Вы находитесь на странице: 1из 9

SOLID STATE

Nuclear Magnetic
Resonance

ELSEVIER

Solid State Nuclear Magnetic Resonance 3 (1994) 49-57

Structural characterisation

of Na,ZrO,

T.J. Bastow a, M.E. Hobday b, M.E. Smith a,c**,H.J. Whitfield d


a CSIRO DiLlisionof Materials Science and Technology, Private Bag 33, Rosebank MDC, Clayton, Victoria 3169, Australia
Department of Applied Chemistry, Royal Melbourne Institute of Technology, Box 2476W, Melbourne, Victoria 3001. Australia
Physics Laboratory, University of Kent, Canterbury, Kent CT2 7NR, UK
d Department of Applied Physics, Royal Melbourne Institute of Technology, Box 2476W, Melbourne. I/ictoria 3001. Australia
(Received 24 June 1993; accepted 7 October 1993)

Abstract
A combination of X-ray and electron diffraction, electron microscopy and solid-state nuclear magnetic resonance
(NMR) has been used to elucidate the structure and the ordering of Na,ZrO,. The diffraction data confirm a
monoclinic crystal structure. A sample prepared by a conventional solid-state reaction of the components is shown
by both X-ray diffraction and electron microscope imaging to have an extremely high concentration of planar defects
associated with stacking disorder of the planes along the c-axis. The incidence of these defects is significantly
reduced in a sample recrystallised from a bismuth oxide flux. NMR indicates that the local coordinations are well
defined in both samples but with some sharpening of the spectra from the recrystallised sample indicative of the
increase of long-range order. The 23Na magic angle spinning (MAS) NMR spectra clearly show three distinct sites
with widely differing quadrupolar interaction parameters that can be related to the known site symmetries. Two
distinct oxygen resonances are observed in the MAS NMR spectrum from an O-enriched sample while the static
9Zr NMR spectrum can be simulated with one set of interaction parameters.
Key words: Magic-angle spinning nuclear magnetic resonance; Structure; Na,ZrO,;

1. Introduction
Structural
characterisation
of solid materials
has developed rapidly over the last decade as new
methods have become available to augment the
established
techniques
of diffraction
and microscopy. A methodology
that employs a suite of
complementary
techniques
that are sensitive to

* Corresponding author. Address for correspondence: Physics


Laboratory, University of Kent, Canterbury, Kent CT2 7NR,
UK.

Diffraction; Electron microscopy

differing aspects of the structure is clearly superior to any single technique.


In this paper a
combination
of diffraction
(both X-ray and electron), to probe the long-range
structure, electron
microscope
imaging, to view spatially localised
information,
and solid-state
NMR, to elucidate
the local (nearest
and next nearest neighbour)
structure,
is used to examine sodium zirconate
(Na,ZrOJ.
Na,ZrO,
is isotypic with Li,SnO,
[ll and
Li,TiO,
[2], crystallising
in a monoclinic
space
group C2/c with unit cell parameters
a = 0.5623
nm, b = 0.9749 nm, c = 1.1127 nm and p = 99.98,

