curves
is within the error limits, which indicates that either the effect of the internal gradients
are sufficiently small, or that due to the fast diffusion in the pores their effect is averaged
out. It can be concluded that the PGSTE method, which produces a much better signal-
to-noise-ratio than the 13-internal-sequence due to the relaxation properties of the
samples, leads to sufficiently reliable results. Moreover, the linear behavior of the
l
normalized echo amplitude versus (L1 - o/~ indicates a Gaussian distribution of
displacements. This has been tested by recording E(q)/E(O) for different delays L1 in the
range between 100 ms and 300 ms. Time-independent diffusion was obtained constants
within error margins, so that the validity of eq.(2) can be assumed.
1".' .... 3.6P:::==t===t==±=;+~
'
0
.. c
3.2 H---aG-*--t---+*""""--1-4

I
<I
I '§ 2.1 rv'-----t--t-?"'y ' l ---t---I
! 0
I ~
0
I 0 2.4 r + - - +- - t - - - + - ---l--4
I I i o I
o 1 2 3 4 I 12 16
192rl(A-a/3) (x10' m·2s) coke content (wt%)
Fig.1 Measurements on diffusivity of n-heptane in Fig.2 Change of the tortuosity during the
fresh catalyst by the PGSTE sequence ( + ) and the deactivation ( • ) and regeneration procedures
I3-interval PGSTE (APGSTE) sequence ( 0). The with initial coke content of 16 wt% (. ) and 6
diffusion encoding time is 266 ms. wt% (_).
The change of the self-diffusion coefficient of n-heptane molecules in coked and
decoked samples shows a similar tendency as the magnetization relaxation times. This
indicates a strong correlation between the surface area and the pore volume as could be
expected. From these self-diffusion coefficients, the variation of tortuosity during
deactivation and regeneration process is obtained from Eq.(2) and shown in Fig. 2. An
increase of the coke content is accompanied by an increase of the tortuosity due to a
reduction of the accessible pore space. During regeneration an intermediate maximum of
the tortuosity is found which indicates an inhomogeneous distribution of the coke layers
within the pore space. After full regeneration, the tortuosity does not reach its starting
value, i.e. that of the fresh catalyst, which means that the physical structure of the
catalysts cannot be fully regenerated, although the coke deposits have been completely
removed. Moreover, the difference between fresh catalyst and fully regenerated catalyst
with an initial coke content of 16 wt% is greater than that between the catalyst with 6
wt% initial coke content. Therefore, higher coked catalysts lead to less efficient
regeneration. Further investigations with more deactivation/regeneration cycles are
needed and currently done to study these phenomena in more details.

3.2. 129Xe NMR SPECTRA

The variation of the chemical shift of restricted 129Xe in the catalyst during the
regeneration process is shown in Fig. 3. The change of NMR line shapes indicates a
607
change of pore morphology and pore size distribution, which is in good agreement with
data obtained from the traditional BET method. The gore morphology of the fresh and
fully regenerated catalyst is different as shown by 9Xe NMR spectra of Fig.3 . The
NMR data of the catalyst during the regeneration process were correlated with average
pore sizes obtained by BET measurements (Fig. 4), indicating an almost linear
dependency, thereby it must be stated that more data are still needed to verify this trend.
9~----------------------~

25%

120 1\0 100 90 80 70 60 65 70 75 80 85 90 95 100 105 110


'2'J Xe chemical shift (ppm) ''''Xe chemical shift [ppm]
Fig.3 129Xe spectra of adsorbed Xe during the Fig.4 Average pore size of the catalyst
regeneration process. (The percentage indicates during the regeneration as a function of
relative amount of cokes removed) the chemical shift of 129Xe

3.3. CHEMICAL COMPOSITION OF THE COKE

The I3 C CP MAS NMR spectra oJ the coked catalyst samples at low and hi~
conversion levels show marked differences. For a low coke content a group of broad I C
resonance lines is identified, which can be assigned to aromatic (110-150 ppm) carbon.
Other lines correspond to spinning sidebands. With the increase of the coke content, this
group of resonance lines is gradually separated showing a shoulder peak at 137 ppm. In
contrast to the deactivation process, the 13C CP MAS NMR spectra for the catalyst
samples de co ked by regeneration show a faster initial disappearance of this shoulder
when the coke is burned off. Therefore, it can be assumed that there are two types of
aromatic coke in catalysts: a very reactive coke, which is rapidly burned off, and a coke
which is less reactive. According to the literature [8] and own measurements [9], the
more reactive coke is formed on the metal sites, and the less reactive coke is formed on

l
the acidic sites.
"c CP MAS 131 ppm 127 ppm
r, \l
/ \\ I1\ \ / .•
C.=16wt% ,~''\ ~ •
A=70% __ ... __.J:. ". __ r _______ ,._...-__ ~

1
C,=16wt% • i ~ 15 1\ "
a _~,.". __ /"\../
..r *
\J ~
oJ \V t\~. . ",,--.~
! C=16wt% /\~ .. j\\ •
i !\ ;\ ..
i
0 , l'. •
r.
'5>
~
C-5wt0'
,- 10 • J" 1~
Ij
1\
!;

f\ •
""/ \~>",--, ., __
2'
a::
A =10%
-----....

/'J '
; 'v l \
/\ . ./\
'J'/\
/ \.

• '-..... • ' - ""--"-


..
.,_ _ _ J/'" •

C,=3 wt% __.___


. ~
:\
'-'
,./\
'_/"~ --"--'-""_
C,=16 wt%
A =0% .-~/ \.../
j~ \ ! \. (\ (\
,j \.../ \.../ \--."-'- __

... 2ao '" 200 ,.. 100 ~ 1'20'" 60 .. 40 " 0 '!lPm

Fig.S 13C CP MAS spectra of the naphtha catalysts for different stages of deactivation and regeneration
processes (Co : initial coke content, A : relative amount of the coke removed). The peaks marked by
stars correspond to spinning sidebands.
608

4. Conclusions

Coking and decoking during catalytic hydrocarbon conversion on naphtha refonning


catalyst have been studied using a variety of NMR techniques. Longitudinal (T1) and
transverse (T2) relaxation times of IH of adsorbed liquid n-heptane have been measured
and correlated with pore morphology changes. Measurements of the self-diffusion
coefficient were performed using a PGSTE procedure and the tortuosity was evaluated
from the ratio of the diffusion coefficient of the confmed and the bulk liquid. It was
found that the physical structure of catalysts can not be fully regenerated, even though
the coke has been fully removed. Higher coke content leads to less efficient
regeneration. These results are conftrmed by those obtained from tortuosity
measurements. The pore size distributions, which show strong changes during the
different stages of the deactivation and reactivation processes as a result of encumbering
and blocking of the micropores have also been investigated from 129Xe NMR line
shapes. These results are in agreement with those obtained from the BET method via the
Kelvin equation and can also be corroborated with those obtained from tortuosity
measurements. It was shown that total elimination of carbonaceous residues by decoking
induces changes in the pore size distribution and tortuosity. By I3C CP MAS NMR
measurement, two types of aromatic cokes were detected. The very reactive coke, which
is rapidly burned off, is formed on the metal sites; the less reactive coke is formed on the
acidic sites.

References

1. Bell, A. T. and Pines, A. (1994) NMR techniques in catalysis, Marcel Dekker, New
York.
2. Bonardet, J. 1., Barrage, M. C., and Fraissard, J. (1995) Use of NMR techniques
for studying deactivation of zeolite by coking. J Mol. Catal. A: Chemical 96, 123-
143.
3. Paweewan, B., Barrie, P. J., and Gladden, 1. F. (1999) Coking and deactivation
during n-hexane cracking in ultrastable zeolite Y. Appl. Catal. A: General 185, 259-
268.
4. Callaghan, P. T. (1991) Principles 0/ Nuclear Magnetic Resonance Microscopy,
Clarendon Press, Oxford.
5. Kimmich, R. (1997) NMR Tomography, DijJusometry, Relaxometry. Springer,
Berlin.
6. Bliimich, B. (2000) NMR imaging o/materials, Oxford University Press, Oxford.
7. Cotts, R. M., Hoch, M. J. R., Sun, T., and Markert, 1. T. (1989) Pulsed fteld
stimulated echo methods for improved NMR diffusion measurements in
heterogeneous systems, J. Magn. Reson. 83, 252-266.
8. Parere, J. M. (1991) in C. H. Bartholomew and 1. B. Butt (eds.) Catalyst
Deactivation. Elsevier Science Publ. B.V., Amsterdam, pp. 103-110.
9. Jess, A., Hein, 0 ., and Kern, C. (1999) Deactivation and decoking of a naphtha
refonning catalyst. Study in Surface Science and Catalysis, Elsevier Science,
Amsterdam, pp. 81-88.
THE EPR AND ELECTRONIC SPECTROSCOPY STUDY OF POLY ARYLENE -
SULFOPHTHALIDES AND POLYDIPHENYLENEPHTHALIDE REDUCTION BY
ALKALI METALS

N.M. SHISHLOV, V.N. KHRUSTALEVA, SH.S. AKHMETZYANOV,


N.G. GILEV A, V.S. KOLOSNITSYN, O.G. KHVOSTENKO
Institute of Organic Chemistry, Vfa Research Center of RAS, Pro
Oktyabrya 71 ,450054 Vfa, Russia, , E-mail:shishlov@anrb.ru

The development of organic high-spin polyradicals is of current interest in basic


research as well as in materials science [1,2]. The structure ofpolyarylenesulfophthalides 1-
J and polydiphenylenephthalide 1, film-forming polymers containing labile cycles in the
main chains, looks very promising for chemical or physical generation of stable
polyradicals of triarylmethyl type (PRTAMT). The kuown chemical properties of phthalide
[3] and sulfophthalide cycles [4] allow to hope for their reductive opening. On the other
hand it is known that reduction by alkali metals is employed for the preparation of
polyanions and polyradicals [5]. So, we tried to obtain radicals or their derivatives in
polymers ! -1 by alkali metal reduction and these reactions were studied with EPR and
electronic spectroscopy methods.

! - R=S02
1-R=CO
The synthesis of the polymers has been described elsewhere [6]. Experiments
were carried out mainly as follows. A polymer solution in freshly distilled dried DMSO or
DMFA was placed in a quartz cuvette 1 cm thick (d = lcm) with a ground-glass stopper. In
a dry box, small pieces of freshly cut lithium or sodium were placed in the cuvette. The
solution was purged with argon for 10 min, the cuvette was plugged with the ground-glass
stopper, and the electronic spectrum of the solution was recorded on a Specord M-400
spectrophotometer. The cuvette containing the solution was agitated in a shaker, and the
spectra of the solution were recorded at intervals during shaking. The portions of solution
were transferred into EPR tubes (0 - 4 rom) under an argon atmosphere by the use of a
glass syringe. EPR spectra were recorded at 77 K and room temperature on SE/X-2544
"Radiopan" spectrometer.
609
1 Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Inteiface Science, 609-614.
© 2002 Kluwer Academic Publishers.
610

Immediately after lithium was placed in the solutions of polymers 1 and 1, and the
solutions were purged with argon, the solutions turned blue and gave rise to the absorption
bands (AB's) at Amax= 568 run for 1 and Amax= 576 run for 1 (Figs. land 2) which are
responsible for this color. The blue color of the solutions became deeper during shaking,
the intensity of the bands in the region of 570 run increased, and the bands at 350 run for 1
and at 360 run for 1 emerged and built up symbatic with the bands at 570 run. The blue
band for polymer 1. in the initial ( after lithium was added) spectrum was much stronger
than that for polymer 1, and the intensity growth rate for this band was also somewhat
higher (Fig. 4). Furthermore, weak AB's were detected for polymers 1 and 1. in the region
of 420 run (Figs.1 and 2), which changed slightly during shaking of the solutions. The
bands at 420 run were hard to observe when the AB's at 570 and 360 run became very
strong. The width and asymmetry of the AB's at 570 run were increased with an increase in
the concentration of the blue color centers (CC's) for either of the polymers. A weak EPR
singlet (g = 2.0028 ± 0.0001; ~H ~ 10 Oe) was detected also for the color solutions of 1
and 1. (Fig. 5).
A n------.-----.----.-...---..-.... ~

1 .
1.6
3
1.2

0.8 2

2
/~l-'--" ~ "
I
I
0.4
-----~
- ' ~
0.0 0.0 .,-----:...,:r----~-- - . -.-..-...-..~- ....-- ..
300 400 500 600 700 800 300 400 500 600 700 800
A, run A, run
Fig. I. Electronic spectra of the system (DMSO + Fig. 2. Electronic spectra of the system (DMSO +
! [10.3 moll1] +Li~: (1) initial system (without Li~; ~ [I 0.3 mol/l] +Li~: (I) initial system (without Li~;
(2) after addition of Lioand purging with argon for (2) after addition of Lio and purging with argon for
10 min; (3) after a 45-min shaking; d = I em. 10 min; d = 1 cm.

A r-r-- 3
7
6
~i--~-------------------'
j
1.6! . 5,
I
;::s :
\.2 1 .ci 4 i
\ .\ li! :
0.8
('
,\ 4 ....~ 32
0.4 1
o
0.0· -l~ ___- -__- -__- -__ -·_--~

300 400 500' 600 700 800


o 100 200 300 400 500
A, run
time, min
Fig. 3. Electronic spectra of the system (DMSO +
J [10.3 molll] +Lio): (I) initial system (without Lio); Fig. 4. Optical density of (polymer.!! [10. 3 molll) +
(2) after addition of Lio and purging with argon for Lio + DMSO) systems vs. shaking time at: (I) 568 nm
10 min; (3) after a 6.5-h shaking; (4) after a 7-h c.!! =1); (2) 576 nm C.!! = l); (3) 430 nm C.!! = J).
shaking; d = I cm.
611

~ lOOe H

Fig.5. (I) EPR spectrum of the system (DMSO +! [0.2 mol/I) +Lio) after 72 h shaking;
(2) EPR spectrum of the system (DMSO +~ [I0-3 molll) +Lio) after 50 h shaking.

