Sei sulla pagina 1di 8

PHYSICAL REVIEW B

VOLUME 58, NUMBER 22

1 DECEMBER 1998-II

Effect of hydrostatic pressure on the crystal structure and superionic behavior


of lead II fluoride
S. Hull and D. A. Keen
The ISIS Facility, Rutherford Appleton Laboratory, Chilton, Didcot, Oxfordshire, OX11 0QX, United Kingdom
~Received 9 June 1998!
The structural behavior of lead ~II! fluoride has been investigated at pressures up to ;0.9 GPa and temperatures up to ;1000 K using the powder neutron diffraction technique. The observed phase transitions between
fluorite and cotunnite structured polymorphs are consistent with the published p-T phase diagram for PbF2. An
irreversible pressure induced transition from the fluorite structured b phase to the cotunnite structured a phase
occurs at p50.61(5) GPa with an 8.5~1!% decrease in volume. On heating, the irreversible ab transition at
ambient pressure occurs at a temperature of 601~4! K, prior to the superionic bb* transition at ;710 K.
Increasing pressure has no measurable effect on the concentration of the Frenkel defects which characterize the
intrinsic disorder within the superionic b* phase. In contrast to previous studies, we find no evidence for
significant lattice disorder within the denser a phase at elevated temperatures and, on increasing temperature at
pressures in excess of ;0.5 GPa, the superionic transition occurs abruptly at the ab* transition.
@S0163-1829~98!06745-9#

I. INTRODUCTION

Ionic compounds with the fluorite crystal structure ~space


m! have been extensively studied owing to their
group Fm3
high-temperature superionic behavior.1 On heating, compounds such as b -PbF2, CaF2, BaF2, and SrCl2 undergo a
broad transition to the superionic state at a temperature T c
which is approximately 0.8T m , where T m is the melting temperature in degrees K. Above T c , the materials are characterized by extremely high values of ionic conductivity ~typically s ;1 V 21 cm21! which are a consequence of
extensive, dynamic Frenkel disorder within the anion
sublattice.2,3 Of the fluorite structured compounds, potential
commercial applications for materials with high ionic conductivities in solid state battery and sensor technologies4 has
focussed attention predominantly on b -PbF2, because it has
the lowest superionic transition temperature in both absolute
(T c ;710 K) and relative (T c /T m ;0.65) terms.5 However,
the use of PbF2 in electrochemical devices is not straightforward, because the material can also exist at ambient pressure
and temperature in the orthorhombic cotunnite structured a
form ~space group Pnma!.
a -PbF2 is approximately 10% more dense than the fluorite structured b form. At ambient temperature the ba
transition occurs at relatively modest pressures ~;0.5 GPa!
and can be induced by grinding.6 The reverse ab transition can only be achieved by heating to temperatures in excess of ;600610 K ~Ref. 6!. The irreversible nature of the
transformations mean that PbF2 can exist indefinitely under
ambient conditions in either the a form, the b form, or, more
commonly, as a mixture of both phases. Indeed, much of the
early literature describing the conductivity and thermal properties of PbF2 is difficult to interpret because the phase purity
of the samples is not reported. Furthermore, it is not clear
which polymorph is thermodynamically stable at ambient
conditions. Using a Pb/b-PbF2//KF(aq)// a -PbF2 /Pb cell,
Kennedy et al.7 measured an open circuit potential of 11.7
0163-1829/98/58~22!/14837~8!/$15.00

PRB 58

mV, indicating that the b form is more stable by ;78 cal/


mol. However, this conclusion has since been questioned on
the basis of high-pressure differential thermal analysis
~DTA!8 and elastic constant9 data.
Extensive high-pressurehigh-temperature ionic conductivity and DTA measurements on structurally characterized
samples have determined the essential features of the
pressure-temperature ( p-T) phase diagram of PbF2 ~Refs.
610! ~Fig. 1!. On increasing pressure at ambient temperature, the ba transition results in an increase in the ionic
conductivity s by ;102 ~Ref. 6!. However, this feature is

FIG. 1. The p-T phase diagram of PbF2 ~based on Refs. 610!.