0926-2040/94/$07.00 0 1994 Elsevier Science B.V. All rights reserved


SSDZ 0926-2040(93)E0045-Y

50

T.J. Basfow et al. /Solid State Nucl. Magn. Reson. 3 (1994) 49-57

with eight formula units per unit cell [3]. This is a


distorted sodium chloride structure; compare the
axial ratios of 1: 1.734: 1.979 and p = 99.98 to
the undistorted values of 1: 1.732 : 1.915 and p =
100 (on the basis of a monoclinic unit cell).
Layers form that are stacked up along the c-axis
with alternating cation layers containing either
only sodium or sodium and zirconium. Such ordered rock-salt structures seem to be susceptible
to polymorphism: monoclinic [3], orthorhombic
[4] and hexagonal [5] forms of Na,ZrO,
have
been reported. The monoclinic forms of Li,TiO,
and Na,ZrO,
are isostructural so that as the
structure of Li,TiO, has been refined [6] (atomic
coordinates are given in Table 1) the fractional
positions of the corresponding atoms are assumed to be approximately the same in Na,ZrO,.
In Li,TiO, order-disorder
effects have been observed and polymorphism in the closely related
Li,SnO, could be reinterpreted on the basis of
stacking disorder [7]. In this present work it is
shown that order-disorder
arising from stacking
faults also occurs in Na,ZrO,.
The local coordination of the different elements can be probed by NMR. All the nuclei in
Na,ZrO, possess isotopes with a magnetic moment so that NMR experiments on each are
possible. All are quadrupolar nuclei but their
differing resonance frequencies and quadrupole
moments demand various NMR approaches.
Sodium is a sensitive NMR nucleus with modest
quadrupolar interactions so that fast MAS [8] is
directly possible. The isotropic chemical shift
range of sodium tends to be small so that substantial overlap of resonances occurs if the second-order quadrupolar broadening is significant.
In fact the presence of the quadrupolar interaction can be advantageous as it also provides information about the local surroundings of the nucleus. For nuclei with small isotropic chemical
shift ranges the quadrupolar interaction is often
more sensitive as a structural probe than the
chemical shift. 91Zr has a nuclear spin I=5/2
and a natural
abundance
of 11.2%. The
quadrupolar coupling constants C, (C, = eqQ/h,
where eq = maximum component of the electric
field gradient (EFG) and eQ = nuclear electric
quadrupole moment) for 91Zr tend to be in excess

of 10 MHz so that only the (l/2, -l/2)


transition is observed which itself can extend over
0.4-0.5 MHz (at a magnetic field of 9.4 T) [9].
These widths are well beyond narrowing by current MAS rates so that a static approach is
adopted. 0 is potentially the most attractive
nucleus in this sample for NMR with a large
chemical shift range [lO,ll] and small nuclear
electric quadrupole moment. The major drawback for 0 is its poor NMR sensitivity due to its
low natural abundance (0.037%). However, even
modest enrichment (2-5%) is sufficient to make
the NMR study of 0 attractive, particularly in
samples where C, is small so that the resonances
tend to be narrow. Such enrichment has been
shown to be straightforward and generally adequate for the MAS NMR study [12,13]. In this
paper the complete multinuclear magnetic resonance information is presented in conjunction
with diffraction and microscopy data.

2. Experimental
Zirconium dioxide and anhydrous sodium carbonate in stoichiometric proportions were intimately mixed and pressed into a pellet. The pellet was heated in a stream of nitrogen to 950C
for 16 h. 0 isotopically enriched Na,ZrO, was
prepared by starting with ZrO, that had been
enriched in 0 by reaction of 10 mol% O-enriched H,O with zirconium isopropoxide, as previously described [12]. Then further annealing for
24-h periods at various temperatures in the range
500-1050C
was performed.
A sample of
Na,ZrO,
powder was then mixed with 10%
(w/w) of Bi,O, powder, pressed into a pellet and
annealed at 950C in either a nitrogen atmosphere or under vacuum until all the Bi,O, had
evaporated from the specimen.
X-Ray powder diffraction
patterns
were
recorded on a Rigaku Miniflex X-ray spectrometer using CuKa radiation. Electron micrographs
and selected area diffraction patterns were obtained with a JEOL 2010 electron microscope,
from crushed powders suspended on holey carbon grids.
The NMR was carried out on a Bruker MSL

T.J. Bastow et al. /Solid State Nucl. Magn. Reson. 3 (1994) 49-57

400 spectrometer (static magnetic field of 9.4 T)