The electronic spectrum of a polymer J solution after addition of lithium, and


argon purging (before shaking), showed the AB at Amax= 430 nm (Fig.3). In the beginning
of the shaking process, this band somewhat increased, but a sharp growth of this AB began
only after - 5 h of shaking. At the same time, a weak band with Amax= 638 nm appeared and
increased symbatic with the AB at 430 nm (Fig. 3). The abrupt increase in the intensity of
the band at 430 nm only after prolonged shaking (4-5 h for polymer concentration 10"3
mol/l) was consistently observed in repeated experiments and was unlikely to be accidental.
An EPR singlet (g = 2.0028 ± 0.0001; Mi ~ 10 Oe) that increases in proportion with AB's
at 430 and 638 nm was observed also for the reduction of polymer J (Fig.5).
For the polymers !-J, the AB's in the region 410 - 430 nm were observed
previously in other reductive systems and were assigned to the allowed electronic
transitions in the radicals of the triarylmethyl type (see [7,8] and references therein). The
weak band at 638 nm in the case of polymer J is most likely to be related to the forbidden
transition in the RTAMT [9], which in this case has the structure of distonic radical anion 2
with an unpaired electron on the quaternary carbon atom and a negative charge on the sulfo
group, and most likely results from the sulfophthalide cycle opening caused by electron
transfer from lithium to the polymer. The parameters of observed EPR signals correspond
the RTAMT's too [10,11]. It is not known yet why the sharp growth in the concentration of
the radicals upon interaction of J with Lio is observed under these conditions only after
prolonged shaking. It should be noted that long (- 24 h) reaction oq with Lio, as estimated
from the intensity of the AB at 430 nm, provides an opportunity to convert almost 100% of
quaternary carbon atoms in the polymer chain into a radical state, i.e., to obtain a
polyradicals with high concentration of unpaired electrons. In our estimates, we assumed
that the value of molar extinction for polyradicals E = 4.104 as for similar monomeric
radicals [9] . As it was said in the introduction already, the preparation of polyradicals and
the study of their properties is associated with the problem of organic ferromagnetic
materials and is an interesting and relevant task [1,2].
612

The blue CC's in polymers! and l resulting from their reaction with lithium are
similar to the centers observed in an aniline-cyclohexanone mixture and upon alkaline
hydrolysis [7,8]. Earlier, they were assigned to quinoid structures of the Chichibabin's
hydrocarbon type [8]. The higher formation rate of the blue CC's for polymer l compared
to that for polymer ! is likely to be due to energy reasons, i.e., the transformation of the
fluorenyl bridge to the quinoid structure is more advantageous over that of the biphenyl
bridge. For polymer J, the formation of the quinoid structure is energetically
disadvantageous, since the Muller's hydrocarbon, which is the model system for J, is most
likely to have a triplet ground state [12]. In our opinion, this clearly pronounced
dependence of the properties of the blue CC's (the formation itself and the rate of
formation) on the nature of the bridge in the polymer chain is important but indirect
evidence for their suggested quinoid structure. Recall that the presence of sulfur or oxygen
atoms between the phehyl rings in the bridge also prevents the formation of blue CC's upon
dissolution of similar polymers in an aniline-cyclohexanone mixture [7]. The above-
mentioned increase in the width and asymmetry of the AB's at 570 nm with an increase in
the concentration of blue CC's is associated with the formation of fragments containing a
quinoid-benzoid conjugation in the main chain of the polymer, as in the case of alkaline
hydrolysis [8]. It should be noted that the formation of the uninterrupted chain of such
conjugation should impart interesting electrophysical and physicochemical properties to
polymers [13,14]. On the other hand, it is clear that in the polymer an isolated RTAMT's
must be formed, and these defects will interrupt the conjugation chain.
Taking into consideration the fact that the quinoid structures, like the
Chichibabin's hydrocarbon, result from electron transfer reactions, we will notice the
following. It should be supposed that the quinoid structures arise upon concerted reduction
of neighboring sulfophthalide cycles, which may be represented by the following scheme:

That is, in this case we observe a brightly pronounced effect of the neighboring group,
which, in principle, is well known for polymer-analogous reactions [15]. This effect should
appear not only in polymers but also in any molecules containing two connected sites of
613

equal reactivity, in particular, in the synthesis of the classical Tiele's and Chichibabin's
hydrocarbons. It is clear that the energetic advantage of the quinoid structure is one of the
components of the driving force for these reactions, but the details of these transformations
are not always explicit and deserve additional investigation. For example, the above scheme
implies that the formation of the radical at one quaternary carbon atom causes a change in
the electronic structure of the neighboring sulfophthalide cycle that favors electron capture
from metal to this cycle. But it is very likely that second electron is captured immediately
by the radical center and then, via the bridge mechanism, which is often observed for
chemical and biological systems [16], transfers to the neighboring sulfophthalide cycle.
Cooperative phenomena in electron transfer reactions currently attract attention of
researches [16] and polyarylenesulfophthalides may be convenient objects for studying
these phenomena.
It may be noted briefly several features of the sodium use for the reduction of
polymers !-J in DMSO: 1. The reduction is accelerated considerably as compared to it by
lithium; 2. The results of the reduction dependence on the sodium content in the reactive
system. As in the lithium case, the quinoid structures and RTAMT's are observed for !-J at
a little sodium content. The higher sodium content leads for! to the consequent appearance
of two new CC's with AB's at 750 and 640 nm. The new CC's with AB at 725 nm are
formed for J in this case. The given new centers are unparamagnetic and are unidentified as
yet; 3. The reduction is complicated by the known [17] interaction of the sodium with
DMSO, that is accompanied by the obvious gassing.
It was of interest to compare the reductive properties of phthalide and
sulfophthalide cycles. Unfortunately polymer ~ is not solved in DMSO, so the experiments
with 4 were performed in DMFA. It is worth to note that Lio and Nao react visibly with
DMF A and this fact complicates a interpretation of the experimental data. The reduction of
~ by lithium and sodium in DMFA gives several CC's with visible AB's at 572,600,676,

805 and 820 nm. These AB's having a complex time


evolution are formed and observed not at the same time. A
paramagnetic species giving EPR quartet with a hyperfine
splitting - 1 Oe and g=2.0049 (Fig. 6a) are fixed for reduction
of ~ at room temperature too. This quartet is transformed to
the singlet with "wings" at 77 K (Fig. 6b). The question of the
PMS's and CC's structure for polymer ~ remains open. One
may see on Fig. 6 the weak asymmetric singlet (~H ~ 0.2 Oe;
AlB ~ 1.5) at g = 2.0023 that is related evidently to the small
colloid lithium particles [18].

Fig. 6 EPR spectra of the system


(1 [0.2 moJll)+DMFA+Lio)
after 6 h shaking: a - at 300 K;
b -at 77 K.

Acknowledgment. This work was supported by the Russian Foundation for Basic
Research ( grant No. 00-03-33088)
614

References
I. Rajca, A. (1994) Organic Diradicals and Polyradicals: From Spin Coupling to Magnetism?, Chem. Rev. 94,
871-892.
2. Crayston, 1.A., Devine, J.N., and Walton J.e. (2000) Conceptual and synthetic strategies for the preparation
of organic magnets, Tetrahedron 56, 7829-7857.
3. Sabitha G., Yadav 1.S. (1998) Reductive cleavage of phthalides with iodotrirnethylsilane, Synthetic
Communications 28, 3065-3071.
4. Roberts, D.W. and Williams, D.L. (1987) Sultone chemistry, Tetrahedron 43,1027-\062.
5. Rajca, S. and Rajca, A. (1995) Novel high-spin molecules: 1t-conjugated polyradical polyanions.
Ferromagnetic spin coupling and electron localization, JAm. Chem.Soc. 117, 9172-9179.
6. Zolotukhin, M.G., Akhmetzyanov, Sh. S., Lachinov, A.N., Shishlov, N.M., Salazkin, S.N., Sangalov, Yu. A.
and Kapina, A.P. (1990) Polyarylenesulfophthalides, Dokl. Acad. Nauk SSSR 312, 1134-1137.
7. Shishlov, N.M., Akhmetzyanov, Sh. S. and Khrustaleva, V.N. (1997) Color reactions of
polyarylenesulfophthalides, Russian Chem. Bull. 46,377-379. .
B. Shishlov, N.M., Khrustaleva, V.N, Akhmetzyanov, Sh. S., Murinov, K.Yu., Asfandiarov, N.L. and Lachinov,
A.N. (2000) Formation of color centers and paramagnetic species by alkaline hydrolysis of
polydiphenylenesulfophtalide, Russian Chem. Bull. 49,298-302.
9. Chu, T.L. and Weissman, SJ (1954) Symmetry classification of the energy levels of some triarylmethyl free
radicals and their cations, 1. Chem. Phys. 22,21-25.
10. Chesnut, D.B. and Sloan, GJ. (1960) Paramagnetic resonance absorption of triphenylmethyl, 1. Chem. Phys.
33,637-638.
II. Braun , D. and Zehman, P. (1976) Stabile Polykohlenstoffradikale, 2: Versuche zur Darstellung von
Polyradicalen des Triphenylmethyl-typs, verkniipft ilber p-Phenylen-Einheiten, Macromol. Chem. 177, 1387-
1400.
12. Shishlov, N.M. and Asfandiarov, N.L. (2000) Color centers in poly(arylenesulfophtalides) and the ground
state of Muller's hydrocarbon molecule, Russian Chem. Bull. 49, 1676-1681.
13. Boudreaux, D.S., Chance, RR, Elsenbaurner, RZ., Frommer, J.E., Bredas, J.Z. and Silbey, R (1985) New
class ofsoliton-suPP'lrting polymers: Theoretical predictions, Phys. Review B 31, 652-655.
14. Chen, W.-Ch. and Jenekhe, S.A. (1998) Model compound studies of small bandgap conjugated
poly(heteroarylene methines), Macromol. Chem. Phys. 199,655-666.
15. Plate, N.A., Litmanovich, A.D. and Noa, O.V. (1977) Macromolekulyarnye reaktsii (Macromolecular
reactions). Khimiya, Moscow
16. Komyshov, A.A., Kuznetsov, A.M., Ulstrup, J. and Stimming, U. (1997) Medium effects on elementary
charge transfer processes in liquid and solid environments, 1. Phys. Chem. BIOI, 5917-5935 .
f7. Butler, J.N. (1967) Electrochemistry in dimethyl sulfoxide, 1. Electroanal. Chem. 14,89-116.
lB. Webb, R.H. (1967) Electron-Spin-Resonance line shape in spherical metal particles, Phys. Rev. 158, 225-
233.
195pt NMR-FOURIER SPECTROSCOPY IN THE ANALYSIS OF THE MECHANISM OF
THE CYTOSTATIC ACTIVITY OF PLATINUM COMPLEXES

V.E.STEFANOV and A.A.TULUB


Department of Biochemistry, St.Petersburg State University, Universitetskaya nab. 7/9,
St.Petersburg 199034, Russia

Interaction of platinum complexes with two model targets (polynucleotides of different length
and GTP-tubulin) was analyzed by means of 195pt NMR-Fourier spectroscopy. Slow hydrolysis
of cis-dichlorodiamineplatinum (complex I) is followed by rapid binding of the released aqua-
form cis-[Pt(NH))zCIHpr (complex II) with polynucleotides, two signals being recorded at 8 =
-1841 and -2304 ppm. Unlike the aqua-form, the monohydroxo-form, cis -[Pt(NH))zOHCll
(complex III), interacts withpolynucleotides very slowly. The signal at 8 = -2304 ppm is shifted
downfield (8 = -2450 ppm). An agreement between the resonance at 8 = -2450 ppm and that of
the tetracoordinated complex of Pt(ll) is supported by the resonance of cis-[Pt(NH)),ll+ 111 the
same spectral region (8 = -2470 ppm). Cyclization of monofunctional adducts of cis-
[Pt(NH 3MN)Clf (IV) into bifunctional adducts is slower than monofunctional binding of the
aqua-form (II). Removing cloride ligands with AgNO) yields cis-[Pt(NH)MN)HP]2+ (complex
V), which immediately forms a chelate giving rise to a resonance at 8 = -2450 ppm. 195pt NMR-
Fourier spectroscopy analysis of interaction of cis-dichlorodiamineplatinum with tubulin bound
GTP showed that originally observed resonance in NMR 195pt spectra at -2060ppm decreases
giving rise to a resonance at -2030ppm, which corresponds to the bidentate coordination of th~
platinum complex. Mechanisms of action of platinum complexes on the intercellular molecular
targets and nature of cytostatic effects of platinum are discussed on the basis of the obtained 195pt
NMR -spectroscopy data.

1. Introduction

NMR-spectroscopy analysis is an efficient tool in the investigation of interactions of


biological macromolecules with physiologically active compounds. Platinum-containing
615
1. Fraissard and O. Lapina (cds.). Magnetic Resonance in Colloid and interjiu:e Science. 615-623.
i(;) 2002 Kluwer Academic Publishers.
616

compounds has attracted interest of researchers since the discovery of antitumor activity of cis-
dichlorodiamineplatinum (cisplatin) [I). Later numerous platinum complexes of different
oxidation states (II-IV) have been successfully synthesized and tested [2-6). Antitumor activity
of platinum complexes is commonly explained by their binding with guano sines of DNA
affecting DNA replication and transcription via producing intrastrand N7(G)-Pt-N7(G) kink
structure [7,8). DNA is considered a major cellular target of platinum compounds due te a very
efficient binding of platinum with DNA [7) . Other plausible cellular targets for platinum, though
100 to 1000 times less efficient than DNA, are RNA, membrane phospholipids, cytoskeleton
microfilaments, etc.[7-9). Of particular interest are nucleotide-dependent G- proteins owing to
their involvement in the processes of cell division and signaling. G-proteins, tubulin including,
manifest their specific biological activity only on binding GTP, whose N7 atom of guanosine can

be attacked by platinum. Similarity observed in chemical mechanism in the above two cases
makes reasonable considering also a model system involving polynucleotides, containing nitrous
bases:

To study biologically significant interaction of platinum complexes with molecules


involving nitrous bases, in the present work, we used the 195pt NMR-Fourier spectroscopy.
Platinum is not contained in the compounds we analyse, it has a low multiplicity of transitions
and acts as a coordination center. This facilitates interpretation of the results. Resonances from
other atomic nuclei are often ambiguous and difficult to interpret.

2. Materials and methods

In the first part of the analysis we used isotopically enriched (97% 195pt) platinum
complex I - cisplatin - and KzPtCl 6 (internal standard) purchased from the Laboratory of Nuclear
Biophysics (CERN, Swetzerland), dinucleotides d(GG) , d(AA) , d(CC) and their mixed
derivatives, d(AG) and d(CG) (G - guanosine, A - adenosine, and C - cytidine) purchased from
SIGMA CHEMICALS. Polynucleotides of the composition d(GGG), d(CCG), d(AAA), d(CCC),
d(S'AGGCCC3), d(S'CGGCCA3), d(5'CCGGCCCA3), d(5'GAGACCGAGC3j,

d(S 'CCCCGGCCCC3) were kindly provided by P. Snezhich (Pharmaceutical firm "SANDOZ"


, Austria). The monoaqua-form (complex II) C95Pt) was obtained in the reaction of complex (I)
with equimolar quantity of AgNO) in the excess of dimethylformamide (DMF). The extracted
617

cis-C95Pt(NHl)2Cl(DMF)f was transfered to polynucleotide solution. DMF is easily replaced by


water molecules.
For the NMR analysis we synthesized the following nucleotide-methylated
complexes : cis-[Pt(NHlMW-G)Clf, cis-[Pt(NHlMW-(G)2l2\ cis-[Pt(NH lMN 7-A)Clf, cis-
[Pt(NH lMN7-A)2l2+, cis-[Pt(NHlMNl-c)CIr, cis-[Pt(NH lMN l -C)2f+ (coordination is over the
atoms in parantheses). Methyl groups were introduced into complexes to prevent plausible
coordination via other atoms of bases (atoms N I and Nl (G, A); for cytidine there is only one N'
atom available for coordination; coordination via oxygen atoms does not occur). The
composition of complexes was determined with a mass-spectrometer (Plasma Quad PQ2-Turbo',
St.-Petersburg, Russia). Cis-DDP, GTP and the compon'ents of magnesium-guanosine (MG)
buffer were purchased from SIGMA CHEMICALS
In the second part of the work we used isotope enriched cis-DDP (97% 195pt), K2PtCI.
(internal standard), GTP (MetN 7-GTP) (98% liP) and HlPO. (internal standard) for NMR
spectroscopy (initial products were from SIGMA CHEMICALS), which were obtained from
Dubna Nuclear Center, Russia. 195pt and lip NMR spectra were registered at 37° C with AM-500
BRUKER using the internal standard [K 2PtCI4 + DCl or Hl P0 4 (0.2 mol/l). The acquisition time
did not exceed 0,8 ms, the decoupling technique was used. Details regarding the methods used
are also provided in [10].

3. Results and discussion


The obtained NMR spectra (Fig. 1,2) reveal complexation of polynucleotides with
complex (I) and its monoaqua-form (II), respectively. In the beginning, solutions of complexes
(I) and (II) manifest resonances at -2149 and -1841 ppm. As complexation proceeds, the intensity
of resonances decreases giving rise to a resonance at -2304 and eventually at -2450 ppm. An
agreement between these resonances and those of synthesized complexes cis-[Pt(NHlh(N '-
G)C!]' (8 =-2302 ppm) and cis-[Pt(NH lMN 7-G)2l 2+(8 =-2452 ppm) suggests that resonances at -
2304 and -2450 ppm correspond to the monodentante and bidentante coordination of
polynucleotides, respectively.
618

-1800 -2100 -2400 -2700


cr, ppm

Fig. 1. Dynamic J9'Pt NMR-Fourier spectrum of binding of complex (I) to d(GG). I - 35 min, 2 - 170 min,
3 - 250 min, 4 - 400 min, 5 - 510 min, 6 - 625 min, 7 - 730 min after the beginning of the reaction

-1800 -2100 -2400 -2700 cr, ppm

Fig. 2. Dynamic 19'pt NMR-Fourier spectrum of binding of aqua-form (II) with d(GG). I - 1 min, 2 - 3
min, 3 - 6 min, 4 - 120 min after the beginning of the reaction

We showed that d(AA) and d(eC) produced resonances at -2352 and -2364 ppm
(monodentante coordination), -2477 and -2484 ppm (bidentante coordination). Resonances are
very specific and correspond to those of structurally identified complexes of cis-[Pt(NH1MN 7-
A)CIf (8 = -2350 ppm) and cis-[Pt(NH1MN1-C)C\t (8 = -2361 ppm), cis-[Pt(NH1MN 7-A)2Y' (8
= -2474 ppm) and cis-[Pt(NH1MW-C)2]2+ (8 = -2482 ppm). The fact that the resonance position
619

depends on the nucleotide type (G, A or C) enables to define the coordination sites in extended
nucleotide sequences.
Analysis of resonances registered for platinum-bonded nucleotides (Table \)
demonstrated that guanosine G (8 "" -2305 ppm) is the most preferred coordination site in the
polynucleotide chain, followed by C cytidine (8 "" -2367 ppm), and then by A adenosine (8 = -

2352 ppm). The polynucleotide length had no significant effect on the position of resonances.
Whatever length the polynucleotide chain was, the platinum complex firstly attacked G and then
C. A is coordinated only if there is no G or C in the chain.