The pressure dependence of the superionic transition ( b b * ) is
not known and is shown as a dotted line.
14 837

1998 The American Physical Society

14 838

S. HULL AND D. A. KEEN

believed to result from the formation of extensive numbers


of defects within the a -PbF2 lattice during the reconstructive
transition and the conductivity of a -PbF2 falls rapidly on
increasing pressure and/or temperature away from the ba
boundary due to annealing of these defects.6 With the exception of this artifact, a -PbF2 is a rather poorer conductor than
b -PbF2 ~Ref. 6!. For example, at 423 K and ambient pressure, values of s 5631025 and s 5531026 V 21 cm21
have been reported for the b and a polymorphs,
respectively.7 In addition, the electronic contribution to the
overall conductivity is extremely small, with s e
;10211 V 21 cm21 at 423 K ~Ref. 7!. Frenkel defects are
accepted to be the dominant species responsible for the ionic
conductivity within b -PbF2 ~Refs. 6 and 10!, though data for
the a phase are ambiguous, with the ~limited! ionic conduction mechanism being explained on the basis of anion
vacancies11,12 and anion interstitials.6,10 In addition, the nature of the superionic transition at elevated pressures remains
uncertain. The high-pressure DTA studies of Klement13
showed a diffuse transition in the temperature range 720
770 K at pressures of 0.51.0 GPa and, as pointed out by
Samara,6 this lies within the stability field of a -PbF2 ~Fig.
1!. However, later DTA studies8 questioned the earlier observations and interpreted a kink in the ab phase boundary
at ;0.20 GPa and ;700 K as the intersection of the superionic bb* boundary. Intuitively, extensive lattice disorder
within the cotunnite structured a phase is surprising, since it
is more densely packed and there are no large empty sites to
accommodate interstitial anions. Nevertheless, indirect support for ~limited! superionic behavior within this structure
has been provided by evidence of an anomalous increase in
the ionic conductivity of PbCl2 ~which adopts the cotunnite
structure at ambient pressure! in a narrow temperature range
just below the melting point14 and by observations of high
F2 mobility in a -PbF2 in recent NMR studies.12
The extent of anion disorder within the a and b modifications of PbF2 and their relative structural stability plays a
central role in both enhancing and limiting the technological
applications of the material, For example, it has been shown
that an ~insulating! passivation layer of a -PbF2 can form at
the anode-electrolyte interface, dramatically reducing the operating life of galvanic cells using b -PbF2 ~Ref. 15!. Conversely, a -PbF2 has recently attracted interest as a potential
high-energy particle detector, though only a -PbF2 is found
to scintillate under irradiation, probably because the higher
anion mobility in the b phase causes nonradiative quenching
of the excitations.16 In this paper we report structural studies
of PbF2 at elevated pressure and temperature which probe the
effects of pressure on the superionic behavior within b -PbF2
and investigate the extent of thermally induced disorder
within a -PbF2. The use of hydrostatic pressure is particularly advantageous for the latter because, as illustrated in the
phase diagram ~Fig. 1!, the stability field of a -PbF2 extends
to higher temperatures as pressure is increased. Intuitively,
one might expect any thermally induced disorder to be favored in this region of p-T space.
II. EXPERIMENTAL

The neutron diffraction experiments were performed using polycrystalline samples of PbF2. Commercially available

PRB 58

material of stated purity 99.99% provided by the Aldrich


Chemical Co. was found to be approximately 90% a phase
by volume. This was converted to pure b phase PbF2 by
heating in an inert argon atmosphere at 650 K for 3 h and
then compressed into a platinum capsule 8 mm in diameter
and 25 mm in length. Diffraction data showed no evidence of
any reconversion to a -PbF2 caused by this compaction process. Details of the heatable pressure cell used in this work
can be found elsewhere.17 Diffraction experiments were performed using the Polaris powder diffractometer at the ISIS
facility, U.K.18 Data were collected using detector banks
which cover the scattering angles 85,62u,95 and provide data over the d-spacing range ;0.3,d(),;4.3
with a resolution Dd/d;631023 . Rietveld profile refinement used the program TF12LS and its multiphase
derivative,19 which are based on the Cambridge Crystallographic Subroutine Library.20 In assessing the relative quality of fits to the experimental data using different structural
models the usual x 2 statistic is used, defined by

x 25

(
N
d

~ I obs2I calc! 2
~ s I obs! 2

~ N d 2N p ! .

N d is the number of data points used in the fit and N p is the


number of fitted parameters. I obs and I calc are the observed
and calculated intensities, respectively, and s I obs is the estimated standard deviation on I obs , derived from the counting
statistics.
III. RESULTS
A. Description of the structures of a -PbF2 and b -PbF2