operating at 37.21 MHz c91Zr), 54.24 MHz (170)
and 105.8 MHz (23Na). For 91Zr, static pointwise-iacquired spectra were obtained by stepping
the spectrometer frequency and measuring the on
resonance intensity of the Fourier transformed
spin-echo (as described in ref. 9). A Hahn spinecho, r/2-T-rr-acq,
was applied with r/2 pulses
of 8 ps and extended-phase cycling [14]. Using a
23-mm-diameter transverse solenoidal coil 8000
scans were accumulated at each frequency, retuning the probe at each step. A recycle delay of 0.25
s was sufficient for full relaxation. The shift scale
was referenced to BaZrO, at 0 ppm. 170 MAS
NMR spectra were acquired using a 7-mm double-bearing (DB) probe using spin rates of around
5 kHz, 5-s recycle delays and 1.6~/LSpulses (N
7r/4). The spectra were referenced to water at 0
ppm. 23Na NMR spectra were collected using a
4-mm DB MAS probe with spin rates of 12 kHz,
short pulses (tip angles <r/8)
with a recycle
delay at 0.5 s, which was sufficient to prevent
saturation. These spectra were referenced to solid
NaCl at 7.2 ppm. To constrain the simulations of
the 23Na centre bands additional data were taken
at magnetic fields of 11.7 T on a Bruker MSL 500
spectrometer operating at 132.3 MHz and at 14.1
T on a Varian VXR-600 spectrometer operating
at 158.7 MHz, with all other experimental parameters being similar to those at 9.4 T. Spectral
simulations were performed using the Bruker
POWDER program.

(a)

51

0
Na
Zr

(b)

000
OO
00

Q
O

000
00

so0
00s

00

Fig. 1. (a) Projection of the structure of Na,ZrO, along [loo]


using the coordinates reported for Li,TiO, (Table 1). The
Zr-0 bonds for the three distinct oxygens in the asymmetric
unit are shown. Unshaded Zr atoms are at x = 0 and the
shaded Zr atoms are at x = l/2. (b) The ABC sequence of
cation layers produced by a stacking fault.

3. Results and discussion


3.1. X-Ray powder diffraction
The X-ray diffraction patterns from the sample formed by direct solid-state reaction gave a
mixture of sharp and diffuse lines. For example
the 002 reflection was sharp and strong whereas
the 020 line was diffuse and asymmetric. Such a
mixture of sharp and diffuse lines has previously
been observed in cobalt powder and was explained in terms of stacking faults of the closepacked layers of the metal [15,16]. Treaty et al.
[17] have recently published a general recursion

Table 1
Atomic coordinates for monoclinic (group C2/c) Li,TiO, [2]
which is isostructural to Na,ZrO,
Atom

Equipoint

Li(l)
Li(2)
Li(3)
Ti(1)
Ti(2)
OfI)
O(2)
O(3)

8f
4d
4e
4e
4e
8f
8f
8f

Coordinates
x

0.238
0.250
0.000
0.000
0.000
0.141
0.102
0.138

0.077
0.250
0.045
0.415
0.747
0.265
0.586
0.906

0.000
0.500
0.250
0.250
0.250
0.138
0.138
0.135

52

T.J. Bastow et al. /Solid State Nucl. Magn. Reson. 3 (1994) 49-57

algorithm for calculating the intensities of powder diffraction patterns in the presence of planar
faulting and shown that the diffraction peak
widths become index-dependent.
The proposed
stacking pattern is shown in Fig. 1. Fig. la shows
the structure of unfaulted Na,ZrO,
using the
fractional coordinates found from the refinement
of the Li,TiO, structure and listed in Table 1.
For this structure the sequence of sodium and
mixed sodium-zirconium
layers up the c-axis is
. . . ABAB . . . Fig. lb shows the cation layers only
of a faulted stacking sequence ABC. Such stacking faults account for the marked variation in the
relative peak heights of the 002 and 020 X-ray

Fig. 2. Electron micrograph


maximise the contrast.

of a crystal of Na,ZrO,

showing

stacking

diffraction reflections observed by other workers


for cobalt [15,161 and for Na,ZrO, in the present
study and commented on by Binner et al. [18].
Such stacking faults are difficult to anneal out
by simple heat treatment and, indeed, on heating
to temperatures between 500 and 1050C no observable change in the X-ray diffraction spectrum
occurred. However, it has been found that recrystallisation from a suitable flux (here Bi,O,) is
effective at removing these faults. The X-ray
diffraction pattern from the final flux-annealed
specimen indexes on a monoclinic cell of the
space group proposed by Lang [1,2] for isotypic
compounds and is in general agreement with the

faults. The crystal

has been tilted in an arbitrary

direction

to

T.J. Bastow et al. /Solid State Nucl. Magn. Reson. 3 (1994) 49-57

53

3.3. Solid-state NMR

pattern reported by Enriquez et al. [19]. Indexing


on other unit cells (e.g., orthorhombic or hexagonal) can be readily explained as changes in the
degree of stacking faulting, which is probably a
function of the powder synthesis conditions.