Table 1.
19Spt chemical shifts (8, ppm) registered for binding of complex (I) with polynucleotides

Adduct
Polynucleotide

monofunctional Bifunctional

-2304 -2450
d(GG)

-2351 -2466
d(AA)

-2367 -2475
d(CC)

-2307 -2456
d(AG)

-2304 -2471
d(CG)

-2305 -2452
d(CGG)

-2304 -2472
d(CCG)

-2354 -2465
d(AAA)

-2368 -2479
d(CCC)

-2304 -2452
d(5 'AGGCCC3 ';

d(5 'CGGCCA3 ';


-2304 -2452

-2305 -2453
d(5 'CCGGCCCA3)

-2307 -2462
d(5 'GAGACCGAGC3)

-2304 -2452
d(5 'CCCCGGCCCC3 )
620

We presumed the following mechanism of binding of complex (I) with polynucleotides.


On dissolving, a slow hydrolysis of (I) occurs (the chloride ligand is replaced by a water
molecule, the rate-limiting process) followed by rapid binding of the generated aqua-form (II) to
the polynucleotide. Only two signals (at Ii = -1841 and -2304 ppm) are registered (Fig. 2) for
monofunctional binding with the "pure" aqua-form (II).
Unlike this complex, the highly inert monohydroxo-form, cis -[Pt(NHl)PHCl] (III),
interacts very slowly with polynucleotides. A high reactivity of the complex (II), compared to (I)
and (III), can also be explained by the positive charge of (II), centered on W(G, A) and Nl(C)
atoms, which favors its interaction with negatively charged DNA bases. Thus , using NMR -
spectroscopy, we have proved that the aqua-form cis-[Pt(NHl)lCIHPf is the most reactive form
of cisplatin. Its reactivity is the result of the increased lability of water molecules in complexes
compared to that of chloride ions. Ordered structure of DNA provides favorable conditions for
efficient binding with such potential tumor killers as platinum complexes ..
We suggested that nucleotide-dependent G- proteins may also be cellular targets for
cytostatic activity of platinum. Their involvement in numerous cellular and intercellular
processes, including cell division and signaling depends on the state of GTP. Colchicinoids [11,
12] were known as antitumor drugs long before the discovery of cisplatin and JM-216. [13,14]
They attack cytoskeletal proteins, mostly such G-protein as tubulin. We assumed that cisplatin
can also affect tubulin via binding with GTP, containing a guanosine moiety with N7 atom,
which is exposed to platinum attack. To investigate the mechanism involved we undertook tQ;p!
and lip NMR spectroscopy studies, using the methods and materials described earlier in [10,15)
and in "Materials and methods". To increase binding rates during platination we used the diaqua
form (cis-DA) of cisplatin - cis-[Pt(NHlMHP)l]l+ , which arises due to hydrolysis of cisplatin.
The kinetic 19Spt NMR study of aggregated platinum bound tubulin (tubulin in platinum
MG buffer, R = 1:1) revealed two distinct resonances at -1590 and -2030 ppm (FigJ, a.) The
first, less intensive, corresponds to cis-DA. Its position is determined [16]. The resonance results
from incomplete binding of cis-DA to tubulin or partial dissociation of platinum bound tubulin
during the formation of platinum bound associates. Its position does not change with time but the
intensity slightly drops, testifying to partial intercalation of the platinum complex into the protein
matrix. The second resonance is broad. Its position, corresponds to magnetic absorbance of
monofunctional or bifunctional adduct (nonsymmetrical coordination on nitrogen, sulfur or
oxygen [10, 17,18]). Monofunctional and bifunctional adducts, when bound to nitrogen
containing ligand, have resonance at -2060 and -2450 ppm, respectively [10,19,20] . Coordination
on sulfur corresponds to a resonance at -2120 ppm (monofunctional adduct) and -2504 ppm
621

(bifunctional adduct) [21]. Complexes of a chelate type with coordination on oxygen produce
resonances in the range -1500"," -1600 ppm [10,20].

SA

cis-OA 1\
~~-----I1~=_~=--=--1'2
a --'--1 ..
cis-OA

---itA
-:::=:-:::-._::
noise
)

3[:.-~ 2

b
OPI' ppm

·1600 ·21IUO
Ol'I'Ppm

Fig.3. Dynamic I9IPt NMR spectra of tubulin-containing solutions (R = I: I. 37°). a - cis-DA-platinated


tubulin (I - 17 Min, 2 - 20 h, 3 - 70 h after the onset of assembly), b - (Met-N')GTP-platinaled
tubulin [/ - 25 s (average over 25 independent tests). 2 - 20 h, 3 - 70 h after the onset of assembly].
c - cis-DA with GTP (I - 30 s. 2 - 45 s, 3 - I h after the onset of assembly): MA - monoadducl of
cis-DA bound tubulin; BA - biadduct of cis-DA bound tubulin

To prove that cis-DA stops tubulin assembly via binding to GTP rather than to other
possible targets in tubulin, we used (N 7)methyl analog of GTP - (Met-N')GTP, which is not
accessible to the attack of platinum on N7 but can bind with tubulin. Unlike in the case of GTP
tubulin, 195Pt NMR spectra of (Met-N 7)GTP tubulin reveal no signal at -2030 ppm (Fig.3, b). The
observed weak gain in intensity (noise level) can be assigned to chaotic binding of small
quantities of cis-DA to amino acid residues in tubulin. A sharp signal at -1590 ppm corresponds
to cis-DA. Constant signal intensity proves that no binding of cis-DA to (Met-N 7)GTP tubulin or
its aggregates takes place.
622

The inhibition of tubulin assembly must result from binding of cis-DA to GTP through
N7 of GTP guanine ring. Analysis of resonances of platinum bound complexes with organic and
bioorganic ligands, including polynucleotides, (10,17,18,20,22-25) suggests that the resonance at
-2030 ppm is most likely due to a chelate formation involving Pt-N and Pt-O bonds. The
assumption is supported by the fact that the monofunctional adduct, cis-[Pt(NHl)iN)OHlt, has
a resonance at -2060 ppm, while bifunctional adducts usually have resonances in the down field
region: -2020 .;. -2045 ppm. 195Pt NMR studies show that binding of cis-DA to GTP (R = I: I)
generates two consecutive signals, at -2063 then at -2031 ppm (Fig.3, c)
Additional evidence in support of chelate formation was obtained froJll lip NMR. studies
of interaction of cisplat in with tubulin-bound GTP. Information on lip NMR spectrum for GTP is
available [26,27]. Three GTP phosphorus, PY, po, and pP, have resonances at -5.6, -10.9, and -
21.4 ppm, respectively (internal standard HlPO.). Interaction of cis-DA with GTP shifts signal
from po downfield by 1.8 ppm, which is the result of a-phosphate perturbation due to
coordination. Resonances from pI and pP almost do not change, the downfield shift for pP and P'
is 0.1 and 0.04 ppm, respectively. The obtained data suggest that cisplatin attack on tubulin,
stopping its assembly, proceeds via a two-step cisplatin binding to GTP in the GTP center of
tubulin. Thus, the use of NMR spectroscopy provide a very efficient tool in the analysis of
molecular mechanisms of cytostatic activity of platinum complexes.

References

(1) Rosenberg B., van Camp L., Krigas T. (1965). Nature 20S, 698.
[2) S. Neidle and M. Waring, editors (1993). Molecular Aspects of Anticancer Drug-DNA interactions. McMillan
Press Ltd., London, 169-212.
[3) Raynaud F., Boxall F.E., Kelland L.R.(1997). CUn. Cancer Res. 3, 2063.
[4) Giandomenico C.M., Abrams MJ., Murrer B.A., Vollano J.F., Rheinheimer M.1. (1995) Inorg. Chem .34, 1015.
[5) Galanski M., Keppler B.K. (1996). Inorg. Chem. 35,1709.
[6) Wong E., Giandomenico C.M. (1999) Chem.Rev. 99,2451.
[7) Jamieson E.R., Lippard SJ. (1999) Chem. Rev. 99, 2467.
[8) Reedijk 1. (1999) Chem. Rev. 99,2499.
[9) Mcintoch D.P., Cooke R.I., McLachlan A.I., Daley-Yaks P.T., Rowland M. 1. (1997) Pharm. Sci. 86, 1478.
[10) Stefanov V.E. , Tulub A.A., Kutin A.A. (1999) International J.Biological Macromolecules, 26, 161
[II) Andreu l.M., GorbunoffMJ., Medrano FJ., Rossi M., TimasheffS.N. (1991) BiochemislJy. 31,: 8080.
[12) Medrano F.J., Andreu J.M., GorbunoffMJ., TimasheffS.N.(1991) Biochemistry. 31, 3770
[13) Wong E., Giandomenico C.M. (1999) Chern. Rev. 99, 2451
[14) Neidle S., Snook C.F. , Murrer B.A., Barnard FJ. (1995) Acta Cryst. 51,822.
[IS] Stefanov V.E., Tulub A.A., Kutin A.A. (2001) /n/erna/ionaIJ.Biological Macromolecules, 28,191
[16) Sherman S.E., Gibson D., Wang A.H. , Lippard SJ. (1988) J. Am. Chern. Soc. 110,7368
[17) Appleton T.G., Bamham KJ., Hall I.R., Mathieson M.T. (1991) Inorg. Chern. 30, 2751
(18) Appleton T.G., Bamham KJ., Byriel K.A ., Hall I.R., Kennard C.H.L .. Mathieson M.T., Penman K.G. (1995)
Inorg. Chern. 34, 6040
623

[19J Watabe M. , Kobayashi T. , Kawahashi T., Hino A., Watanabe T., Mikami T., Matsumoto T., Suzuki M. (\999)
1. Inarg. Biachem.73, 1
[20) Bancroft D.P., Lepre C.A., Lippard SJ. (1990)1. Am. Chem. Sac.1I2, 6860
[21) Lin Z., Hall M.B. (1991) Inarg. Chem. 30,646
[22) Barnham KJ ., Djuran M.l., Murdoch P.S., Ranford J.D., Sadler PJ. (1996) Inorg. Chem. 35, 1065
[23) Talman E.G, Bruning W., Reedijk J., Spek A.L., Veldman N. (1997) Inarg. Chem. 36, 854
[24] Matsunami J., Urata H., Matsumoto K. (1995) Inarg. Chem. 34,202
[25] O'Halloran T.V., Lippard SJ. (1989) Inarg. Chem.28, 1289
[26] Kemp W. (1986) NMR in Chemistry. London: McMillan.
[27] Bovey F.A., Mirau P.A. (1996) NMR a/Polymers. San Diego. Acad. Press
NITROXYL RADICALS AS SPIN PROBES FOR THE STUDY OF LEWIS AND
BRONSTED ACID SITES OF OXIDE CATALYSTS

O.Yu. Ovsyannikova l.2, A.V. Timoshok l, and A.M.Volodinl


I Boreskov Institute of Catalysis
Pro Ak.Lavrentieva 5. Novosibirsk 630090, Russia
2Novosibirsk State University
Pirogova st. 2, Novosibirsk 630090, Russia

1. Introduction

The method of paramagnetic probe is widely used in the investigation of catalysts such
as silica, alumina, alumosilicates, zeolites etc. which do not possess their own
paramagnetic sites. The stable nitroxyl radicals are often used as such spin probes [1].
Direct interaction of >N-O· group of nitroxyl radical with the catalysts of Lewis
acid type in which the cation nuclei possess non-zero spin tends to appear alternative
superfine splitting [2,3). The magnitude of AI changes depending on the electron-spin
density on the nitrogen atom of the nitroxyl radical. In many cases the analysis of the
anisotropic constant A ft allows to receive information about the nature of adsorption
sites.
Thus, received ESR spectra of adsorbed nitroxyl radicals allows to analyze the
interaction of these molecules with adsorption surface sites and in a number of cases to
observe directly the adsorption sites.
On the other hand, it is known that nitroxyl radicals can be protonated by nitroxyl
groups in a strong acid solution to form the paramagnetic complexes [4]. These
complexes are strong oxidizing agents and are able to disproportionate and form
oxoammonium cations and hydroxylamines according to next scheme (1):

""N-O'+ H+ - + -OH
""N ,
/ / (1)

""N-OH + N - Q ' -'"N-OH + '"N=O


+, ' " +
/ / / /
The possibility of this process depends on many factors: acid strength, presence of more
basic functional groups in radical, nitroxyl group basicity.
625
1. Fraissard and O. Lapina (eds.). Magnetic Resonance in Colloid and Interface Science, 625-630.
© 2002 Kluwer Academic Publishers.
626

That 3-imidazoline derivatives are more stable to protonation in homogeneous


systems than piperidine nitroxyl radicals is well known [4J. In the present work we try to
observe the similar regularity in heterogeneous system. For this purpose, the correlations
between the stability of adsorbed nitroxyl radicals, their structure, and acidic properties
of the catalysts were studied.

2. Experimental

Fluorinated alumina (surface area -200 m 2 /g) was prepared by treatment of


y-A1 2 0 3 (surface area 226 m 2/g) with hydrofluoric acid (y-AI203/HF-1.4
wt.%). Then, the sample was dried at 120°C for 2 h and calcined in air
at 600°C for 2 h.
Sulfated alumina (y-AI203/S03-8.6 wt.%, surface area -150 m 2/g) was
prepared by the procedure described in our previous article [5]. For dehydroxylation of
the catalysts the samples (30-50 mg) were activated in air at 500°C for several hours.
Spin probes used for study of the active sites of the catalysts were the following
stable nitroxyl radicals with essentially differing oxidative-reductive properties:

TEMPO TEMPON NR

£) ):NN'x
0

0 I-
0 0
I-
0
I-
2,2,6,6-tetramethyl-piperidine-I-oxyl (TEMPO, 0.535 mmolll);
2,2,6,6-tetramethyl-piperidine-4-oxo-I-oxyl (TEMPON, 0.534 mmolll);
2,2,5,5-tetramethyl-4-phenyl-3-imidazoline-I-oxyl (NR, 0.539 mmolll).

Nitroxyl radicals were adsorbed from their solution in hexane at normal


conditions.
It is to note that for all used radicals spin density is mainly localized on N-O
fragment.
X-band ESR spectra were recorded at normal conditions on an ESR-221
spectrometer.