Before discussing in detail the structural behavior of PbF2


on increasing pressure and temperature it is necessary to describe the structures of the two polymorphs. a -PbF2 possesses the orthorhombic cotunnite structure21 ~space group
Pnma,
a56.440 ,
b53.899 ,
c57.651 ,
V
5192.1 3 , Z54! which contains two alternate layers of
atoms perpendicular to the b axis @Fig. 2~a!#. All the atoms
are in Wyckoff 4(c) sites, at 6(x, 41 ,z; 21 2x, 43 , 12 1z),
with x Pb50.2530, z Pb50.1042, x F1 50.8620, z F1 50.0630,
x F2 50.4660, and z F2 50.8460. This leads to a highly distorted close packing of F2, with each Pb21 surrounded by an
irregular coordination shell containing nine anions at distances in the range 2.4053.035 . Anions on F1 sites are
coordinated to 43Pb21 at relatively similar distances
~2.4472.641 ! and an average Pb21-F2 contact of 2.518
. The cation arrangement surrounding each F2 position is
more distorted, with 53Pb21 at distances in the range
2.4053.035 and an average of 2.770 .
In contrast, the cubic fluorite structure of b -PbF2 ~space
m, a55.9273 , V5208.2 3 , Z54! is relagroup Fm3
tively simple22 and can be considered as a simple cubic array
of anions in which the cations occupy alternate cube centres.
Pb21 are situated in the 4(a) sites at (0,0,0)1fc @where fc
denotes the face centering operations 1~ 21 , 12 ,0!, 1~0, 21 , 12 !, and
1~ 21 ,0, 12 !# and F2 in 8(c) positions at ( 41 , 14 , 14 ),( 41 , 14 , 34 )1fc.
As illustrated in Fig. 2~b!, each Pb21 is surrounded by eight

EFFECT OF HYDROSTATIC PRESSURE ON THE . . .

PRB 58

14 839

FIG. 3. Schematic diagram of the dynamic Frenkel defect cluster observed in the superionic phase of fluorite structured compounds such as b -PbF2 ~Ref. 2!. The larger spheres are cations,
with an asterisk labeling the cation at the unit cell origin. The anion
FIG. 2. Schematic diagram of the structures of ~a! cotunnite
structured a -PbF2 and ~b! fluorite structured b -PbF2. The larger
spheres are Pb21 and the smaller ones are F2. In the case of
a -PbF2, the light and dark smaller spheres denote the symmetry
independent F1 and F2 anions, respectively. The ninefold and
eightfold anion coordination shell around Pb21 is illustrated for
a -PbF2 and b -PbF2, respectively.

anions and each F2 lies at the centre of a tetrahedron formed


by 43Pb21. The single Pb21-F2 distance is )a/4
52.566 .
The nature of the thermally induced disorder within superionic b * -PbF2 has been determined by neutron scattering
studies.2 The dynamic Frenkel defects ~illustrated in Fig. 3!
comprise an anion interstitial located in a 48(i) site at
(u,u, 12 ) with u;0.32, which causes the two nearestneighbor anions to relax away from their regular lattice sites
into 32( f ) sites at ~w,w,w! with w;0.32. The location of
the charge compensating vacancy within the defect has been
determined by static energy calculations.23 Detailed powder
neutron diffraction studies subsequently showed that measurable concentrations of these Frenkel defects (n Frenkel
.;0.5%) appear at ;650 K, and that the concentration
increases to n Frenkel;4% at ;900 K before levelling off.3
The suppression of the disorder is probably due to repulsive
interactions between the defects. The onset of disorder is
accompanied by anomalous increases in the both the linear
lattice expansivity

a l5

Frenkel interstitial ~I! is located in one of the 48(i) sites at (u,u, 21 )


with u;0.32. This causes the nearest-neighbor anions at ( 41 , 14 , 14 )
and ~ 41, 14, 34! ~R! to relax away in ^111& directions into two of the
32( f ) sites which are located at (w, 21 2w, 21 2w) and ( 21 2w,w, 21
1w) with w;0.32. The location of the charge compensating anion
vacancy ~V! at ( 41 , 14 , 74 ) has been determined by static energy calculations ~Ref. 23!.

the former as a pressure calibrant.24 These results are illustrated in Fig. 4. Owing to the highly reactive nature of PbF2
at high temperatures, these calibration curves were used to
determine the pressure in subsequent diffraction studies using pure PbF2. On increasing pressure starting with b -PbF2,
the ba transition was observed to occur over the range

1
da
dT a ~ 293 K!

and the anion isotropic thermal vibration parameter B F


58 p ^ u 2 & , where u is the mean thermal displacement from
the lattice site.3
B. Overview of high-pressure behavior of PbF2

The equation of state of both a and b forms of PbF2 was


determined by pressurizing an approximately 1:2 mixture by
volume of NaCl and PbF2 and using the compressibility of

FIG. 4. The compressibility of the a and b polymorphs of PbF2


under hydrostatic pressure, illustrating the bulk moduli of the two
modifications (B 0 ) and the volume change at the ba transition on
increasing pressure (DV b a /V b ).

S. HULL AND D. A. KEEN

14 840

FIG. 5. Neutron powder diffraction data collected from PbF2 on


increasing temperature at p50.29(2) GPa. The ba and ab*
transitions are clearly seen at T;400 and T;740 K, respectively.