Diffraction and microscope imaging establish


the long-range order within the structure. NMR
complements these measurements by examining
the local structural configurations present. The
9Zr static NMR spectrum shows the (l/2, - l/2)
transition extending over approximately 350 kHz
(Fig. 3). This spectrum can be simulated by a
single set of NMR interaction parameters of C,
= 14.6 MHz, 77= 0.15, &, = - 0 ppm and a
smoothing of N 20 kHz. Although there are two
zirconium sites in the crystal structure they are
each in position 4(e) with identical nearest and
next nearest neighbours and similar bond lengths
so that similar NMR interactions at each site can
be anticipated, and any differences will not be

3.2. Electron diffraction and microscope imaging


A low-resolution electron microscope image of
Na,ZrO, prepared by direct solid-state reaction,
without subsequent recrystallisation from a Bi,O,
flux, shows directly the presence of a high degree
of stacking faults (Fig. 2). Electron diffraction
patterns normal to the c-axis show very sharp
spots confirming that the sodium and zirconium
ions are perfectly ordered within individual layers.

co

0
0
0

00
00

0
0

00

80

0
0
0

0
0

0
0
0

260

160

-lb0

Fig. 3. 91Zr static NMR spectrum of Na,ZrO,.

-2bo

54

T.J. Bastow et al. /Solid State Nucl. Magn. Reson. 3 (1994) 49-57

(b)

1000

750

500

250

-250

-500

170 Chemical Shift in ppm.

320

310

300

290

280

170 Chemical Shift in ppm.


Fig. 4. 0 MAS NMR spectrum of Na,ZrO, showing (a) an
expansion of the centre bands and (b) the side band manifold.

resolvable within the static spectrum. It should be


noted that although one set of parameters can
simulate the overall features of the spectrum the
observed spectrum is not as smooth as expected
for a single site, as observed for polymorphs of
zirconia with one site (e.g., tetragonal) [9], perhaps indicating minor differences between the
two zirconium sites in Na,ZrO,. Recognition of a
quadrupolar line shape does show that locally the
structure is ordered as any significant variation in
sites through the structure (e.g., as in locally
disordered compounds) gives rise to a distribution of powder patterns that blur the observed
line shape into one with no distinct features, as
observed for 91Zr in cubic ZrO, [9].
170 MAS NMR spectra reveal two peaks with
positions 308.9 k 0.1 and 286.0 + 0.1 ppm, and
approximately equal widths (FWHM) of 55 Hz in
the flux-annealed sample (Fig 4a). This half width
places an upper limit on C, ( < (20~,/3)(154/
2)12, vr is the Larmor frequency, A is the rms

(root mean square) half width of the centre band,


relative to vI, and 77 is assumed to be 0 [ll]) of
650 kHz. In fact C, is probably considerably
smaller than this value as the prominent side
bands from the (& 3/2, + l/2) transitions would
extend over a range of approximately k 97.5 kHz
[from V~= 3C,/21(21l)] for C, = 650 kHz and
even though the signal-to-noise ratio is not good
enough to accurately measure this value (Fig. 4b)
the extent of side bands is clearly considerably
smaller than k97.5 kHz. C, could be more accurately deduced from a slowly spinning sample to
accurately register the envelope of the outer transitions but this would require a much more highly
enriched sample. As C, is at most 650 kHz this
means that any second-order quadrupolar corrections of the peak position to give the isotropic
chemical shift will be less than 0.8 ppm [20]. The
small C, value is to be expected from the high
ionicity of sodium and zirconium bonds with oxygen [21]. The sample prepared by direct solid-state
reaction gave a line width (FWHM) of 200 Hz.
This factor of four decrease in the line width
between the directly prepared and flux-annealed
samples could be explained by a decrease in C,.
However, a more likely explanation is a decrease
in the chemical shift dispersion in the flux-annealed sample which is a local manifestation of a
decrease in the number of stacking faults, thereby
increasing the amount of order throughout the
structure. This is a much more subtle example of
the decrease in 170 line width reported recently
in the crystallisation of titania from amorphous
gels [131.
In the crystal structure (Table 1 and Fig. 1)
there are three oxygen sites but from I70 NMR
spectra only two distinct resonances can be resolved. Spectral integration gives 23.6 & 0.5 : 12.5
* 0.5, effectively 2: 1, indicating that locally two
of the sites must be very similar. All three oxygens are in eight-fold positions being locally coordinated in an octahedron of four sodiums and
two zirconiums, with the zirconiums on adjacent
corners. The average (Na,Zr)-0
distances are
0.235, 0.227 and 0.239 nm for O(l), O(2) and
O(3), respectively. The orientation of the local
coordination octahedra relative to the crystal axes
for each of the oxygen sites is also informative.