3. Results and discussion

Earlier we have shown that adsorption of nirtoxyl radicals TEMPO, TEMPON and NR
is accompanied by sufficiently strong interaction with surface Lewis acid sites of the
catalysts.
The ESR spectra of these nitroxyl radicals adsorbed on the y-Ah03 surface are
shown in Figure I. In the case of TEMPO adsorption, the multiplet spectrum with
hyperfme splitting between nuclei 14N and 27AI is observed (Fig. I,a). This spectrum is
probably attributed to the formation of paramagnetic complexes TEMPO with tbree-
627

coordinated aluminium ions (Ae+) [2]. TEMPON and NR radicals have been adsorbed
on the y-Ah03 surface with stronger basic fragment. The possible structures of
adsorptive complexes of nitroxyl radicals with Lewis acid site are shown below (2):
o

O~ o
~y

o. .
x:)< (2)
:
AI
. 0
AI
:
AI
N-"
'Ph

if.lo "0 if0I "0 if I "0


o
This explains an absence of hyperfine splitting in the case the two latter radicals.
Typical anisotropic triplet spectra of these radicals are presented in Figure I (Spectra
b,c).

x 7.7
a

x3 .6 'Y·Ah03 +TEMPON
b 25%

xl
c 100%

H,G
3230 3290 3350 3410

Figure J. ESR spectra nitroxyl radicals adsorbed on y-AhOl

As followed from obtained results, NR is the most stable nitroxyl radical toward
the protonation (the quantity of adsorbed radical observable in the ESR spectra is about
100%); TEMPO possesses smaller stability (75%) and TEMPON shows the least one
(25 %).
628
Up to the present time, the detailed mechanism of the processes occurring with
nitroxyl radicals on the surface of the catalysts, causing the loss of their paramagnetism,
is unknown. Most probably, the protonation is the fIrst stage of such processes as for
homogeneous systems. Protonation can occur on a Bronsted acid site located near the
adsorption site (i.e. Lewis acid site). According to our results, after decomposition of
adsorbed radicals, diamagnetic products occupy the same sites as before. This data
indicates there is no opportunity for other radicals to be adsorbed on these sites. Just by
this reason, observable concentrations of Lewis acid sites (i.e. adsorptive capacity
toward nitroxyl radicals) are equal as in the case when all adsorbed radicals remain
paramagnetic (NR), or when the signifIcant part of them loses their paramagnetism
(TEMPON).
Thus, we support that adsorbed nitroxyl radicals lose their paramagnetism exactly
at the expense of interaction with sufficiently strong neighbouring surface proton sites
instead of among themselves. Also our hypothesis is confInned by the fact that upon
addition of the next portion of nitroxyl radical its concentration increases but the
percentage observable by the ESR method remains constant (Fig. 2).

c 5
'0.
","!- 4
"b
-.3
oil.

2
AIP/SO.+TEMPON

O+-~--~~--~~--r- __--~~~~~
o 2 4 6 8 10

Figure 2. Dependence of the radicals quantity observed by ESR method on their absorbed quantity

The possible structures of the initial complex fonned at protonation are


represented below (3):

M:
,P"'H)
AI
ell "0
0
(3

AI
ello "0
It was interesting to compare stability of the radicals using the catalysts with
signifIcantly differing acid properties. The results of the measurement for y-Al;03,
629
sulfated, and fluorinated alumina are presented in Figure 3. The highest percent of
radicals appearing in the ESR spectra 'is observed for the y-Ah03 sample while the least
is for the sulfated ones. Thus, in the case of heterogeneous catalysis the catalysts acidity
essentially influences the radical stability toward the protonation processes as for
homogeneous system.

293K

-NR
-TEMPON

a
-100%
--25%

--100%
--30%

H,G
3260 3310 3410

Figure 3, ESR spectra of nitroxyl radicals adsorbed on (a) y-Ah03. (b) AhO~F and (c) Ah03/so1·

4. Conclusion

The use of series of nitroxyl radicals possessing different stability with regard to the
protonation has permitted us to develop a simple experimental technique for
measurement of the quantity of the Lewis acid site-strong Bronsted acid site pairs and
for the estimation of the strength of such acid sites. Lewis acid site acts as the site of the
radical adsorption while neighboring Bronsted acid site accounts for the radical
protonation and the following loss of paramagnetism.
630

Acknowledgment: This work was supported by Russian Foundation for Basic Research
(Grants 00-03-32441, 01-03-06385 and 01-03-06386) and by Programm of Russian
Academy of Science for Support of Young Russian Scientists (Grant 185).

5. References

I. Kuznetsov A.N. (1976) The method o/paramagnetic probe, Nauka, Moscow.


2. Lunina, E.V., Zacharova, M.N., Markaryan, G.L. and Fionov, A.V. (1996) The application pf
paramagnetic complexes of probe molecules for the investigation of the Lewis acidity of
aluminas, Colloids and Surfaces. A: Prysicochemical and Engineering Aspects. 115, 195-206.
3. Lunina, E.V. (1996) EPR spectroscopy of adsorbed nitroxides to investigate catalyst surface,
Applied Spectroscopy. SO, 1413-1420.
4. Volodarskii, L.B., Grigoriev, I.A., Dikanov, S.A., etc. (1988) Imidazoline nitroxyl radicals,
Nauka, Novosibirsk.
5. Bedito, A.F., Timoshok, A.V. and Volodin, A.M. (2000) Formation of tetramethylethylene
radical cation after pentane adsorption on sulfated zirconia, Catalysis Letters 68, 209-214.
POROUS STRUCTURE OF CELLULOSE FIBER WALLS STUDIED WITH NMR
DIFFUSOMETRY

DANIEL TOPGAARD AND OLLE SODERMAN


Physical Chemistry 1
Chemical Center, Lund University, P.o.Box 124, S-221 00 Lund, Sweden
e-mail: daniel.topgaard@jkemJ.lu.se

1. Introduction
Paper has several structural levels (cf. Figure 1). The paper sheet consists of a two-
dimensional network of cellulose fibers having the fonn of flattened tubes. The fiber wall is
built of closely packed micro fibrils consisting of crystalline and amorphous regions of
cellulose. Water in a hydrated paper system has different properties depending on its
position: between fibers, inside the fiber wall or dissolved in the amorphous cellulose
regions. The differences are due to the varying size of the pores that the water occupies and
the different degrees of interaction with the macromolecules in the pore wall. During the
chemical pulping process, pores are created in the fiber wall by the dissolution of lignin and
hemicellulose. These pores are affected by the mechanical treatment, beating, used to
improve fiber flexibility and paper strength [1]. The dry fiber wall is a non-porous solid [2],
which implies that pores exist only when water is present.

Figure I. The structure of paper.

631
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 631-635.
© 2002 Kluwer Academic Publishers.
632

Here we use NMR diffusometry to separate between water in nm-scale and J.lm-scale
pores inside the fiber wall, and NMR relaxometry to separate between water inside and
outside the fiber wall. At low amounts of sorbed water the transverse relaxation time is a
few milliseconds, which means that the Pulsed Field Gradient STimulated Echo (PFG STE)
[3] technique must be used in the measurement of water self-diffusion. The results obtained
with this technique is affected by cross-relaxation, i.e. exchange of longitudinal
magnetization between water and cellulose [4]. It is possible to correct for this exchange if
the rate of cross-relaxation is known. In order to evaluate the diffusion experiment properly,
we quantify the cross-relaxation with the use of the Goldman-Shen (GS) experiment [5].

2. Experimental
The samples were viscose and paper sheets made of bleached kraft pulp fibers. Two kinds
of kraft pulp were used: beaten and unbeaten. NMR experiments were performed at 25°C
on a Bruker DMX 200 spectrometer equipped with a Bruker gradient probe. The amount of
sorbed water was determined from the free induction decay (FID) following a 3.6 J.ls 900
pulse using 0.5 fls dwell time and 4.5 flS receiver dead time. The CPMG decay curve was
sampled at the midpoint of even echoes using a 400 fls delay between the 1800 pUlses.
Self-diffusion measurements were performed with the PFG STE technique using 0.3 ms
gradient pulse length and diffusion times from 10 to lOOms. The gradient strength was
stepped up linearly to a maximum of 9.36 Tim. The cross-relaxation parameters were
measured with the Goldman-Shen pulse sequence as described in ref. [6].

3. Results and Discussion

3.1 FREE INDUCTION DECAY


In Figure 2 we display an experimental free induction decay from a hydrated paper sample.
The signal arising from protons on the solid cellulose decays within some tenths of
microseconds while the water signal lasts some milliseconds. Extrapolation of the two
components to time zero, taken as the midpoint of the 900 pulse, yields an estimate of the
ratio between the number of water and cellulose protons, which can be converted to the
moisture content (MC) = g water I g cellulose [7].
633

1...---r---r----r---r----,
~
§ 0.8
~ cellulose
~ 0.6
~
/'
~ 0.4
I/)
c:
~ 0.2

oL-~--~--~~==d
o 0.2 0.4 0.6 0.8 1
t/ms

Figure 2. Experimental free induction decay a from a paper sample with MC = 0.20 gig.

3.2 CPMG TRANSVERSE RELAXATION


Experimental CPMG echo decay curves obtained on a viscose sample can be seen in Figure
3. At low MC, when all water is inside the fiber wall, the signal decays exponentially with
T2 on the order of a few milliseconds. At high MC the decay is biexponential. The slow
component originates from water outside the fiber wall.

0.01 ....._w..a...-....~..&..Io............__oL.Ioo.a.;&..Iu...&..........,
o 0.04 0.08 0.12
2tn I s

Figure 3. Experimental CPMG echo decay curves from a hydrated viscose sample at varying Me.

3.3 SELF-DIFFUSION AND CROSS-RELAXATION


For free diffusion the normalized signal intensity E depends on the water self-diffusion
coefficient D and the experimental parameter k according the Stejskal-Tanner equation [8]
E =exp(-kD) (I)
The definition of k is k = (rG~2 t, where r is the gyromagnetic ratio, G is the gradient
strength, 0 is the gradient pulse length and t = /). - 0/3 is the effective diffusion time.
634

For gaussian diffusion D is independent of I. When there are barriers for the diffusive
motion on the length scale of the mean square displacement <t> = 2Dt, D is decreasing
with increasing I and can be determined from the initial slope of a Stejskal-Tanner plot of E
vs. k. For water diffusing some tenths of milliseconds in cellulose fibers <t> 1/2 is a few
micrometers. Standard evaluation using Eq. (1) results in D decreasing with I at all Me and
for all samples, implying that pores with dimensions of micrometers would exist even at
very low Me. However, the GS experiment shows that cross-relaxation occurs on the same
time-scale as the I-values used in the diffusion experiments. Longer t leads to more
complete cross-relaxation. An expression analogous to Eq. (1), taking cross-relaxation into
account, was presented in ref. [4]. Since this formula is quite involved it is not reproduced
here. When evaluating the diffusion data using this expression no t dependence was
observed for the kraft pulp samples up to Me = 0.25 gig, and for the viscose sample up to
Me = 0.7 gig. The absence of a t dependence means that the dimensions of the restrictions
is either much larger or much smaller than micrometers. From what is known about the
fiber ultrastructure it is obvious that the pores are on the nm-scale. When D is independent
of t and the dimensions of the pores are much smaller than <t>1I2, the obtained D is a
measure of the permeability of the porous network. We refer to these pores as nm-scale
pores.
For the kraft pulp samples a t dependence of D could be observed above Me = 0.25 gig,
even when the experiment was corrected for cross-relaxation. This is consistent with the
large pores created during the chemical pulping process. With electron microscopy it has
been shown that dissolution of lignin leads to the formation of elongated pores between
bundles of microfibrils [9]. These pores are some tenths of nanometers wide and several
hundred nanometers long. Free diffusion in one dimension on the order of a micrometer is
sufficient to explain the observed I dependence of D. We refer to these pores as !lm-scale
pores.
In Figure 4 we show D vs. Me for the three samples used in this study. Note that D is a
function of t above Me = 0.25 gig for the kraft pulp samples. Only an average value is
shown in Figure 4. The coincidence of the curves for beaten and unbeaten kraft pulp below
Me = 0.25 gig shows that the beating has no effect on the nm-scale pores. The absence of
!lm-scale pores in viscose indicate that they are an effect of the chemical pulping process.

open symbols:
o independent of t
filled symbols:
o decreasing with t

o 0.2 0.4 0.6 0.8 1


9 water I 9 cellulose

Figure 4. Self-diffusion coefficient of water sorbed in cellulose fibers.


635

4. Conclusions
By perfonning an array of NMR experiments, including diffusometry and relaxometry, on
hydrated cellulose samples we can observe water with different properties. Water inside
and outside the fiber wall is separated with the CPMG experiment. Water residing in nm-
scale and J.l.m-scale pores inside the fiber wall is separated with self-diffusion
measurements. The results of this study are summarized in Figure 5. At high MC all types
of water are present. Drying removes water between the fibers until this water disappears at
MCI. Further drying removes water from the J.l.m-scale pores until they are closed at MC2.
For the kraft pulp samples MCI = 0.25 gig and MC2 = 0.7 gig. For viscose MCI = MC2 = 0.7
gig.

bulk water

,"un-pores
nm-pores

o Me
Figure 5. Removal of different water fractions during drying.

5. References
I. Stone, J.E., A.M. Scallan, and B. Abrahamson. (1968) Influence of beating on cell wall swelling and internal
fibrillation. Svensk papperstidning 71(19), 687-694.
2. Stone, lE., A.M. Scallan, and a.M.A. Aberson. (1966) The wall density of native cellulose fibers. Pulp Paper
Magazine Canada 67(5), T263-T268.
3. Tanner, J.E. (1970) Use of stimulated echo in NMR diffusion studies. 1. Chern. Phys. 52,2523-2526.
4. Peschier, LJ.C., et al. (1996) Cross-relaxation effects in pulsed-field-gradient stimulated-echo measurements
on water in a macromolecular matrix. J. Magnetic Resonance B 110,150-157.
5. Goldman, M. and L. Shen. (1966) Spin-spin relaxation in LaF3' Phys. Rev. 144(1),321-331.
6. Topgaard, D. and O. SOderman. (2001) Diffusion of water absorbed in cellulose fibers studied with I H-NMR
Langmuir 17, 2694-2702.
7. Hartley, 1.0., F.A. Kamke, and H. Peemoeller. (1994) Absolute moisture content determination of aspen wood
below the fiber saturation point, Holz/orschung 48, 474-479.
8. Stejskal, E.O. and J.E. Tanner. (1965) Spin diffusion measurements: Spin echoes in the presence of a time-
dependent field gradient, J. Chern. Phys. 42(1),288-292.
9. Hafren, l, T. Fujino, and T. Itoh. (\999) Changes in cell wall architecture of differentiating tracheids of Pinus
thunbergii during lignification. Plant Cell Physiology 40(5), 532-541 .
ALTERNATION IN THE STRUCTURE OF WATER-IN-OIL
MICROEMULSIONS BY THE ACTION OF POLY(ETHYLENE GLYCOL)

N.N. VYLEGZHANINA, B.Z. IDIYATULLIN, YU.F. ZUEV and


V.D. FEDOTOV
Institute of Biochemistry & Biophysics, Russian Academy of Science
POB 30, 420503, Kazan, Russia; e-mail: vylegzhanina@mail.knc.ru

1. Introduction
One of the ways of changing the properties of water media in lipid systems is the
replacement of a part of water by some lipid with a different polarity. For this purpose
water-soluble poly( ethylene glycol) (PEG) with low molecular weight can be used. It
has been found that PEG produces dehydration of the phosphatidylcholine (PC) bilayers
due to water binding to polymer [1]. However in all studies of PEG effects on the lipid
systems the bilayer structures or direct micelles dispersed in water were mostly used.
Until recently the PEG interactions with the lipid-based reverse micelle dispersions
were not studied. The reverse micelles are the spherical droplets with water in a central
core surrounded by a monolayer of amphipatic lipid molecules, which have thier polar
head groups facing the water, and thier hydrophobic tails oriented towards the
continuous oil phase.

2. Purpose of the work


is to study the effect of PEG on the structure of water-in-oil (w/o) microemulsion based
on PC, particularly on the phospholipid-formed monolayer.

3. Samples
To prepare the reverse micelle solutions in oil the following ingredients were used:
phosphatidylcholine (SPC) (mol.wt. 750), tricaprylin C8:0, n-hexanol, poly(ethylene
glycol) 400Da. and 1500, twice distilled water. The main series of the samples has a
volume fraction of dispersed phase ~=0.27 and SPC concentration Cspc =0.32M. In the
control contrl sample without PEG the water amount is equal to 6 wt.% referred to the
overall microemulsion weight. Therewith the water-to-surfactant molar ratio is
Wo==10.4. In all samples containing PEG a part of water is replaced by polymer, the
mass portion of PEG plus water in the microemulsions remaining constant
(Xwater+XPEG == 1, where Xi are the molar fractions of the components) and equal to the
water portion in the control PEG-free sample. To obtain the real size of the reverse
micelles we used the dilute microemulsions with the same Wo=10A but ~=0.086.
The molecules of I-palmitoyl-2-steroyl-(7-doxyl)-glycero-3-phosphocholine
(7-SLPC) were used as spin label. The stock solution (O.OIM) of spin label in oil was
added to oil of the microemulsions to yield a final spin label concentration of 5·1O·4 M
referred to the total sample volume.
637
1. Fraissard aruJ o. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 637-642.
© 2002 Kluwer Academic Publishers.
638

4. Methods of investigation
Fourier-transfonn pulsed-gradient spin-echo (FT PGSE) lH nuclear magnetic resonance
(NMR) method allows to measure the self-diffusion coefficients of the microemulsion
components (Fig.l) and to obtain an infonnation about micelle size and distribution of
the microemulsion compounds between different compartments. The measurements
according to procedure described in [2] were perfonned on a modified Tesla BS 576A
100MHz) and Tesla BS 587A (80MHz) NMR spectrometers. All self-diffusion
experiments were carried out at 30°C.
a H,o
o PEG
• SPC
• T ricaprylin
X Hexano/
;?
0-

i
C
.!I!
1

c:
o
~
~
3l 0, 1 -I--........-+~-+---+--I---<
0,00 0,Cl.1 0,D2 0,03 0,04
molar fraction X. EG
Figure 1. Self-diffusion data as afunction of molar fraction a/polymer, XP£G, in the PEG-mixture
(XP£G+X.",., =1).