0.570.65 GPa. Conversely, on heating a -PbF2 at ambient


pressure the ab modification was observed at 601~4! K.
High-pressurehigh-temperature data on PbF2 were collected in three different runs. The first two started with
b -PbF2 material which was pressurized to 0.25~2! and
0.29~2! GPa ~i.e., below the ba phase boundary! and then
heated in temperature steps of between 8 and 20 K to temperatures of 9001000 K. The evolution of the neutron diffraction pattern for PbF2 at 0.29 GPa is illustrated in Fig. 5.
The third run at a higher pressure of 0.85~3! GPa started with
pure a -PbF2 material and was heated to ;950 K. The p-T
values of the transitions observed in this work are listed in
Table I. However, as illustrated in Fig. 5, it is important to
note that the transitions are not abrupt and there is coexistence of the two phases for typically 2030 K. The sluggish nature of the ab transition has been observed
previously6 and probably results from the highly reconstructive nature of this structural change.
Finally, on a practical note, the definition of the superionic transition bb* in PbF2 is somewhat ill defined. Conventionally, T c is taken to be the temperature at which the
anomalous peak in the specific heat which accompanies the
onset of Frenkel disorder occurs.1 This is roughly coincident
with the peak in the lattice expansivity a l ~Ref. 3!. Whilst
the latter is, in principle, straightforward to measure accurately using neutron diffraction, the increasing encroachment
of the a phase with increasing pressure makes this approach
TABLE I. The observed structural phase transitions between the
a and b polymorphs of PbF2 in this work. At the highest pressures,
the transition on increasing temperature occurs from the a phase to
the superionic fluorite structured phase labeled b*.
Increasing p
Increasing T

ba
ba

Increasing T

ab~b*!

p50.61(5) GPa at T5295(2) K


T5406(21) K at p50.25(2) GPa
T5400(28) K at p50.29(2) GPa
T5601(4) K at p50.1 MPa
T5730(26) K at p50.25(2) GPa
T5742(22) K at p50.29(2) GPa
T5841(16) K at p50.85(3) GPa

PRB 58

FIG. 6. Diagram illustrating ~a! the anisotropic lattice compression of a -PbF2 under hydrostatic pressure and ~b! the anisotropy
thermal expansion of the same phase on increasing temperature at
p50.85(3) GPa. The values in ~a! are relative to the lattice parameters measured at ambient pressure and T5295(2) K @a 0
56.4389(9) , b 0 53.8977(6) , and c 0 57.6459(11) # and
those in ~b! are relative to p50.85(3) GPa and T5295(2) K @a 0
56.418(1) , b 0 53.886(1) , and c 0 57.621(2) #.

unreliable. Instead, we shall consider the extent of disorder


~concentration of Frenkel defects n Frenkel! and infer from this
the behavior of T c with p ~see Sec. IV C!.
C. Phase a -PbF2

As illustrated in Fig. 6~a!, the compression of a -PbF2


under hydrostatic pressure is somewhat anisotropic, with the
c direction being approximately twice as stiff as the other
two axes. Within the accuracy of our measurements, there is
no observable change in the positional parameters x Pb , z Pb ,
x F1 , z F1 , x F2 , and z F2 over the pressure range studied ( p
<;0.9 GPa). On increasing temperature at elevated pressure @see Fig. 6~b!#, the c axis also behaves differently, with
its expansivity being roughly half that observed along a and
b. The behavior of the six positional parameters shows no
dramatic changes, with only x Pb and z F2 both showing a
slight decrease with increasing temperature at p
50.85(3) GPa. The remaining four parameters are essentially independent of temperature. The variation of the three
isotropic thermal vibration parameters B Pb , B F1 , and B F2 is
shown in Fig. 7. As might be expected on mass grounds, the
anion values are somewhat higher than their cation counterpart. However, the thermal vibrations associated with the F2
site are significantly higher than that of F1 over the entire
temperature range investigated. This may be indicative of a
lower potential gradient at the F2 position, as a consequence
of the rather looser surrounding coordination shell of
Pb21, as described in Sec. III A. Attempts to refine the thermal vibrations anisotropically did not produce any significant
improvement in x 2 and were not considered further.
We now turn our attention to the question of anion disorder ~superionic behavior! within a -PbF2. As discussed earlier ~see Sec. I!, there is debate concerning the nature of the
mobile carriers in a -PbF2 ~anion interstitials and/or vacancies! but, to our knowledge, there are no theoretical studies
of possible defect cluster geometries within cotunnite struc-

PRB 58

EFFECT OF HYDROSTATIC PRESSURE ON THE . . .