T.J. Bastow et al. /Solid State Nucl. Magn. Reson. 3 (1994) 49-57

From Fig. la for both O(1) and O(3) there are


zirconiums attached above and below (only Zr-0
bonds are drawn for clarity) the plane of the
oxygen, while O(2) has both zirconiums below the
plane. On the basis primarily of the similarity of
the orientation of O(1) and O(3) coordination
octahedra and the marked difference of O(2) we
assign the peak at 308.9 ppm to O(1) and O(3).
Additionally these two also have the more similar
average bond lengths.
The sodium MAS NMR spectrum showed a
highly structured centre band due to overlapping
second-order quadrupolar line shapes from the
three sites. However, there is sufficient resolvable
structure to allow simulation of the spectra which
is constrained by performing experiments at three
applied magnetic fields, 9.4 T (Fig. 5a), 11.7 T
(Fig. 5f) and 14.1 T (not shown). At both 9.4 T
and 11.7 T experiments were performed on two
independent samples, each being simulated several times to ascertain the errors quoted in Table
2. It is clear that although there are distinct shifts
for each site it is the quadrupolar coupling constant that exhibits the more marked variation.
The simulations at 9.4 T (Fig. 5b) and 11.7 T (Fig.
5g) are shown for the single set of parameters
that resulted in the best fit at all fields. The three
individual components for the 11.7-T simulation
are shown separately (Fig. 5c-e).
Careful quantification of the spectra has also
been performed. Essentially only the (l/2, - l/2)
transition contributes to the centre band, so that
it is the fraction of the (l/2, -l/2)
transition
that appears in the centre band that determines
the observed intensities which depend on the
parameter vi/vrv, (where V, is the MAS rate)
[22]. This fraction can be extracted from the
graphical plots of this function that appear in the
paper by Massiot et al. [22] and is listed for the
three sodium sites in Table 2. The integrated
intensities measured from the spectra can then be
suitably corrected by this factor and renormalised
to give the actual distribution of sodium between
the three sites in the structure. The three sodium
sites occur in the structure in the ratio 2 : 1: 1 (i.e.
50% : 25% : 25%) for Na(l), Na(2) and Na(31, respectively. Although 44% : 22% : 34% (Table 2)
does not accurately agree with the expected dis-

Cc)
50

25

-25

-50

-75

-100

23Na Chemical Shift in ppm.

50

25

-25

-50

-75

23Na Chemical Shift in ppm.


Fig. 5. Z3Na MAS NMR centre bands of Na,ZrO, showing
(a) the experimental data, (b) the full simulation at 9.4 T,
(c-e) the individual simulation components at 11.7 T, (0 the
experimental data and (g) the full simulation at 11.7 T. (Note
that the intensities of the individual components c-e are not
those used in the full simulation g.)