ESR spin label method allows to obtain an infonnation about dynamic


stmcture of the phospholipid interface separating the internal polar core of the reverse
micelles and the external organic phase. To characterize the lipid chain dynamics we
used the maximum hyperfme (hf) splitting, 2Amax. in the ESR spectra since this
parameter is sensitive both to the amplitude and to the rate of the PC molecule spin-
labeled chains [3].
ESR spectra of the microemulsions spin-labeled by 7-SLPC were registered on
radio spectrometer RE 1306 (USSR) over the temperature range -lO++70°C. At high
temperatures ESR spectra of all samples consist of three line with different width and
intensity typical for rapid anisotropic motion of paramagnetic moiety of spin label. As
the temperature decreases the outer wide peaks (OWP) ·arise in the spectra. Their
appearance implies that the motion of the SPC acyl hydrocarbon tails with nitroxide
moiety becomes sufficiently slow and does not average the hf interaction anisotopy.
When the temperature is lowered to a value Ttr at which the OWP appear in ESR spectra
there is a sharp increase of hf splitting 2Amax. The smaller the local mobility of the SPC
lipid tails in the micelle interface layer the higher temperature Ttr and the greater the
2Amax-value.

5. Results and discussion


In the first place, let us consider and discuss in detail the data and a proposed model for
PEG400. In the absence of PEG the system studied represents a dispersion of the
639

reverse micelles or water droplets covered by the layer of SPC molecules and floated in
the oil medium. If practically all SPC molecules form the reverse micelle envelopes the
surfactant self-diffusion coefficient determines the diffusion motion of these micelles.
The reverse micelle size have been obtained by a procedure on the basis of the
modified Stokes-Einstein equation [4]:
D kT
=---
1- fjJ 61!1]RM

where D is the self-diffusion coefficient of the reverse micelle, Y] is the oil medium
viscosity, RM is the micelle radius. We estimated that in the absence of PEG (molar
fraction of polymer XPEG=O) the droplet radius was equal to 46A. The replacement of
minimum amount of water by PEG resulted in a decrease of micelle radius down to 40A
which remains constant up to XPEG =O.022. At XPEG =O.022, RM reduced up to 3lA.
According to the simple geometric model it is possible to expect that a
decrease of the micelle radius (an increase of the SPC layer curvature) results in an
increase in the free volume for the motion of the SPC hydrocarbon tails. However ESR
data show the opposite tendency (Fig.2). We suppose that it is a result of the changes in
SPC packing in the reverse micelle interface and try to find the explanation of this
phenomenon.

45
()
40
f " I\.
,
.,. ,
0

I- 35 1'..
~
.aell 30
8.
E 25
.9l
20
0,000,01 0,020,030,040,05
molar fraction XpeG
Figure 2. Dependence of the temperature T" on the PEG amount in the microemu/sions.

The self-diffusion coefficients of water and PEG are greater than that of SPC
(Fig.l). It means that not all water and PEG molecules are present within the reverse
micelle internal core. A part of them is dissolved in the oil phase. In the framework of
the two-site model, which is successfully used for microemulsions [5] and on the basis
of self-diffusion coefficients, it is possible to estimate the fractions of substances
present in different phases of the system. In these calculations we considered the
following states of PEG and water. PEG bound - a part of PEG within the reverse micelle
water pool; PEG lTee - a part of PEG in the oil medium; Wbound - a water within the
reverse micelles; W lTeel - a water in the oil medium; W ITee2 - a water bound to the "free"
PEG molecules. The number of water molecules bound by each of the ether-linked
oxygen atoms in PEG400 is equal to 1.8 [I]. As a result, the number of water molecules
640

bound to every free PEG molecule is equal to 16. The data characterizing a
redistribution of PEG and water between the reverse micelle core and the oil phase are
depicted in Fig.3 and 4.
- -r - -

! 0,04
~ I ~r
()
c: 0,03
7
~
~ 0,02
I(
§
/
8 0,Q1

o
o
- .J
0,01 0,02 0,03 0,04 0,05
molar fraction X PEG
Figure 3. Molar concentration of PEG present in the oil phase of the microemulsions.

a) Stage I Stage 2

tdhydrocarbon chains
-"- polar head

Figure 4. Scheme of changes in the reverse micelle size (a) and the SPC molecules orientation (b) on
substitution of water by PEG.

We suggest the following mechanism of the PEG400 acting in the solution of


the SPC-based reverse micelles. When in the microemulsion the PEG molar
concentration, XPEG , increases up to 0.022 (the water-to-PEG wieght ratio 2: 1) its
concentration in the oil remains constant (Fig.3) and a major part of polymer locates
inside the reverse micelle core together with water. It results in a slight growth of the
reverse micelle core size, but therewith the overall radius of the micelle slightly
decreases and the water-to-PEG molar ratio inside micelle strongly falls to the absolute
PEG hydration value (Fig.4).
This reduction of the overall radius under the growth of micelle core is
possible only when the thickness of the SPC interface is lowered through a gradual
enlargement of lipid-tail tilt angle relative to a normal to the reverse micelle core
surface. The ability of PC molecules to change the hydrocarbon chains tilt with respect
641

to the surface of the polar head groups is associated with specific features of their
molecular structure [2]. In the PEG-free samples the SPC hydrocarbon tails are disposed
aloqg the normal to the core surface, at XPEG =O.022 the tilt angle is about 44°C and at
maximum PEG concentration it increases up to 61°C. This, in turn, has to result in the
reduction of chain mobility, which is just registered by the ESR measurements.

45++-"'-+-'-........-+-"'-+-1
. -e-PEG400
~ 40 - 0 - PEG1500

~:b 35++--f--,y:.....;...:::o-__r/---;---4-l

1:++I"--I--'--I--'-+-'--I--1
20~~--~~--~

0.0 0,1 0,2 0,3 0,4


Z = [~~..().] / [~O]
Figure 5. Dependence 0/ the temperature T". on the PEG 's-amounts in the microemulsions.

The internal core dimensions determine the quantity of PEG, which can enter
the reverse micelle. When the greatest possible quantity of PEG dissolves inside the
inner core, the polymer amount within the oil phase will progressively increase as the
overall PEG content in the microemulsion increases (Fig.3). Some water molecules can

PEG 1500
-o-PEG
-6.-SPC
0,1 +-+--;....-+-;....-;.--;....+-~~
0,0 0,1 0,2 0,3 0,4
Z = [~~..().] / [~O]
Figure 6. Self-diffusion data as a/unction o/parameter Z/or the PEG's-microemulsions.

be expected to come out of the reverse micelle water pool and to bind to the PEG
molecules in oil. As a consequence, the reverse micelle radius reduces and free volume
in which the SPC hydrocarbon chains move expands. The redistribution of water and
PEG between the reverse micelle core and the oil phase induces significant changes in
the core polarity, which, in turn, causes the alteration in the phospholipid layer.
642

Let us see what happens when we use PEGlSOO instead PEG400. These two
polymers have no fundamental differences from the water-binding point of view [1] . In
the equal weight portions both polymers have approximately same quantities of
monomer units (-CH2-CH2-0-). For comparison of them, we use a ratio of molar
concentrations of monomer units and water in the microemulsions Z = [-CH r CH 2-O-] /
[H 20]. The main difference between two PEG's used is their size. The maximum linear
size of PEG400 is about 3sA and for PEG IS00 - 13sA. In the PEG-free microemulsions
the reverse micelle water core radius is equal to 21A. Flexible as this polymer may be, it
is rather difficult to imagine that more then one PEGlSOO-molecule can enter the inner
core. This is consistent with value of Z""O.OS. One can see from Fig.S that for PEG ISOO
the TIT -temperature reaches its maximum approximately at this concentration.
Therefore, it is evident that for PEGlSOO we see the same process as for PEG400
belowX PEG =O.022, namely, the changes in phospholipid packing and the enlargement of
lipid-tail tilt angle relative to a normal to the lipid polar head surface. The only
distinction between PEG's is that for PEG1S00 these alterations finish at lower
concentration than for PEG400, since the reverse micelle internal cores are saturated in
the average by one PEG1S00-rnolecule for one reverse micelle.
The NMR self-diffusion data demonstrate the same picture (Fig.6). One can
see that below Z""O.OS the self-diffusion coefficients Dspc"",Dmicelle"",DPEG. It means that
all PEG lS00-molecules are inside the reverse micelles. But above this concentration
DPEG>Dmicelle' Most likely, the reason is that additional PEG1500 arranges within the oil
phase. In our opinion, the apparent difference in the PEG's-behavior at low
concentrations may be the results of some reasons: firstly, there is significant difference
in their self-diffusion coefficients (for PEGlS00 DPEG"",Dmicelle), secondly, an exchange
between bound and free states foe PEG 1500 has more difficulties than for PEG400. As
for water in the microemulsions with PEG 1500, we did not fix the marked changes in
its diffusion behavior in comparison with PEG400.

This work was supported by the Russian Foundation for Basic Research (grant no.99-03
-32037).

6. References
I. Tilcock, O.P.S. and Fischer, D. (1982) The interaction of phospholipid membranes with poly(ethylene
glycol) vesicle aggregation and lipid exchanege, Biochem. Biophys. Acta 688, 645-652.
2. Zuev, Yu .F., Vylegzhanina, N.N., Fedotov, V.D., Idiyatullin, Z.Sh. and Archipov, V.P. (2000) Poly
(ethylene glycol) and phospholipid packing in the structure of reverse micelles, App/. Magn. Reson. 18,
275-288.
3. Fedotov, V.D., Vylegzhanina, N.N., Altshuler, A.E., Shlenkin, V.I., Zuev, Yu.F. and Gat"ti, N. An
electron spin resonance study of the soy bean phosphatidylcholine-based reverse micelles, Appl. Magn.
Reson. 14,497-512.
4. Fedotov, V.D., Zuev, Yu.F., Archipov, V.P. and Idiyatullin, Z.Sh. (1996) Self-diffusion in
microemulsion and micellar size, Appl. Magn. Reson. 11,7-17.
5. Fedotov, V.D., Zuev, Yu.F., Archipov, V.P., Idiyatullin, Z.Sh. and GaTti, N. (1997) A Fourier transform
pulsed-gradient spin echo nuclear magnetic resonance self-diffusion study of microemulsions and the
droplet size determination , Colloid and Surfaces. A: Phosphochem. and Engin. Aspects 128,39-46.
IN SEARCH OF THE NATURE OF THE ANISOTROPIC DIFFUSION IN
NERVOUS TISSUE - MR MICROSCOPY OF THE EXCISED RAT SPINAL
CORD

W.P. WfiGLARZ, A. HILBRYCHT', D. AOAMEK2, J:PINOEL, A. JASINSKI.


Institute of Nuclear Physics, KrakOw, Poland
JAcademy of Mining and Metallurgy, Krakow, Poland
1Collegium Medicum, Jagiellonian University, Krakow, Poland,

Abstract. Anisotropic diffusion in the excised rat spinal cord saturated with 0.9%
saline was investigated using MR microimaging with b-values up to 4000 s/mm2 . Two-
exponential transversal diffusion decay was found in white matter (WM) and gray
matter (GM). In contrast, the longitudinal diffusion is two- and monoexponential in
GM and WM respectively. Slow transversal components have similar values in WM
and GM. It was concluded that observed diffusion ~nisotropy in the spinal cord is
entirely due to presence of the slow transversal component, arising from restricted
diffusion.

1. Introduction

Diffusion weighted MR imaging of the nervous tissue undergoes a rapid development


during last decade due to high sensitivity of the anisotropic diffusion to structural
changes in the tissue. Monoexponential diffusion model was used for data analysis
until it was found that in the brain tissue diffusion measured with high b-values is non-
exponential [1]. Two components of the diffusion were attributed to extracellular and
intracellular water. Considerable effort was done in the following years to characterize
non-exponential diffusion in the brain and the spinal cord tissue and to correlate its
components with physiological water pools [2-5]. Serious difficulties were noticed
with finding direct correspondence between component fractions and known
composition of the brain tissue. This discrepancy was explained by influence of
exchange between water reservoirs, but no undoubted model was proposed. In other
papers the role of restrictions was pointed out [6,7]. The diffusion in bovine optic nerve
was explained entirely on the basis of the analytical model of the restricted
diffusion [6].
The spinal cord represents tissue of intermediate complexity between single nerves
and brain tissue. It consists of white matter (WM) forming its outer part, and gray
matter (GM) inside it. WM contains highly oriented axons of cylindrical symmetry
parallel to each other, surrounded by impermeable myelin sheaths. The cells in GM
form a much more isotropic structure.
The aim of this work was to use MR microscopy for detection and quantification
of non-exponential anisotropic diffusion in an excised rat spinal cord, and to correlate
the results with structure dominating in spinal cord WM and GM.
643
1. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Interface Science, 643-648.
© 2002 Kluwer Academic Publishers.
644

2. Materials and Methods

Male Wistar rats were anaesthetized and the laminectomy at the Th9 spine level was
performed. Animals were intracardiac perfused with 4% formaline. A part of the
thoracic spine was excised and kept in formaline. Before measurements, the spinal cord
was intensively rinsed in 0.9% saline for 4 hours to remove formaldehyde which
significantly shortens T2• This increases SIN ratio which is an important factor,
especially when collecting data for 2D images. Orientation of the long axis of the
spinal cord was set approximately along slice imaging direction. Diffusion
measurements were performed on an 8.4 T MR microscope, using standard DWSE
imaging sequence capable to apply diffusion gradients in any direction [8]. Two sets of
lD profiles (with phase imaging gradient set to 0), for slice of thickness 400 /lm,
perpendicular to spinal cord axis (transversal plane), were measured for 18 values of b-
factor ranging from 0 to 4000 s/mm2• The first set was done with diffusion gradients
parallel to spinal cord axis to measure longitudinal apparent diffusion coefficient
(longitudinal ADC - D1oog), while the second one was done with diffusion gradients
perpendicular to spinal cord axis, along the phase direction (transversal ADC - Dtrans).
The phase direction was choosen for measurements of Dtrans to avoid influence of the
read imaging gradient for the corresponding b value [8]. Diffusion time (d) was set to
26 ms, and echo time (TE) was 65 ms. Length of the diffusion gradient (0) was 7 ms,
and the gradient amplitudes were incremented to give the appropriate b-value. The
number of acquisitions was 100 and the repetition time was 2s. The data were analyzed
for two-exponential decay using non-linear least squares fittin~. 2D MR images
(128x128) were acquired for six b values (up to 3240 s/mm). The number of
acquisitions was decreased to 30 to reduce the total time of experiment (over 2 orders
in magnitude longer than for ID measurements). Other parameters were set as in ID
experiment. The experimental setup and data analysis for 1D measurements are
identical to that we used for measurements of the diffusion in the spinal cord in
formaline [9].