FIG. 7. Variation of the isotropic thermal vibration parameters


B Pb and B F for both a and b modifications of PbF2 on increasing
temperature at p50.85(3) GPa. The open and closed symbols correspond to the a and b phases, respectively. The solid lines represent the corresponding values of B F ~upper line! and B Pb ~lower
line! for b -PbF2 at ambient pressure ~Ref. 3!.

tured compounds. Furthermore, the structure is rather


densely packed and simple hard sphere considerations
show that there are no obvious empty sites for thermally
activated Frenkel interstitials to occupy ~of the type found in
fluorite structured b -PbF2!. Any such anions would, therefore, require extensive relaxations of the surrounding anion
and cation sublattices. In light of this, we adopt the most
general approach to analyse the diffraction data and allow
the mean occupancy of the two anion sites to vary during the
least-squares refinement process. However, we do not constrain those ions which leave the lattice site to occupy any
specific locations within the lattice. While not physically realistic, this approach gives a valuable first indication of significant lattice disorder and has been extensively used in the
study of other superionic compounds @see, for example, Li2O
~Ref. 25!, CuCl ~Ref. 26!, and AgI ~Ref. 27!#. As illustrated
in Fig. 8, there is no evidence for significant quantities of
anions leaving the lattice sites at any temperature. The excellent quality of the fits to the data ~typically x 2 ;1.1 1.2!
is illustrated for the case of the highest temperature dataset
@ T5805(2) K# at p50.85(3) GPa in Fig. 9. The fit shown
in Fig. 9 has values for the conventional R factors of R p
58.38%, R wp56.54%, R exp56.02%, and R Bragg53.01%.
The corresponding value of x 2 51.18 compares very favorably with the value x 2 51.07 obtained by a modelindependent fit in which only the background, unit cell constants and peak intensities are fitted parameters.
D. Phase b -PbF2

The refined values of the lattice parameter a of b -PbF2


versus temperature are illustrated in Fig. 10. For clarity, only

14 841

FIG. 8. The mean occupancy of the anion sites in a -PbF2 on


increasing temperature at p50.85(3) GPa, determined by leastsquares analysis of the neutron diffraction data.

the data collected at 0.29~2! and 0.85~3! GPa are shown in


the figure, together with the ambient pressure data taken
from our previous work.3 The presence of a -PbF2 over portions of the temperature region does not allow detailed analysis of the lattice expansivity anomaly at ;T c to be made,
though the lower pressure data in Fig. 10 @ p50.29(2) GPa#
suggest that it is still present.
We now consider the question of disorder within the
b -PbF2 phase. Ignoring the effects of thermal vibration, the
structure factors for the fluorite structured modification of
PbF2 are

FIG. 9. The least-squares fit to the diffraction data collected


from a -PbF2 at p50.85(3) GPa and T5805(3) K. The dots are
the experimental points and the line is the calculated ~fitted! profile.
The difference plot ~observed minus calculated! shown in the lower
trace is on the same scale and illustrates the quality of the fit.

S. HULL AND D. A. KEEN

14 842

FIG. 10. The variation of the cubic lattice parameter a of


b -PbF2 on increasing temperature at ambient pressure ~data from
Ref. 3! and pressures p50.29(2) GPa and p50.85(3) GPa.

F4n 54b Pb18b F for ~ hkl ! reflections with


~ h1k1l ! 54n,

F4n12 54b Pb28b F for ~ hkl ! reflections with


~ h1k1l ! 54n12,

F4n61 54b Pb for ~ hkl ! reflections with


~ h1k1l ! 54n61.

The coherent scattering lengths @b Pb59.405 fm and


b F55.654 fm ~Ref. 28!# are almost in the ratio 2:1 and, as a
consequence, the structure factor F 4n12 is small. This is
borne out by the absence of observable ~200! and ~222!
peaks at ambient temperature in Fig. 5. However, as the
compound becomes superionic, a significant fraction of anions leave the lattice sites and the time-averaged value of b F
falls. As a result, b Pb exceeds 23b F and the presence of
measurable intensity at the ~200! position at elevated temperature is a direct illustration that the sample is in its superionic ~b*! state. A more quantitative measure of the extent
of the intrinsic ~thermally induced! lattice disorder is provided by least-squares refinements of the diffraction patterns.
In addition to the lattice parameter a ~shown in Fig. 10!, the
fitted parameters comprised a scale factor, five coefficients of
a polynomial representing the background scattering, a
Gaussian width parameter describing the broadening of the
Bragg peaks, two isotropic thermal vibration parameters B Pb
and B F ~shown in Fig. 7! and the concentration of Frenkel
defects n Frenkel . The latter are determined by allowing the
occupancy of the lattice fluorine site to vary but with the
constraint that those F2 leaving the regular 8~c! position are
distributed over the 48(i) and 32( f ) sites in the ratio 1:2 so
as to represent the Frenkel defect clusters shown in Fig. 3.