T.J. Bastow et al. /Solid State Nucl. Magn. Reson. 3 (1994) 49-57

56

tribution (and is much less convincing than some


examples for 27Al(e.g. refs. 22 and 2311,given the
errors quoted, it is shown (vide infra) that this
intensity distribution allows unequivocal assignment of the site with the highest intensity. The
major error in the distribution comes from the
site with the largest C, since this produces the
broadest resonance so that the integral is sensitive to the baseline correction (much more so
than the other two lines) and the effect is exaggerated as it has the lowest intensity in the centre
band. To test the uniqueness of this fit rather
than assigning structural intensities of 2 : 1: 1 to
the rows of Table 2 the other two possibilities
were simulated, i.e., 1: 1: 2 and 1: 2: 1, which
would give observable centre band intensities at
9.4 T of 31%:44%:25%
and 27%:63%:11%,
respectively. These intensity distributions produce much poorer simulations of the observed
spectra. On the basis of this it is with some
confidence that the site with C, = 2.52 MHz can
be assigned to Na(1) (the 8f site).
The three sodium sites in the structure all
have an octahedron of nearest neighbour oxygens. At the next nearest neighbour level Na(3)
(in 4e) has a hexagon of zirconiums in the same
layer and a triangle of sodium cations in the
layers above and below. Na(1) (in site Sf) and
Na(2) (in site 4d) have identical next nearest
neighbours with six sodiums in the plane and
triangle of one sodium and two zirconiums in the
layers above and below. There appears to be no
correlation of this structure to the isotropic chemical shift (Table 2). However, for C,, although it

is difficult to calculate absolute values of the


EFG directly as this tends to be dominated by
charges close to the nucleus, it should be somewhat related to the local site symmetry. One
crude measure of this might be expected to be
the difference in the maximum and minimum
Na-0 bond lengths which are 0.026, 0.017 and
0.056 nm for Na(l), Na(2) and Na(3), respectively. The site which has been assigned to Na(1)
on the basis of its intensity does have the intermediate C4 value and corresponds to the intermediate difference in Na-0 bond length. Then,
as Na(3) has the largest distortion, this can be
assigned to the site with C, = 4.20 MHz, leaving
Na(2) to be assigned to the site with C, = 2.08
MHz. The presence of second-order quadrupolar
structure here definitely aids structural assignment of the resonances. The observation of distinct quadrupolar line shapes is further evidence
that the structure is locally ordered with the
planes being essentially perfect at the nearest
neighbour level, emphasising that, even if the
stacking is faulted, changes in the next nearest
neighbours do not significantly change the EFG.
The effects of stacking faults on the sodium spectrum are more subtle than for the 0 spectrum
(where second-order
quadrupolar
effects are
small) causing merely a slight smoothing of the
singularities and discontinuities of the powder
pattern in the NMR spectra from samples prepared by direct solid-state reaction compared to
those that have been flux-annealed.
It has been shown that a combination of microscopy, diffraction and NMR is a powerful ap-

Table 2
23Na NMR interaction parameters that gave the best simulations for all magnetic fields and the measured intensities, fractions of
magnetisation of the (l/2, - l/2) transition in the centre band at 9.4 T and the corrected intensities
Site designation as for the crystal structure; A = Gaussian smoothing used for simulations; Int. intensity = intensity derived from
the simulations at 9.4 T; CB fraction = the fraction of the (l/2, - l/2) transition intensity in the centre band at 9.4 T calculated
using functions of intensity against v~/Y,Y, in ref. 22, corrected intensity = intensity corresponding to the sodium distribution in
the actual structure.
Site
Na(l)
Na(2)
Nat31

77

siw

(ppm)

$Hz)

15.0 + 0.2
27.0 f 0.2
19.5 rf:0.2

2.52 + 0.02
2.08 + 0.02
4.20 f 0.04

0.67 f 0.02
0.05 + 0.05
0.27 f 0.02

A
(Hz)

Int.
intensity (%)

CB
fraction

Corrected
intensity (%)

150
150
150

53 f 3
31+2
16f4

0.79
0.91
0.30

44+5
22 f 4
34 f 8

T.J. Bastow et al. /Solid State Nucl. Magn. Reson. 3 (1994) 49-57

preach for the elucidation of crystal structures.


The long-range techniques point to disorder in
Na,ZrO, produced by direct solid-state reaction
due to stacking faults along the c-axis. This has
only a minor effect on the NMR spectra that
show locally the structure is well defined even in
the initial sample. Flux annealing removes the
stacking faults increasing the long-range order as
is observed in diffraction and microscopy and
more subtly in the NMR spectra.