3. Results

Fig 1. presents image of the spinal cord with orientation of the phase imaging axis
shown. The spinal cord was oriented in a way allowing for easy distinction of the
regions in profile corresponding exclusively to saline solution from that corresponding
mainly to spinal cord tissue. The series of diffusion weighted profiles with diffusion
gradients applied in the phase direction (Fig. 2a) and the slice direction (Fig. 2b) show
characteristic differences in regions corresponding predominantly to spinal cord tissue
(i.e. approximately between 1.8 and 4.5 mm). In both cases diffusion in the spinal cord
is smaller than in saline solution but this difference is much larger for direction
perpendicular to spinal cord axis (transversal diffusion). Numerical analysis of the data
show that for the range of b values used, transversal diffusion decays in spinal cord are
non-exponential, in contrast with purely exponential longitudinal diffusion in the WM
(range 2.1-2.4 mm on Fig. 2b) as well as diffusion in saline. Longitudinal diffusion in
the region of the spinal cord where signals from WM and GM are not resolved in the
profiles (i.e. 2.7-4.5 mm) shows small deviation from exponential decay, which is
645

much smaller than in the transversal diffusion. In the Fig. 3 results for WM region of
the spinal cord are shown. Transversal diffusion is described by two components: fast
one with I/,rans = (1.3 ± 0.2).10.3 mm2/s comprising 53% of the total signal and the
slow one If,rans = (1.0 ± 0.4)-10:4 mm2/s with relative abundance equal to 47 %. The
longitudinal diffusion component is equal to D/ong = (1.10 ± 0.08)-10. 3 mm2/s.
Diffusion coefficient measured for saline is (1.50 ± 0.05) pO·3 mm2/s in both
directions. Bulk water diffusion coefficient measured at the same conditions for
calibration purpose was (1.90 ± 0.05),10.3 mm%, which agrees with literature value
for water diffusion coefficient at 20 °C [3].

Fig. I. MR Microimage of the excised rat spinal cord. Inner sample tube diameter is equal to 5 mm. Phase
and read image directions are shown.

4,Oxl0·

..... 3,5xl0·
::l
.!i 3,oxl0·
Q)
'0
:E 2,5xl0·
a.
E 2,oxl0·
<
(ij
c: l,5xl0·
Ol
Ci5 l,oxl0·

5,oxl0 s

0,0
0 2 3 4 5 6 7 0 234 567
Position along read axis [mm)

Fig. 2. The set of diffusion weighted MR profiles (with phase imaging gradient set to 0) of the excised rat
spinal cord for diffusion gradients in (a) phase and (b) slice directions.
646

o
o

Spinal Cord

DIOng
0.9% Saline

o 500 1000 1500 2000 2500 3000 3500 4000

b[s1mm 2]

Fig. 3. Amplitudes of the signal from different regions of the profiles as a function of b. Data points for
spinal cord were averaged within range 2.1-2.4 mm (WM), while for saline within ranges 1.3-1.8 mm and
4.8-6.1 mm of the corresponding profiles. The lines were fitted using one or two exponential decays with
parameters indicated in the text.

In the table 1 the fitted parameters of the diffusion decays for three different
regions of the spinal cord are shown. For longitudinal diffusion only mono exponential
results were fitted with confidence in all three regions. For regions II and III small
deviations from exponential decay were observed indicating presence of the slow
component. However, this deviation was too small to fit the slow component with
confidence. In general, the longitudinal and fast transversal diffusion have similar
values while the slow transversal component is about one order in magnitude slower.

Region Longitudinal Transversal diffusion


[mm] diffusion ---F=a-s-t-----=Sl=-o-w---
1O-3[mm2/s] 10-3[mm2/s] 1O-4[mm2/s]
I (2.1-2.4) 1.10±0.08 l.3±0.2 (53%) 1.0±0.4 (47%)
II (3.0-3.4) 0.99±O.07 l.3±0.2 (52%) 1.0±0.4 (48%)
III(3.9-4.2) 1.03±O.07 1.2±O.2 (54%) 1.0±OA (46%)

Tab. I Parameters of the diffusion decays for three different regions of the spinal cord.

Fig. 4 presents microimages of the investigated spinal cord, with the diffusion
gradients oriented perpendicularly and parallelly to the spinal cord axis, for different
values of the diffusion gradient amplitudes (different b-values). The difference between
transversal and longitudinal diffusion in WM and GM is well recognized for high 'b-
647

values (above -1000 s/rrun2 ), where the slow ADCs are dominant. The slow
transversal AOC is present in WM and GM, while the slow longitudinal one is present
only in GM.

Fig. 4. The set of diffusion weighted MR images of the excised rat spinal cord. Diffusion gradients oriented
in transversal (upper row) and longitudinal (lower row) direction, with b-values equal to: 0, 360, 810, 1440,
2250 and 3240 slmm2, respectively.

4. Discussion and conclusions

The 10 micro imaging of appropriately oriented spinal cord enables collection in


relatively short time sufficient number of good quality data to fit more complex decay.
However, some details like resolution between WM and GM are possible only on full
20 images. The time required for the 20 experiment is about two orders of magnitude
longer, so maximizing SIN is necessary. Rinsing out formaldehyde which shortens T2
with 0.9% saline increases observed signal and assures that tissue is measured in
conditions similar to physiological ones. For echo time we used (i.e. 65 ms) increase of
the signal amplitude was -2 [10]. Use of the saline rise however question concerning
the correspondence with results obtained earlier for spinal cord in formaline [9].
Therefore, the experimental setup for 10 measurements was the same in both cases.
The obtained results are similar in general: presence of the significant deviation from
the mono-exponential decay is observed for transversal diffusion while the longitudinal
one is practically mono-exponential. The fast transversal component and the
longitudinal component have similar values while the slow one is one order slower.
However, fraction of the slow component for tissue in saline (-47%) is higher than in
formaline (-30%) [9).
In general, the apparent diffusion in the spinal cord tissue measured with b-values
up to 4000 s/rrun2 may be described by a two-exponential decay with "fast" and "slow"
components. The fast component comes from the water molecules experiencing
approximately isotropic diffusion as may be concluded from similar values of the fast
component in both directions. The average distance of free diffusion for this
component is -7~m, which is comparable to typical axon's diameter, and is much
smaller than the axon's length [7). Thus the slow component observed in the
transversal diffusion in WM is due to fact that in the direction perpendicular to the
spinal cord axis the water diffusion is restricted to much higher extent than in the
longitudinal direction. From 20 microimages the difference between diffusion in WM
648

and GM is easily seen for b-values above -1000 slrrun2, where the slow diffusion
component is dominant (Fig 4.). From the presence of the transversal as well as
longitudinal slow component in GM it may be concluded that water diffusion in GM is
restricted in both directions. No significant contrast between WM and GM for slow
transversal component suggests that the mean distance between restrictions in GM is
similar to that in the direction perpendicular to the axon's axis in WM. Due to the
tissue structure the diffusion in WM is highly anisotropic - fairly free along tubules
while limited across them. In contrast, the lack of the dominant long cells structures in
GM causes restriction of water diffusion in both directions. The diffusion in GM may
also have some degree of anisotropy, but much smaller than in white matter.
The above results agree with the conclusions obtained from the investigations of
the bovine optic nerve [6], where restrictions were shown to produce non-exponential
diffusion decay. In the view of difficulties with binding diffusion components to
different water compartments within nervous tissue [I -5] it seems reasonable to
conclude that the observed non-exponential diffusion decay may be explained entirely
by the restrictions experienced by the water molecules diffusing across cellular
barriers. The high diffusion anisotropy in the spinal cord WM, which is the result of
such restrictions in the direction perpendicular to the axon's axis, is best visualized by
diffusion imaging with b-values above 1000 slrrun2• This observation is of great
importance for utilizing diffusion anisotropy imaging for monitoring the spinal cord
damage and recovery.

Acknowledgments
This work was supported by the State Committee for Scientific Research (KBN) of Poland under grant No: 2
P03B 102 18.

References
J. Niendorf T., Dijkhuizen RM., Norris D.G., van Lookeren Campagne M. and Nicolay K. (\ 996)
Biexponential Diffusion Attenuation in Various States of Brain Tissue: Implications for Diffusion-
Weighted Imaging, Magn. Reson. Med. 36,847-857
2. Assaf Y. and Cohen Y. (1998) Non-mono-exponential attenuation of water and n-Acetyl Aspartate
signals due to diffusion in brain tissue, J. Magn. Reson., 131,69-85.
3. Mulkern RV., Gudbjartsson H., Westin C-F., Zengingonul H.P., Gartner W., Guttmann e.RG.,
Robertson R.L., Kyriakos W., Schwartz R, Holtzman D., Iolesz F.A. and Maier S.E. (\999) Multi-
component apparent diffusion coefficients in human brain, NMR Biomed., 12,51-62
4. Assaf Y and Cohen Y. (2000) Assignment of the Water Slow-Diffusing Component in the Central
Nervous System Using q-Space Diffusion MRS: Implications for fiber tract Imaging, Magn. Reson.
Med., 43, 191-199
5. Inglis B.A., Bossart E.L., Buckley D.L., Wirth III E.D. and Mareci T.H. (2001) Visualization of Neural
Tissue Water Compartments Using Biexponential Diffusion Tensor MR!, Magn. Reson. Med., 45, 580-
587
6. Stanisz G.I., Szafer A., Wright G.A. and Henkelman R.M. (1997) An analytical model of restricted
diffusion in bovine optic nerve, Magn. Reson. Med. 37, 103-11 J.
7. Ford J.e., Hackney D.8., Lavi E., Philips M. and Patel U. (1998) Dependence of Apparent Diffusion
Coefficients on Axonal Spacing, Membrane Permeability, and Diffusion Time in Spinal Cord White
Matter, J. Magn. Reson. Imag., 8, 775-782.
8. Krzyzak A.T., (2000) Investigations of Water Dynamics in Biological Systems using Diffusion Tensor
Imaging. PhD Thesis, Krak6w,
9. Weglarz W.P., Jasinski A., Adamek D., Pindel I., Kulinowski P., Hilbrycht A., Sk6rka T., Sulek Z. and
Szybinski K. (2001) MR Microimaging of the anisotropic multicomponent water diffusion in the excised
rat spinal cord, Mol. Phys. Rep., 33, 216-219.
10. Pindel J., Jasinski A., Weglarz W.P., Krzyzak A.T., Adamek D. and Mareci T.H. (2001) Influence of
formaline fixation on relaxation and diffusion in spinal cord samples, Mol. Phys. Rep., 33, 200-203.
STRUCTURAL FACTORS IN MICELLAR CATALYSIS:
NMR SELF-DIFFUSION STUDY

YU.F. ZUEV\ B.Z. IDlY ATULLIN\ V.D. FEDOTOV\


A.B. MIRGORODSKAYA 2, L.YA. ZAKHAROVA2,
L.A. KUDRY AVTSEV A2
J Kazan Institute of Biochemistry & Biophysics, Russian Academy of
Science
420503, POB 30, Kazan, Russia; e-mail:zuev@mail.knc.ru
2 A.E. Arbuzov Institute of Organic and Physical Chemistry, Russian
Academy of Sciences
420088 Kazan, Akademik Arbuzov Str., 8, Russia

1. Introduction

The use of surfactants is one of the most important ways affecting reactivity of organic
compounds. The micellar effect in reaction kinetics is mainly due to increase in
concentration of reagents in the micellar pseudophase which in tum is accompanied by
changes in the microenvironment, solvation and orientation of reacting species. The
distribution of reactants between solvent and micelles is governed by their polarity and
hydrophobicity .
The microemulsions are transparent, isotropic, thermodynamically stable
dispersions of oil and water, stabilized by surfactant and co-surfactant. The high
solubilizing ability of microemulsions and dramatically large area of oil/water interface
provide the effective contact and interaction of reagents characterized by different
solubility in water and oil. It makes possible to use successfully such systems as the
medium for various chemical reactions [1-3].
In the present work we show two characteristic examples of compartible study of
microemulsion structure and chemical kinetics. Several reactions of hydrolysis of
phosphorus acids esters were studied in microemulsions and micellar solutions based on
cationic surfactants.

2. Experimental

2.1. METHODS

2.1.1. NMR self-diffusion


Fourier-transform pulsed-gradient spin-echo (FT PGSE) IH nuclear magnetic resonance
method [4] allows to measure the self-diffusion coefficients of the microemulsion
649
J. Fraissard and O. Lapina (eds.), Magnetic Resonance in Colloid and Inteiface Science, 649-654.
© 2002 Kluwer Academic Publishers.
650

components and to obtain an information about micelle size and distribution of the
microemulsion compounds between different compartments. The measurements were
performed on a modified Tesla BS 576A (lOOMHz) and Tesla BS 587A (80MHz) NMR
spectrometers equipped with home-built field-gradient units producing a field gradient
up to 50 G cm" [5,6]. All self-diffusion experiments were carried out at 30°C.

2.1.2 . Reaction kinetics


The reactions were monitored by observing the p-nitrophenolate-anion absorption using
'Specord M-400' spectrophotometer equipped with temperature controlled cell holders.
The kinetic data obtained were treated in terms of pseudo phase model [1-3].

3. Results and discussion

3.1. CETYLPYRIDINIUM BROMIDE-BASED MICROEMULSIONS

In cetylpyridinium bromide (CPB)-based microemulsion we studied the hydrolysis of


O,O-(bis-p-nitrophenyl)methylphosphonate (1) in the presence of primary n-
alkylamines depending on the structure of the reaction medium, namely, an oil-in-water
microemulsion of the following composition: cetylpyridinium bromide (CPB), 9.5; n-
butanol, 9.1; n-hexane, 2.0; and water, 79.4 wt.%. Water containing 98% 0 20 (Ferak)
was used in the NMR measurements.

TABLE I. Self-diffusion "Coefficients of microemulsion components (30±O.S"C) and the second-order rate
constants of I in the presence of primary amines

DCPB D Bu Ow k2
Amine
(J 0·lI m2s·') (J 0·9 m2s") (J 0·9m2s") (dm 3 mor' s·')
3.6 0.46 1.74
Butylamine 4.2 0.44 1.74 I.3
Octylamine 4.1 0.44 1.74 0.85
Decylamine 4.0 0.46 1.77 0.55
Dodecylamine 3.2 0.51 1.84 0.56
Cetylamine 2.9 0.5 1.83 0.55
Octadecylamine 3.0 0.5 1.85 0.58

The most intense 'H NMR signals were used for the self-diffusion measurements
of the microemulsion compounds (TABLE I). The water and butanol self-diffusion
coefficients were determined with the help of signals from water protons and a-CH2
protons of butanol, respectively. The diffusion decay of the (CH 2)n protons signal
651

showed ·that CPB and hexane molecules are characterized by the same translational
mobility.
If the total surfactant concentration in ME is much higher than the critical micelle
concentration, the surfactant diffusion is close to the drop diffusion, Dsurf ~ Ddrop [4] .
Thus, the effective drop radius R.trop can be estimated by the Stokes-Einstein equation.
We found R.trop ::::: 3 nm in the absence of amines by extrapolation of the concentration
dependence of CPB self-diffusion coefficients to the infinite dilution.
In oil-in-water microemulsions, a co-surfactant and a hydrocarbon take part in the
formation of drops along with surfactant. Our self-diffusion data suggest that this is true
for hexane: the diffusion decay of the (CH 2)n proton signals from hexane and CPB
cannot be divided. Thus, CPB and hexane form joint structural aggregates.
Nevertheless, the self-diffusion coefficients of butanol in microemulsion are higher than
Ddrop ' This may result from the partial presence of butanol in the bulk phase. According
to the two-site model (fast exchange between two states on the NMR time scale) the
observed self-diffusion coefficient for butanol is the average from its diffusion in the
free state in bulk medium and bound with drop. From our results it follows that disperse
phase includes only 34% of butanol presented in amine-free microemulsion. The
analysis of water self-diffusion coefficients shows that 6.5 % of total water is included
in the drops. Evidently it is the result of water binding to the polar surface of drops.

0.4
~ .
OJ
.--------
0.2

0.1
....~ . .

0.0
2 4 6 8 10 12 14 16 18 20
n

Figure I. Weight fraction of (I) butanol and (2) water included in the drops ofmicroemulsion as afuction of
carbon-chain length (n) in the amine hydrocarbon residue. Dotted lines show the corresponding values for
aminejree microemulsion.

Amines may act as co-surfactants and compete with butanol at the interface of
drops. According to self-diffusion data (TABLE 1, Figure 1), the amines examined can
be divided into two groups according to their influence on the microemulsion structure.
An increase in self-diffusion coefficients of CPS (approximately by 15%), i.e., a
652

decrease in the drop size, and the increase in the weight fraction of butanol included in
disperse phase are the characteristics of the first group (butylamine and octylamine).
Starting from decylamine, the changes in the size and composition drops take place: the
radius of drops increases, and butanol and water are forced out from the disperse phase
into the bulk one. The hydrophobic amines tend to incorporate with droplets and to
increase their size. Then the weakening of drop hydration may result in decrease of the
drop surface charge density because of the incorporation of neutral amine molecules
between the charged CPB species. Probably, the surface dehydration of micelles by
hydrophobic amines leads to the separation of a nucleophile (water activated by amines)
and an oil-soluble substrate and hence results in a decrease in the rate constant of
hydrolysis for 1 (TABLE 1).