PRB 58

FIG. 11. The variation of the Frenkel defect concentration


n Frenkel of b -PbF2 on increasing temperature. The data collected at
pressures of p50.29(2) GPa and p50.85(3) GPa in this work
~closed and open symbols, respectively! are compared with that
determined previously ~Ref. 3! which is shown as a solid line.

Figure 11 illustrates the effect of pressure on the disorder


within b -PbF2, once again taking the ambient pressure data
of Goff et al., for comparison.3 Further discussion of the
results presented in Fig. 11 will be provided in Sec. IV C.
However, in addition to the absence of any appreciable pressure effect on the defect concentration, we find no evidence
for significant changes in the defect geometry at elevated
pressure. In particular, attempts to vary the positional parameters u and w which describe the locations of the interstitial
and relaxed anions provided no significant improvement in
the quality of the fit.
IV. DISCUSSION
A. High-pressure behavior of PbF2

The values of the transition pressures and temperatures


observed in this work ~tabulated in Table I! are generally in
accord with the published phase diagram of PbF2 ~Refs. 6
10! illustrated in Fig. 1. The exception is the ba transition
at ambient temperature, which we observe at a slightly
higher pressure @ p50.61(5) GPa# than previous workers.
However, this may be explained by a more hydrostatic pressure environment in our experiments. The ability to promote
the ba conversion by grinding the material illustrates the
sensitivity of the reconstructive transition to shear stresses.
Although the compressibility data shown in Fig. 4 cover
only a limited pressure region ~especially in the case of
b -PbF2! it is possible to make approximate ~linear! estimates
of B 0,a 5117(4) GPa and B 0,b 564(2) GPa for the bulk
moduli of the a and b polymorphs, respectively, using the
expression @ V 0 2V(p) # /V 0 5 p/B 0 . The latter is in broad
agreement with the value B 0,b 563 GPa at 300 K determined
by Jamieson et al. on the basis of elastic constant data9 and is

PRB 58

EFFECT OF HYDROSTATIC PRESSURE ON THE . . .

FIG. 12. The variation of the width of the Bragg diffraction


peaks of a -PbF2 on increasing temperature at p50.85(3) GPa.

comparable to the value of 57 GPa recently reported for fluorite structured BaF2 ~Ref. 29!. The value for cotunnite structured a -PbF2 is somewhat higher, though this might be expected because the relatively densely packed cotunnite
structure is well suited to low compressibility @for example,
the cotunnite structured phase of ZrO2 has B 0 5332 GPa,
approaching that of diamond, B 0 5442 GPa ~Ref. 30!#. The
volume change associated with the ba transition is found
to be DV/V 0 50.085(1),
in line with other
fluoritecotunnite transformations in CaF2 ~DV/V 0 50.093
at ;10 GPa!, SrF2 ~DV/V 0 50.089 at ;5.5 GPa!, and BaF2
~DV/V 0 50.098 at ;3 GPa! ~see Ref. 31, and references
therein!.
B. Phase a -PbF2

On increasing temperature at modest pressure ~i.e., below


the ba transition!, the diffraction data indicate a significant broadening of the orthorhombic a -PbF2 diffraction lines
immediately above the ba transition. This behavior is visible in Fig. 5 and illustrated in Fig. 12 by the variation of the
Gaussian peak width parameter for a -PbF2 with temperature
at a pressure of 0.85~3! GPa. This observation is consistent
with the tendency of the ionic conductivity to increase at the
ba transition6 due to the formation of large concentrations
of ~static! lattice defects during the reconstructive structural
transition. The subsequent annealing of these defects, observed as a rapid fall in the ionic conductivity,6 is supported
by our observation of a decrease in the Gaussian width parameter as the temperature is increased above the ba transition in Fig. 12.
The anisotropic compression of a -PbF2 under hydrostatic
pressure is clearly illustrated in Fig. 6~a! and the relative
stiffness parallel to the c axis is similar to that observed in
the cotunnite structured phase of BaF2 ~Ref. 29!. Indeed, a
decrease in c/a has been predicted in the recent theoretical