4. Conclusion
l0, 23Na and 91Zr NMR spectra have been
reported for Na,ZrO,. A single zirconium site
can be resolved within the static NMR spectrum.
For 0, despite there being three sites in the
crystal structure, two of these are locally identical
as only two separate oxygen resonances are seen
in the ratio 2: 1. For 23Na three sites can be
resolved that are consistent with the structural
distribution 2: 1: 1, with widely differing C, values. On the basis of the intensity distribution and
C, values the resonances can be assigned to
specific sites within the structure. NMR spectra
of the sample prepared by direct solid-state reaction and that flux-annealed in Bi,O, show only
minor differences so that the local coordinations
are very similar in both samples. Diffraction and
electron microscopy data show that flux-annealing removes disorder due to stacking faults. A
methodology combining diffraction, microscopy
and NMR holds much promise for characterising
solid materials.

57

an NMR probe. Bruker Analytische Messtechnik


and Dr. Stefan Steuernagel are thanked for access to the MSL 500.

6. References
1 G. Lang, Z. Anorg. Allg. Chem., 276 (1954) 71.
2 G. Lang, Z. Anorg. Allg. Chem., 348 (1966) 246.
3 Powder Diffraction File (JCPDS), File No. 35-770, Pub.
International
PA, 1992.
4 J. DAns and
1.
5 S.O. Ampian,
6 J.F. Dorrian

Centre for Diffraction

Data, Swarthmore,

J. Loeffler, Z. Anorg. Allg. &em.,

191 (1930)

J. Am. Ceram. Sot., 51 (1968) 607.


and R.E. Newnham, Mater. Res. Bull., 4

(1969) 179.
7 M. Troemel
8
9

10

11
12

and J. Hauck, Z. Anorg. Allg. Chem., 373


f 1970) 8.
E.R. Andrew, ht. Rev. Phys. Chem., 10981) 195.
T,J. Bastow and M.E. Smith, Solid State Nucl. Magn.
Reson. l(1992) 165.
E. Oldfield, C. Coretsopoulos, S. Yang, L. Reven, H.C.
Lee, J. Shore, O.H. Han, E. Ramh and D. Hinks, Phys.
Rev. B, 40 (1989) 6832.
T.J. Bastow and S.N. Stuart, Chem. Phys., 143 (1990) 459.
T.J. Bastow, M.E. Smith and H.J. Whitfield, J. Mater.
Chem., 2 (1992) 989.

13 T.J. Bastow, A.F. Moodie, M.E. Smith and H.J. Whitfield,


J. Mater. Chem.. 3 (1993) 697.

and E. Oldfield. J. Magn.


Reson., 69 (1986) 124.
1.5 O.S. Edwards and H. Lipson, Proc. Roy. Sot. A, 180 (1942)
14 A.C. Kunwar, G.L. Turner

268.
16 A.J.C. Wilson, Proc. Roy. Sot. A, 180 (1942) 277.
17 M.M.J. Treaty, J.M. Newsam and M.W. Deem, Proc. Roy.
SK A, 433 (1991) 499.
18 J.G.P. Binner, R. Stevens and S.R. Tan, Trans. J. Br.
Ceram. Sot., 82 (1983) 98.

19 J.L. Enriquez, P. Quintana and J. Vasquey, Trans. J. Br.


Ceram. Sot., 81 (1982) 118.

20 E. Lippmaa, A. Samoson and M. Magi. J. Am. Chem. Sot.,


106 (1986) 1730.
21 S. Schramm and E. Oldfield, J. Am. Chem. Sot., 106

5. Acknowledgements
M.E.S. thanks the SERC
T service at the University
Parkinson for his help there
(University of Warwick) for

for access to the 14.1


of Edinburgh, Dr. J.
and Prof. R. Dupree
the generous loan of

(1984) 2502.

22 D. Massiot, C. Bessada, J.P. Coutures and F. Taulelle, J.


Magn. Reson., 90 (1990) 231.
23 M.E. Smith and S. Steuernagel, Solid State Nucl. Magn.
Renon., l(1992) 175.

Вам также может понравиться