3.2 . AQUEOUS SOLUTIONS OF DODECYL PYRlDINIUM BROMIDE

Correlation between salt induced structural transitions of the dodecyl pyridinium


bromide (DPB) micelles and their catalytic effect on the basic hydrolysis of O-ethyl 0-
p-nitrophenyl chloromethylphosphonate (2) were studied in 0.05 M DPB aqueous
solution.

0.16

0.12

0.08
.,
'"
J
0.04

0.00

-2.0 -1.6 -1.2 -0.8 -0.4 0.0 0.4


IgC"h

Figure 2. The dependence of the observed rate constant of the basic hydrolysis of2 (0.005 M NaOH) in the
DPB (0.05 M) micellar solution on logarithm of salt concentration, 25°C (\ -KCI, 2-KBr, 3-NaSal).

One of the most important effects in micellar catalysis is the influence of the salt
induced micellar transitions. The addition of electrolytes results in a change in such
micellar characteristics as critical micellar concentration (cmc), aggregation number,
degree of counter-ion binding, micellar shape, etc. At a definite electrolyte
653

concentration gradual changes in the 'above parameters turns into the sharp changes,
which is revealed by a break in the 'property' vs.'lg Csalt ' plot. These threshold
electrolyte concentrations are associated with sphere-to-rod micelle transitions.
In Figure 2 the kinetic data are presented, which demonstrate a decrease in the
observed rate constant with addition of several salts. According to the pseudo phase
model such inhibition can results from a decrease in the micellar surface potential with
increase in counter-ion concentration which in tum is responsible for weakening of the
electrostatic attraction of hydroxide-ions with the positively charged micelles where the
reaction takes place. In addition the displacement of the substrate by counter-ions from
the micellar surface is probably observed.

70 70

60 60

50 50

40 40

0
'0;
.... 30 30
Vl •
><
'" 20 20

10 10

0 0

0.01 0.1
CNllsal

Figure 3. The dependence of the DPB micelle axis ratio on the NaSal (M) concentration, COPB = 0.05 M.

The micelle size and shape were analyzed on the base of the DPB self-diffusion
coefficients obtained from diffusion decay of proton signals from the (CH 2)n groups in
NMR spectra. The use of Stokes-Einstein equation gives the effective radius (R) of the
DPB micelles in the absence of NaSal equal to 1.54 nm, which is very close to the DPB
molecule length calculated according to parameters of its bonds (1.55 nm), thus
indicating spherical shape of DPB micelles without added NaSal. In the presence of
NaSal we received improbably great R values that is why we used DPB self-diffusion
coefficient to obtain the axis ratio P=2B/2A examining the DPB micelles as prolate
ellipsoids. Here 2B is the length of ellipsoidal micelle and 2A is taken equal to twice
radius of the spherical micelle in the absence of NaSal. The results presented in Figure 3
654

show that an increase in the Sal' concentration results in the changes of the micellar
shape from spherical to intermediate sphere-cylindrical (P=I+S) at the first stage and to
strongly stretched above CNaSal-0.05 M.
So, the NMR self-diffusion data give evidence for the structural transition in the
DPB micelles near the 0.05 M NaSal concentration. Comparing the data in Figures 2
and 3 one can see, that Ccr values in Figure 2 are in the concentration range of NaSal
corresponding to the sharp structural alteration in the OPB micelles, thus confirming
their relationship.

This work was supported by the Russian Foundation for Basic Research (grants nO.99-
03 -32037 and 01-03-33222).

4. References

I. Mirgorodskaya, A.B., Kudryavtseva, L.A., Zuev, Yu.F., Archipov, V.P. and Idiyatullin, Z.Sh. (1999)
Catalysis of hydrolysis of phosphorus acids esters by the mixed micelles of long-chain amines and
cetylpyridinium bromide, Mendeleev Communications 196-198.
2. lakharova, L.Ya., Kudryavtsev, D.B., Kudryavtseva, L.A., Konovalov, A.I., luev, Yu.F., Vylegzhanina,
N.N., Zakhartchenko, N.L. and Idiyatullin, l .Sh . (1999) The influence of sodium salicylate on the
micellar rate effect and the structural behavior of the dodecyl pyridinium bromide micelles, Mendeleev
Communications 245-248.
3. Mirgorodskaya, A.B., Kudryavtseva, L.A., luev, Yu.F., Idiyatullin, B.Z. and Fedotov, V.D. (2000)
Cetylpyridinium bromide-based microemulsions as media for the hydrolysis of phosphorus acids esters in
the presence of pro mary amines, Mendeleev Communications 205-206 .
4. SMerman, O. and Stilbs, P. (1994) NMR studies of complex surfactant systems, Progress in Nuclear
Magnetic Resonance, 26, 445-482.
5. Fedotov, V.D. , Zuev, Yu.F. , Archipov, v.P. and Idiyatullin, l.Sh. (1996) Self-diffusion in
microemulsion and micellar size, Applied Magnetic Resonance 11, 7-17.
6. Fedotov, V.D., Zuev, Yu.F., Archipov, V.P" Idiyatullin, Z.Sh. and Garti, N. (1997) A Fourier transform
pulsed-gradient spin echo nuclear magnetic resonance self-diffusion study of microemulsions and the
droplet size determination, Colloids and Surfaces. A: Physicochemical and Engineering Aspects 128, 39-
46.
INDEX

acetate ion 133 bead packs 553


acetic acid/sodium acetate 133 benzene 71, 75, 285
acetonitrile 486 benzenelHZSM-5262
acetonitrile diffusion 488 . benzoin 176
acid AISBA-15 mesoporous solids 307 benzophenon 176
acrylamide 586 BET method 607
adenosine 616 beta-diketonate ligand 121
adsorbed organic substances dynamics 485 bicontinuous microemulsions 130
adsorbed state 219 bidentate molecule 146
adsorbed xenon 255 bimetallic particles 455
adsorption 84, 197, 285 binding of platinum with DNA 616
aggregation 384 biological membranes 173
aggregration number 277 biomaterial mineralisation 209
AIBNIMMA 145 bioprostheses 215
AiMCM-41 448-450 Bioran 185
AlP04 aluminophosphate 108 bioresorable ceramics 209
AIP04 -11 aluminophosphate 108 birdcage resonator 116
AlP0 4 -CHA III bis(triethoxysilylpropyl)tetrasulfane 146
AIP04 -CHA aluminophosphate 110 bone structures 209
21Al MAS NMR 494 bone substitutes 209
21Al NMR 546, 559 brain tissue diffusion 643
AI203-Ga203-Zr02 501 breast implants 210
A120 3-Zr02 501 breast protheses 209
AIF3562 Bronsted acid sites 625
alfa-methyldeoxybenzoin 176 Bronsted acidity 565
alkali metals, reduction by 609 B-ZSM-5220
alumina 224, 559 I-butene 85,88,91,93,94
alumino phosphates (AlPO) 545 l-butenelNaX system 84
aluminum 308
AIV04 359 l3C 75
ammonium group 140 13C 13C spectra 75
anatase 297 l3C 2D exchange NMR spectrum II
Anderson-Wiess model 574 l3C chemical shift analysis 140
anion motion 29 l3C chemical shifts 84
anisotropic diffusion in nervous tissue 643 l3C CP MAS 149
anisotropic rotation 90 l3C CP MAS NMR spectrum 144
anthraquinone adsorption 501 l3C MAS 74, 80,409
antitumor activity 616 l3C MAS NMR 74, 77, 80, 137,409,546,559
APGSTE sequence 604 l3C MAS spectra 74
Arrhenius curve 77 l3C NMR 74, 89, 219, 513, 603
asphaltenes 591 l3C single pulse experiment (SPE) MAS NMR
auro-correlation function 185 138
l3C single pulse experiment MAS 145
liB NMR 220 l3C SNP 176
l31Ba 43 l3C solid state NMR 140, 604
Ba2 V20 1 359 l3C SPE MAS 145, 149, 151

655
656

13C spin-lattice relaxation times (T I) of the C60 conductivity spectroscopy 29


515 confined liquid crystals 190
13C{ 19F} 409 connectivity pores 163
13C{IH} 409 controlled pore glasses (CPO) 469
I3C_ IH dipolar interaction 80 convection 328, 338
40Ca 43 copper oxide 482
capillary transport of liquid 200 copper sulfate trihydrate 519
carboxylic acid 586 comer filling 162
cardiac valves 215 correlation function 186
Carr-Purcell-Meiboom-Oill 593 correlation times 150
Carr-Purcell-Meiboom-Oill (CPMO) sequence cosmetic emulsions 375
403 counterflow 425
catalyst tortuosity 605 C.PIMAS 44, 145, 149, 375,409
catalysts of Lewis acid type 625 CPMAS NMR 145
catalytic reactions 219 CPMO sequence 6, 322
CCI 2F2 559 CPMO train 7
Ce(S04)2 480 CP-MQMAS 112
cellulose fiber 631 cross polarization (CP) 107, 109, 141, 308
Ce02365 cross polarization dynamics 137, 148
Ce04 366 cross-linked gels 134
ceramic phosphates 210 cross-linking agents 136
Ceria 479 cross-peaks 90, 92
cetyltrimethylammonium 138 cross-relaxation 90, 185, 633
cetyltrimethylammonium chloride 144 cross-relaxation rates 90, 92, 93
channel architecture 57 crown/saddle system 75
CHCI 3 562 crude oils 591
CHCIF2 559 cryoporometry 157
chemical pulping process 634 crystal-field parameters 370
chemical shielding anisotropy 538 Cs+ cations 268
chemical shielding Hamiltonian (HCS) 16 CS 2V4011 360, 544
chemical shift anisotropy (CSA) 44, 45, 597 Cu(II) 445
chemical shift tensor 115 Cu/ZrO(OH)2 579
chemically induced dynamic nuclear polariza- Cu2+ ions 519
tion (CIDNP) 173, 174 Cu-Ce oxide 479
chemotherapy 350 CuS04 481
CHF3 562 cyclohexane 159
chlorofluorocarbons 224, 559 cyclononene ring 75
cholesterol 507 cytidine 616
35CI43 cytostatic activity 615
clay mineralogy 43
C032- 409 10 27 AI NMR 219
coke formation 603 20 IH NOESY NMR 95
coke, chemical composition 607 20 IN NOESY MAS NMR 94
colchicinoids 620 20 3QMAS 27AI NMR 223
colloidal aggregation 591 20 cross section 8
colloidal science 123 20 OAS-CPMAS 109
colloids 383 20 exchange 77
columns, paraelectric 72 20 EXSY 115
columns 72 20 HETCOR MQMAS NMR 109
columns, anti ferroelectric 72 20 image 198
columns, ferroelectric 72 20 MQ MAS 27AI NMR 219, 223
computer simulations 185 20 NMR (20 WISE) 546
657

2D NOESY 89 EHEC solution 137


2D NOESY NMR 90 Einstein law 594
2D spin-echo sequence 198 Einstein relation 605
2D VEXSY 430 elastomers 525
DAS 108 electron microscopy 634
DCP-MAS 409 electron paramagnetic resonance (EPR) see
DD"MQMAS III EPR
deactivation of catalysts 603 electron spin relaxation 173
density matrix 175, 572 electron spin resonance 445
deoxybenzoin 176 electronic spectroscopy 609
deuterium NMR 73, 74 electron-spin density 625
dibenzyl ketone 176 emulsions 124,317,325,337
dichlorodiamineplatinum 615 ENDOR 268
didodecyltrimethylammonium 128 EPR 268, 277, 365,479, 519, 577,609
differential scanning calorimetry (DSC) 74, 76, EPR adsorbed 2,2,6,6-tetramethylpiperidine-N-
145 oxyl 500
diffusion 124, 285, 383 EPR adsorbed anthraquinone 501
diffusion coefficients 324, 587 EPR measurements 299
diffusion diffractograms 132 EPR spectra 173
diffusion equation 385 EPR study 577
diffusion time 135 ESR 531
diffusitivity 191 eta(e)-caprolactame (CPL) 148, 149
diffusivities (D) of confined samples 486 eta-amrninocapronic acid 148, 149
diffusivity of n-heptane 606 ethyl(hydroxyethyl)cellulose (EHEC) 134-137
diffusometry 123 Euclidean geometry 67
dinuc1eotides 616 EXAFS 33
dipolar correlation effect 188 exchange NMR 10
dipolar coupling 107, 571
dipolar Hamiltonian (HD) 16 17F MAS 409
dipole-dipole interaction 175 19F MAS NMR 215
direct micelles 637 19F NMR 219, 559
di sc-like cores 71 F-ions 409
discotic liquid crystals 71 falling drop 327
DNA 616 falling liquid films 424
double PGSE pulse sequence 555 fast dynamic processes 115
double resonance 308 fast exchange 92
double rotation (DOR) 107 Fe(III) 445
drop 327 Fe/ZrO(OH}z 579
droplet coalescence 375 ferrite particles 383
droplet-size distribution 337, 343 field gradient 328
drying 318 field gradient NMR 3
drying process 393 Floquet method 77
DWSE imaging sequence 644 flow dispersion in porous media 439
dynamic angle spinning (DAS) 107 flow imaging 197
dynamic model of polymer chain 573 flow velocity maps 197
dynamic of spin system 573 flow-compensating PFG pulse sequences 338
dynamics of adsoption 84 fluorinated alumina 629
dynamics of diffusion 603 fluorinated surfactant 165
dynamics of the template 548 foodstuffs 124
Fourier coefficients 4
Echo-decays 135 Fourier transformation 3
EHEC gel 137 fractal 591
658

Fredholm equation 339 hexadecane 128, 142, 143, 145


frying of alumina pellets 200 hexadecyltrimetbylammonium chloride 546
fullerene solubility in organic solvents 513 hexaheptyloxy TBCN 73
fullerenes 513 hexasubstituted TBCN 72
HFI-tensors 175
g tensor 173 high-resolution IH 74, 75
GaP3-Zr02501 Hodgkin's disease 350
GARField (Gradient at Right-angles to the Hofmann elimination 139
Field) 317 Hofmann reaction 145
gas hydrates 116 HP Xe 121
Gaussian diffusion 126 HR (high resolution) MAS NMR 83
Gaussian line shapes 83 HR MAS spectroscopy 83
gels 123, 124 HSQC 89
general cross-validation (GCV) 340 HSQC experiments 85
Gibbs-Thomson equation 155 hydrocarbon conversion 603
gradient coils 4 hydrocarbons 285
gradient echoes 3 hydrochlorofluorocarbons 559
gradient fields 7 hydrodynamic dispersion 197
gradient pulses 133 hydrodynamic radius 134
guanosine 616 hydrodynamic size 389
hydrogen water 403
IH 75, 319, 375, 409 hydroxyapatite 210
IH MAS NMR 212, 546 hydroxylamines 625
IH NMR 84, 85, 485, 603 hydroxysodalite 491
IH NMR relaxation 593 hyperpolarized xenon 116, 121
IH NMR relaxometry 591 HZSM-5 zeolite 285
IH NMR transverse relaxation rate 403
IH NOESY NMR 84, 89 imaging 162, 321
IH SNP 176 implants 209
IHf 7AI} TRAPDOR 309 inelastic neutron scattering 29
IH_13C CPIMAS dynamics 546 inorganic vanadates 597
IH_13C HSQC 84 interatomic connectivities 107
IH_ 29Si cross polarization (CP) 137 interfacial tension 375
I-hexadecene 140 internuclear connecti vities 107
IH-MAS NMR 89, 94 intracrystallite diffusion coefficient 285
IH-NOESY 84, 91 ion-exchanged 445
2H NMR organic nanopartic1es 507 isomerization of n-hexane 566
2H NMR of CD3COCD 3 511
2H NMR of CDCI3 510 joint two-time probability density 333
2H NMR of deuterated water 508
Hahn echo 3, 126 39K 43

Hahn-echo decays 526 4-K electron paramagnetic resonance (EPR)