14 843

investigations of PbF2 under pressure by Lorenzana et al.32


This behavior is mirrored in the thermal expansion, which is
relatively small in the c direction @Fig. 6~b!#. With reference
to the crystal structure of a -PbF2 @Fig. 2~a!#, there is no clear
explanation for the anisotropic compression of the orthorhombic unit cell under pressure, because the c axis does not,
for example, correspond to any particularly short F2-F2 contact. The effect of pressure is to increase the distortion of the
cation environment surrounding F1 and decrease it around
F2. In contrast, the refined values for the positional parameters x Pb , z Pb , x F1 , z F1 , x F2 , and z F2 show no exceptional
changes on increasing temperature at p50.85(3) GPa, with
only marginal decreases in the values of x Pb and z F2 .
Analysis of the diffraction data allowing a fraction of the
anions to leave their regular F1 and F2 sites gave no significant improvement in the goodness-of-fit parameter x 2
over that provided by a fully ordered structure. Similarly, the
variation of the unit cell constants and anion thermal vibration parameters show linear behavior as a function of temperature. Therefore, the neutron diffraction data provide no
evidence of superionic behavior within the cotunnite structured a phase of PbF2. Previous ionic conductivity and differential thermal analysis ~DTA! studies13 supporting the onset of extensive lattice disorder within a -PbF2 appear to be
in error, probably owing to uncertainties over the pressure/
temperature calibration and resultant misinterpretation of the
phase~s! presents. In particular, the assignment of a diffuse peak observed in DTA scans to a gradual transition to
a superionic state is not justified, because, as illustrated in
this work and previous studies,6 the ab transitions are
themselves rather gradual.
C. Phase b -PbF2

The three characteristic features of the superionic transition within b -PbF2 observed in a neutron diffraction experiment are the onset of significant Frenkel disorder within the
anion sublattice and anomalous increases in the anion isotropic thermal vibration parameter B F and the lattice parameter
a. The latter leads to a peak in the linear expansivity a l at a
temperature close to the transition temperature T c . However,
as discussed in Sec. III B, the intervention of the a phase at
elevated pressures make the latter two features difficult to
detect with any degree of certainty ~see Figs. 7 and 10!. The
critical information concerning the effect of pressure on the
superionic behavior within b -PbF2 is, therefore, contained in
Fig. 11. In particular, there is no evidence that pressures up
to 0.8 GPa significantly reduce the concentration of Frenkel
defects, despite the reduction in unit cell volume by ;1.5%.
In the context of ionic conductivity measurements at elevated pressures, this effect has been interpreted by Samara33
as evidence that the activation volume for Frenkel defect
formation DV F ~i.e., the contribution to the lattice volume
due to the formation of an F2 interstitial and vacancy! is
relatively low compared to the molar volume V M . This is
presumably a consequence of the relaxations of the nearest
neighbour ions to accommodate the interstitial. Indeed, interpretation of the coherent diffuse neutron scattering data from
b -PbF2, CaF2, and BaF2 in their superionic phase provides
evidence that small displacements of the six next nearestneighbor anions around the defect also occur.2

S. HULL AND D. A. KEEN

14 844

In light of the above discussion, the superionic transition


temperature T c appears to be essentially independent of pressure. As a result, the dotted line in Fig. 1 denoting the bb*
transition is close to horizontal and it is reasonable to assume
that it meets the kink in the ab phase boundary observed at p;0.20 GPa and T;700 K by Klement and
Cohen.8 At pressures greater than p;0.20 GPa the transition
from the a phase on increasing temperature is then directly
to the superionic b* state. This is consistent with the results
of our neutron diffraction data on increasing temperature at
p50.29(2) GPa and p50.85(3) GPa in Fig. 11. We would,
therefore, expect this transition to be accompanied by an
abrupt increase in the ionic conductivity ~i.e., it becomes a
type I superionic transition in the notation of Boyce and
Huberman1! rather than a gradual one ~type II! as observed at
;710 K under ambient pressure.
V. CONCLUSIONS

The results presented in this paper resolve the ambiguity


present in the previous literature concerning the nature of the
thermally induced disorder within the two polymorphs of
PbF2 as a function of p and T. There is no evidence of superionic behavior within cotunnite structured a -PbF2 and the
application of hydrostatic pressure has no significant effect
on the superionic behavior within fluorite structured b -PbF2,
at least to p50.85(3) GPa. While impedance spectroscopy
measurements of the ionic conductivity are more sensitive in
systems with relatively low numbers of mobile ions, uncer-