Hahn-spin echo 34 479
heart valves 209 K3VSO I4 544
HETCOR MAS 409 kaolinite 491
HETCOR NMR 108, 109 Keggin PW 1204Q3- ion 567
heterogeneous catalytic reactions 197 Knight isotropic shift (K) 455
heteronuc1ear correlation (HETCOR or WISE) Knight shifts 473
107-109, 308 kraft pulp 632
heteronuc1ear recoupling via REDOR 107 Kramers doublet 370
heteropolyacids 565
hexabenzocoronene 71
659

139La NMR 231 mechanical dispersion 554


labeled mesogens 713 mechanisms of relaxation 185
Langevin function 385 melting transition 76
Langevin law 384 mesophases 71, 72, 74, 75, 77, 79, 80,
Lanthanum vanadate (LaV04 ) 358, 597 mesoporous 445
Larmor frequency 110 mesoporous materials 43, 57, 485
laser light 97 mesoporous silica 139
laser-polarized xenon 97 mesoporous silica MCM-41 145
L-curve method 340 metacyclophanes 71
Legendre polynomial 3 metakaolinite 491
levyne structure 222 metal 470
Lewis acid sites 625 metal ion 445
Lewis acidity 497 metavanadates 540
Li+ cations 268 methanol 226
7Li resonance 35 methylcyclohexane 116
(LiNb03)I-x-(W03)x 15 MFI-type zeolites 57,60
LiNb02-W03 17 micellar catalysis 647
Line width (Delta V) 455 micelles 127, 173,277,592
Liouville equation 175 micellized radicals, diffusitivity of 178
lipid chain dynamics 638 microemulsions 126,507,637,649
liquid crystals 185 micro-imaging 118
liquid interactions at solid surfaces 166 microporous solids 255
liquid/gas interface 423 mixed alkali effect (MAE) 15,31
local field gradient 571 MMA 145
local resonance offsets 188 25Mg 43
longitudinal diffusion 643 95Mo 43
longitudinal relaxation 188 55Mn NMR 231, 236
longitudinal relaxation rate 385 55Mn NMR in zero external magnetic field
longitudinal relaxation time 167 235
long-time diffusion 134 modulated gradient spin echo method (MGSE)
Lorentzian line shapes 83 433
low-field l3C MAS 79 molecular motion 571
lubricants 71 molecular motion in porous media 434
lubrication approximation 317 molecular orientation 80
molecular reorientation 73
macrocyclic molecules 97 molecular traffic control 60
magic angle spinning (MAS) techniques 83 molecular-scale technique 116
magnetic field gradients of a porous system molecule diffusion 186
162 monometallic particles 455
MAS 112 Monte Carlo simulations 60
MAS NMR 307 motion, ultras low 185
mass transport 197, 201 MQCP-MQMAS 108
maxillofacial surgery 209 MQMAS 110-112
MCM-41 65, 137-152 MQ-REDOR 107, 112
MCM-41 silica 137 MQ-tl-REDOR 110, III
MCM-41IMMA 138, 145 MQ-trREDOR 110
MCM-41INylon-6 composite 149, 152 MR microimaging 643
MCM-41IPEO 151 MR197,285
MCM-41IPMMA 140 multidimensional exchange NMR 7
MCM-41IPMMA nanocomposites 145, 148, multidimensional NMR 3, 8
152 multinuclear NMR techniques 604
Me-AIMCM-41 445,451 multiple echo sequence 319
660
multiple quantum magic angle spinning non-Oaussian diffusion 126
(MQMAS) 107,355 NTBCN 73, 79
nuclear magnetic relaxation 525
Na+ cations 268 nuclear magnetic resonance (NMR) techniques
NaA zeolite 496 16, 17
14N MAS NMR 44, 48 nuclear Overhauser effects 188
14N MAS NMR spectroscopy 43
14N MAS NMR spectrum 45, 54 11043
14N quadrupole coupling 45 110 NMR 29
(Na3P04) l-x-(Na2S04)x 15, 21 offset fields 3, 7, 10
23Na NMR 470 oil/gas formation in the North Sea 43
23N a NMR spin-lattice relaxation 24 oil-in-water emulsions 342, 375
nafion 267, 268 oil-in-water microemulsions 651
n-alkanes 144 optical anisotropy 72
nanochannels 144 optical microscopy 74, 76
nanocomposites 150, 151 optical pumping 97, 115
nanophases 137 orientational anisotropy Ill, 190
nanopores 140 orthocyclophanes 71, 79
nanoporous materials 57 orthopaedic surgery 209
nanosubstituted TBCN (NTBCN) 73, 79 orthovanadates 599
nanosystems 97 osteoformation 209
naphthalene 71 othovanadates 538
NaX 89 outer sphere diffusion theory 384
NaX zeolite 84 oxoammonium cations 625
Neel relaxation time 386
network modifier species 32 31p 347, 375
NH4N0 3 51 31p CPIMAS 212
NH4V03 356 31p MAS NMR 211, 212, 565
n-hexane 285 31p NMR 352, 546
Ni(II) 445 31p NMR lineshape measurements 23
N-isopropylacrylamide 586 31p SNP 176
nitrogen adsorption-desorption isotherms 309 31p{l9F} 409
nitroxyl radicals 625 I-palmitoy 1-2-steroyl-(7 -doxy 1)-glycero-3-
NMR 4, 57, 307 phosphocholine 637
NMR cryoporometry 155, 163 paddle wheel mechanism 21
NMR diffusion measurement 162 paper structure 631
NMR diffusometry 124, 126, 127, 130, 134, paramagnetic probe 625
137,631 pathological calcification 209
NMR imaging 5, 197, 327, 393, 423 pathological mineralisation 214
NMR microscopy 328 Peclet number (Pe) 318, 554
NMR pulsed field gradient velocity encoding pentyl-4-cyanobiphenyl 185
techniques 423 percolation process 591
NMR relaxometry 123,632 perhydrotriphenylene (PHTP) 148
NMR self-diffusion 647 pericardium valves 211
NMR shift 161 petrochemical industry 603
NMR studies 71 petroleum colloids 591
NMR surface scanner 525 PFO displacement spectroscopy 204
NMR-MOUSE 6, 525 PFO NMR 245, 268
NMR-relaxation 268 PFO NMR instrumentation 58
NOESY NMR experiment 188 PFO NMR sample 67
non-alkanoyloxy TBCN (NTBCN-Cn) 74 PFO NMR self-diffusion 61
non-cubic solids 115 PFO NMR studies 65
661

PFG-NMR diffusion coefficients 585 propagator 205


phase transition 585 proteins 97
phosphatidylcholin (PC) 347 proton relaxation rates 185
phosphatidylinositol (PI) 347 protons 308, 395
phospholids 347 pseudorotation 75, 79
phospholipid interface 638 PtfAI20 3 604
photoionization on N-alkylphenothiazines 445 (Pt(NH 3MN)Clt 615
photolysis 176 (Pt(NH 3MN)H20f+ 615
platinum complex I -cisplatin 616 (Pt(NH3h(N 3-C)2)2+ 617
platinum complexes 615 (Pt(NH 3MN 3-C)Clt 617
p-nitrophenylolate-anion absorption 650 (Pt(NH 3MN 7-A)2)2+ 617
poisoning agent for catalysts 480 (Pt(NH 3h(N 7-A)Clt 617
polarization 97, 116, 118 (Pt(NH 3MN 7-G)2)2+ 617
polarization transfer 108 (Pt(NH3MN 7-G)Clt 617
poly(ethylene glycol) 637 (Pt(NH3)2CIH20t 615
poly(ethylene oxide) 137 (Pt(NH 3hOHCL) 615
poly(methyl methacrilate (PMMA) nanocom- (PtlRe-AI 20 3) 604
posite 138 195pt NMR 615
polyamine copper complexes 519 pulse gradient spin echo (pGSE) 433
polyanions 609 pulsed field gradient (PFG) 126,245,327,337,
polyarylenesulfophthalides 609 586
polycondensation 148 pulsed field gradient method (PFG NMR) 3, 57
polydiphenylenephthalide 609 pulsed field gradient spin echo (PSGE)
polyethylene 144 methods 328
polyethyleneoxide (PEO) 134, 151 pulsed field gradient stimulated echo 632
polyisoprene/silica compound 146 pulsed gradient fields 7
polymer nanocomposites 139 pulsed gradient spin echo (PGSE) 376
polymer solution 136 pulsed gradient spin echo (PGSE) NMR 553,
polymers 123, 124, 139 638
polymers coils 587 pulsed gradient stimulated echo (PGSTE)
polymers, crosslinked 571 sequence 604
poly-N-isopropylacrylamide (PNIPAM) 585 pyrovanadates 539
polynucleotides 616
polyradicals 609 lQCP-3QMAS 108
polystyrene beads 556 lQCP-MQMAS 108
porcine valves 211 3QCP-3QMAS 108
pore filling factor 84, 87, 89, 94 3QMAS NMR 598
pore filling regimes 162 3Q-~-MQMAS 110
pore size 603 5QCP-5QMAS 108
pore size distribution 155, 200, 486 7QCP-7QMAS 108
porosity 155 quadrupling coupling (QC) tensors 537
porous catalysts 197 quadrupolar contribution 107
porous glasses 185 quadrupolar Hamiltonian (HQ) 16
porous materials 115, 119, 155, 185 quadrupole interaction 140
porous media 553
porous silica 139 radical pairs 173
porous structure 433, 631 radical reactions 173
porous system 130 Raman spectroscopy 577
position exchange spectroscopy (POXSY) 8, random walk procedure 186
327,333 Rb2V60 16 361
POXSY experiments 9 Rb2V60 16 544
profiling 317 Rb 3V5014 544
662
ReIAl20 3 604 small-angle neutron scattering (SANS) 277,
real refractory mortar 393 591
REDOR 107, 110, 112 SNP 173
REDOR experiment 110 sodium aluminate 138
redox 369 sodium filaments 469
reduced 4D exchange NMR 10 solid catalysts 219
reform catalyst (PtlRe-AI 20 3) 603 solid polymers 124
reforming naphtha catalyst 604 solid state NMR 10, 16, 139, 565
reforming of naphtha 603 solid state NMR techniques 15
regeneration of catalysts 603, 604 solid-solid phase transitions 43
relaxation 158 solid-state 14N MAS NMR 43
relaxation mechanism 175 solid-state NMR 545
relaxation time analysis of a liquid 156 solid-state spectroscopy 4
relaxation times 486 sol vent/surfactant/solid system 166
relaxivity 383, 388 sorbents 197
relaxometry 190 Sorbsil 485
reorientation processes 79 spatial distributions of flow velocity 202
reverse micelles 637 SPE 149
rhodiarome 507 speciation 311
rocks 124 spectral densi ty Jij (w) 90
rotating-frame Hamiltonian 571 sphingomyelin (SM) 347
rotational diffusion 77 spin density 198
rotational echo double resonance (REDOR) spin density profile 327
107, 110, 112 spin echoes 126, 434, 437
ruthenium sulfide 531 spin echo double resonance (SEDOR) 34
ruthenium-vanadium sukfides 531 spin echo pulse sequence 458
spin exchange 115, 188
Safaniya asphaltenes 591 spin label 637
sand grains, beds of 65 spin polarization 97
Satellite transition spectroscopy methods spin probes 625
(SA TRAS) 356, 537 spin-lattice relaxation 118, 470
SDC 136 spin-lattice relaxation time 198
SEDOR experiments 34 spin-lattice relaxation time Tl 458
self-diffusion 65,124,159,253,337,375,434, spin-rotational coupling 175
632 spin-spin relaxation 593
self-diffusion IH 486 standard saturation-recovery technique 458
self-diffusion coefficients 124, 271, 332, 554, standard surfactant 163
604,639 STARS 44, 45
self-diffusion coeffricients of the micro- STARSNNMR 44
emulsion components 649, 650 steel-making converter 393
self-diffusion of macromolecules 248 Stejskal-Tanner equation 339, 633
sera 347 Stejskal-Tanner plot 126, 135
SERPENT experiment 8 stimulated echo 188, 341
shim coils 4 stimulated spin echo 571
29Si solid state NMR 145 stochastic molecular motions 185
29Si MAS-NMR 137,492 Stokes-Einstein equation 639, 651
29Si NMR 219, 220, 559 Stray Field Imaging 393
(Si(OH)2t(H 2PW 120 40t 567 structure of transport channels 268
silicas 157 sulfated alumina 629
SimFonia Bruker program 579 superficial velocity 344
single-file diffusion 60 superparamagnetic (SPM) nanoparticles 383
skin formation 323 surface coating 162
663

surface interactions 163 translational dynamics of macromolecules 250


surface relaxi vity 156 transmission-line tuning (TLT) CPIMAS 43
surfactant aggregates 123, 130 transversal diffusion 643
surfactant diffusion 124, 651 transverse magnetization 4
surfactants 649 transverse relaxation 188
susceptibility 159 transverse relaxation time 168
susceptibility effects 140 tribenzocyclonatriene (TBCN) 7 I , 72
switched external magnetic field (SEMF) 173, tricalcium phosphate 210
174 trimethylamine 139, 145
synchrotron radiation 110 triphenylene 7 I
synergy phenomenon 531 triple-quantum MAS NMR 597
tris(o-phenylenedioxy )cyclotriphosphzene 15 I
TI measurements 167 tris(triphenylsilyl) vanadate 597
TI relaxation time 167, 604 tris-o-pheny lenedioxycyclophosphazene (TPP)
T lpC 546 148
T lpH 546 tubulin 615
T 2 592 tubulin in platinum 620
T2 measurements 168 tungstophosphoric acid supported on silica 565
T 2 profile 322 two-dimensional flow maps 203
T2 relaxation time 168, 604 two-site exchange model 191
99Tc NMR 455 type B carbonate fluoroapatite 409
Taylor dispersion 553
technetium nanopartic1es 455 ultraviolet-visible diffuse reflectance spec-
technetium-ruthenium oxide-supported cata- troscopy 445
lysts 455
templating agent 138 V20 3 365
TEOS 307 V20 S 365
tetraethylorthosilicate 138 V20S.MxOy 356
tetramethyl-4-phenyl-3-imidazoline-l-oxyl 626 SIV chemical shielding tensors 537
tetramethylammonium hydroxide 546 SIV MAS NMR 597
tetramethyl-piperidine- l -oxyl 626 SIV NMR 531, 537
tetramethyl-piperidine-4-oxo-l-oxyl 626 SIV quadrupole coupling 597
tetramethylpiperidine-N-oxyl 497 vanadia based catalysts 355
tetrapropylammonium bromide (TPAB) 46 vanadium cerium oxides 365
tetrathiovanadates 534 vanadium oxide 534, 537
thermal analysis 577 vanadium paramagnetic centres 372
thermoreversible polymers at interfaces 585 vanadium pentoxide V20S 544
thermoreversible polymers in solution 585 vanadium sulfide 532
thin films 319 vanadium-based heterogeneous catalysts 597
three-pulse sequences 573 vapor deposition 469
three-pulse Z-filter sequence 355 Varian VNMR 44
Tikonov regularisation 339 velocity correlation 437
time-invariant fields 5 velocity correlation function 435
time-resolved fluorescence quenching (TRFQ) velocity distribution 332, 337
277 velocity exchange spectroscopy (VEXSY) 8,
time-resolved SNP kinetics 175 553
Ti0 2 297-299 velocity profile 327
TMA 139 velocity self-correlation function (VCP) 433
toluene-in-water emulsion 344 VEXSY experiments 9
tortuosity 155, 603 viscosimetry 591
TPR 366 viscosities 593
translational diffusion 188 Vleck theory 33
664

v04 polyhedra 601 129Xe NMR spectroscopy 115


VOCl3 598 x-ray diffraction 74
VOP04544 x-ray scattering 379
Vo/zr0 2 system 362 xylene-in-water emulsion 343
Vycor 185

Waltz16 sequence 377


Wang' s method 109 Zeeman Hamiltonian (Hz) 16
water 159, 268 zeolite HZSM-5 563
waterbone caotings 319 zeolites 43, 84, 85, 121,219, 224, 559
water-in-oil 637 zero field splitting 372
zero-loading 115, 118
Xe adsorbed in NaA 257 zeta-potential 403
Xe NMR spectroscopy 121 zirconia catalysts 497
129Xe chemical shift 607 67Zn 43
129Xe NMR 255, 603, 605, 606 Zr02 catalyst 577

Potrebbero piacerti anche