J. B. Boyce and B. A. Huberman, Phys. Rep. 51, 189 ~1979!.


M. T. Hutchings, K. Clausen, M. H. Dickens, W. Hayes, J. K.
Kjems, P. G. Schnabel, and C. Smith, J. Phys. C 17, 3903
~1984!.
3
J. P. Goff, W. Hayes, S. Hull, and M. T. Hutchings, J. Phys.:
Condens. Matter 3, 3677 ~1991!.
4
S. Chandra, Superionic Solids. Principles and Applications
~North Holland, Amsterdam, 1981!.
5
W. Schroter and J. Nolting, J. Phys. Colloq. 41, 20 ~1980!.
6
G. A. Samara, J. Phys. Chem. Solids 40, 509 ~1979!.
7
J. H. Kennedy, R. Miles, and J. Hunter, J. Electrochem. Soc. 120,
1441 ~1973!.
8
W. Klement Jr. and L. H. Cohen, J. Electrochem. Soc. 126, 1403
~1979!.
9
J. C. Jamieson, M. H. Mangnani, T. Matsui, and L. C. Ming, J.
Geophys. Res. B 5, 4643 ~1986!.
10
J. Oberschmidt and D. Lazarus, Phys. Rev. B 21, 2952 ~1980!.
11
M. Mahajan and B. D. N. Rao, Chem. Phys. Lett. 10, 29 ~1971!.
12
F. Wang and C. P. Grey, J. Am. Chem. Soc. 117, 6637 ~1995!.
13
W. Klement, Jr., J. Phys. C 9, L333 ~1976!.
14
F. E. A. Melo, K. W. Garrett, J. Mendes-Filho, and J. E. Moreira,
Solid State Commun. 31, 29 ~1979!.
15
J. H. Kennedy and R. C. Miles, J. Electrochem. Soc. 123, 47
~1976!.
16
D. F. Anderson, M. Kobayashi, C. L. Woody, and Y. Yashimura,
Nucl. Instrum. Methods Phys. Res. A 290, 385 ~1990!.
17
S. Hull, D. A. Keen, R. Done, T. Pike, and N. J. G. Gardner,
Nucl. Instrum. Methods Phys. Res. A 385, 354 ~1997!.
18
R. I. Smith and S. Hull ~unpublished!.
1
2

PRB 58

tainty over the phase under investigation can arise. Powder


neutron diffraction is, therefore, a valuable technique for the
study of systems such as PbF2 which exhibit extensive structural disorder and complex polymorphism.
The absence of a high ionic conductivity within the a
phase of PbF2 may have important consequences for the synthesis of solid state electrochemical devices. The use of aliovalent doping to enhance the ionic conductivity of b -PbF2 at
temperatures close to ambient ~i.e., reduce T c ! via the formation of extrinsic defects has been widely studied ~see, for
example, Refs. 3436!. However, the effects of doping on
the ba transition at elevated pressure have been relatively
poorly investigated. The introduction of a dopant to form a
nonstoichiometric fluorite structured solid solution would
clearly be advantageous if it also moved the ba transition
pressure to higher values, thereby reducing the likelihood
that quantities of the ~poorly conducting! a form is produced
by grinding and compaction during device manufacture.

ACKNOWLEDGMENTS

The work presented in this paper forms part of a wider


research project investigating the high-pressure behavior of
superionic conductors funded by the Engineering and Physical Sciences Research Council ~reference P:AK:113 C2!. We
are grateful to T. Cooper and N. J. G. Gardner for technical
assistance with the operation of the heatable high-pressure
cell.

W. I. F. David, R. M. Ibberson, and J. C. Matthewman ~unpublished!.


20
P. J. Brown and J. C. Matthewman ~unpublished!.
21
P. Boldrini and B. O. Loopstra, Acta Crystallogr. 22, 644 ~1967!.
22
R. W. G. Wyckoff, Crystal Structures ~Krieger, Malabar, FL,
1982!.
23
C. R. A. Catlow and W. Hayes, J. Phys. C 15, L9 ~1982!.
24
D. L. Decker, J. Appl. Phys. 42, 3239 ~1979!.
25
T. W. D. Farley, W. Hayes, S. Hull, M. T. Hutchings, and M.
Vrtis, J. Phys.: Condens. Matter 3, 4761 ~1991!.
26
S. Hull and D. A. Keen, J. Phys.: Condens. Matter 8, 619 ~1996!.
27
D. A. Keen, S. Hull, W. Hayes, and N. J. G. Gardner, Phys. Rev.
Lett. 77, 4914 ~1996!.
28
V. F. Sears, Neutron News 3, 26 ~1992!.
29
J. M. Leger, J. Haines, A. Atouf, O. Schulte, and S. Hull, Phys.
Rev. B 52, 13 247 ~1995!.
30
J. Haines, J. M. Leger, and A. Atouf, J. Am. Ceram. Soc. 78, 445
~1995!.
31
K. F. Siefert, Fortschr. Mineral. 45, 214 ~1978!.
32
H. E. Lorenzana, J. E. Klepeis, M. J. Lipp, W. J. Evans, H. B.
Radousky, and M. van Schilfgaarde, Phys. Rev. B 56, 543
~1997!.
33
G. A. Samara, Solid State Phys. 38, 1 ~1984!.
34
C. C. Liang and A. V. Joshi, J. Electrochem. Soc. 122, 466
~1975!.
35
C. Lucat, G. Campet, J. Claverie, J. Portier, J. M. Reau, and P.
Hagenmuller, Mater. Res. Bull. 11, 167 ~1976!.
36
J. M. Reau, P. P. Federov, L. Rabardel, S. F. Mater, and P.
Hagenmuller, Mater. Res. Bull. 18, 1235 ~1983!.
19

Potrebbero piacerti anche