Sie sind auf Seite 1von 7

J. Am. Ceram. Soc.

, 94 [6] 1817–1823 (2011)


DOI: 10.1111/j.1551-2916.2010.04319.x
r 2011 The American Ceramic Society

Journal
Conductivity Enhancement in Lanthanum Phosphates
Satyajit R. Phadkew and Juan C. Nino
Materials Science and Engineering Department, University of Florida, Gainesville, Florida 32611

Current proton exchange membranes (PEM) are based on poly- As such, the mechanism of proton conduction in these materials
meric materials such as perfluorosulfonic acid (Nafiont), which is much less understood.
have an upper limit to their temperature of operation (o1001C). In the phosphate-rich region of the La2O3–P2O5 binary-phase
To overcome this limitation, ceramic PEM materials are being diagram, four stoichiometrically different lanthanum phosphate
investigated for transportation applications, where operating materials exist, namely LaPO4 (orthophosphate), La2P4O13
temperatures in the range of 2001–6001C are desired. In this (tetraphosphate), La(PO3)3 (metaphosphate/polyphosphate),
study, the conductivity behavior of lanthanum orthophosphate and LaP5O14 (ultraphosphate).21 The difference in stoichiome-
(LaPO4) has been compared and contrasted with lanthanum ul- try in the above compounds is accommodated crystallochemi-
traphosphate (LaP5O14) in order to better understand crystal cally by changes in the degree of interlinking or condensation
structure—proton conduction relationships in ceramic materi- (trend shown in Fig. 1) of the phosphate anion. This leads to a
als. The conductivity of the lanthanum phosphates (doped and different yet closely related crystallochemical arrangement. For
undoped) was measured using impedance spectroscopy in the example, in LaPO4, the phosphate anion consists of isolated
temperature range 3001–6001C. The conductivity of 5 mol% tetrahedra of PO4 but in LaP5O14, the phosphate anion consists
Sr21-doped LaP5O14 (1.01  104 S/cm, 6001C) was found to of condensed corner sharing PO4 tetrahedra arranged in a rib-
be an order of magnitude higher than similarly doped LaPO4 bon-like morphology.
(7.00  106 S/cm, 6001C), which is a well-investigated proton Owing to the nature of bonding and marked structural differ-
conducting material. In addition, it was observed that the acti- ences in the arrangement of PO4 tetrahedra, both LaPO4 and
vation energy for protonic conduction was much lower for doped LaP5O14 present quite different avenues and prospects for pro-
LaP5O14 (0.8070.01 eV) as compared with LaPO4 (1.0970.01 ton conduction in the material, which will be discussed in detail
eV). A hypothesis relating the oxygen-to-oxygen ion distance in later in the text.
a material to the activation energy for proton conduction is pre- It is well known that in LaPO4 acceptor, doping leads to the
sented and the experimental results obtained have been critically formation of charge compensating oxygen ion vacancies, which
examined on the basis of the hypothesis and other relevant lit- is equivalent to formation of a pyrophosphate group.18 The
erature. From this analysis, it is shown that the condensed pyrophosphate is formed by the local condensation of two ad-
nature of the phosphate anion in LaP5O14 can provide low- jacent PO4 tetrahedra, one of which is missing is an oxygen ion.
energy avenues for proton transport within the material leading In the presence of ambient water vapor, the hydrolysis of the
to enhanced conductivity in the material. Limitations of the cur- pyrophosphate leads to the formation of protonic defects as
rently proposed model for proton conduction along with some represented by the equation below:
other plausible explanations for the conductivity enhancement
have also been discussed. H2 O þ ½P2 O7 4 ! 2½HPO4 2 (1)

These protons (which are associated to an oxygen ion) are


I. Introduction free to hop around (Grötthuss mechanism) between neighboring
oxygen ions in the structure. Under suitable electrochemical
P ROTON exchange membranes (PEMs) have been widely in-
vestigated in recent years for hydrogen fuel cell applications,
which are sometimes referred to as PEM fuel cells. From the
gradient, they can diffuse along the gradient leading to charge
transport within the material. Owing to the nature of proton
transport within solid-state materials, it is evident that the spa-
conduction point of view, PEMs are solids that exhibit high
tial arrangement of oxygen ions and hence of PO4 tetrahedra
protonic conductivity. Currently used PEMs are based on poly-
can have a significant effect on the proton conductivity.
meric materials such as perfluorosulfonic acid, which have
Because the diffusion of charged species (proton jump that
an upper limit to their temperature of operation (o1001C).1,2
leads to electrical conduction in this case) is a thermally acti-
This restricts their use in transportation applications where
vated process, it generally follows the well-known Arrhenius
higher temperatures of operation are desirable. Thus, for achiev-
relationship as shown below:
ing intermediate temperatures of operation, numerous investi-
gations (in the last decade or so) have focused on ceramic-based  
materials.3–12 Ea
sT ¼ A exp  (2)
In this respect, several different orthophosphate materials kT
have been investigated as PEMs, including RPO4 (R 5 La, Sm,
Nd, Ce), Ba3Ce(PO4)3, and some pyrophosphate materials such where s is the conductivity, T is the absolute temperature, A is
as SnP2O7 and TiP2O7.13–18 However, the number of studies on the preexponential, k is the Boltzmann’s constant, and Ea is the
higher condensed phosphate materials such as metaphosphates activation energy for the proton jump. It is clear from the above
(A(PO3)n) and ultraphosphates (A3(P5O14)n) is quite limited.19,20 equation that a higher preexponential term and a lower activa-
tion energy is required for fast proton conduction.
T. M. Gur—contributing editor It has been noted previously from computational investiga-
tions on proton-conducting perovskite-type oxide materials that
a lower oxygen-to-oxygen ion distance results in a smaller value
of the activation energy for the proton jump.22,23 Hereby the
Manuscript No. 27793. Received April 4, 2010; approved November 9, 2010. same concept for proton conduction is extended to phosphate
This work was financially supported by the U. S. National Science Foundation (NSF) as
part of the CBET-0730900 grant. materials. A qualitative graphical demonstration of the above
w
Author to whom correspondence should be addressed. e-mail: satyajit.ufl@gmail.com statement is presented in Fig. 2. The graph was plotted using the
1817
1818 Journal of the American Ceramic Society—Phadke and Nino Vol. 94, No. 6

rier (Ea) to jump to the adjacent potential well (close to the


neighboring oxygen ion). In the central region between the
oxygen ions, the potential is high because the proton is far
away from either of the ions. As seen in Fig. 2, lowering of the
oxygen-to-oxygen ion distance leads to a reduction in the height
of the activation energy barrier (assuming the other factors
affecting the shape of the well remain unchanged). It is thereby
here proposed that tailoring a material to attain a lower jump
distance for the proton should lead to a smaller activation en-
ergy and consequently higher proton conductivity, provided all
other relevant factors in the material are minimally disturbed.
Based on this, a scientific evaluation of the above crystallo-
chemical approach toward the development of enhanced
proton-conducting materials is presented by contrasting the
properties of strontium-doped LaPO4 with identically doped
LaP5O14. The obtained results have been critically analyzed on
the basis of the proposed model for proton conduction and dis-
cussed in the background of other relevant literature available.

II. Experimental Procedure


(1) Synthesis
Single crystals of undoped LaP5O14 were synthesized by precip-
itation from concentrated phosphoric acid solutions as previ-
ously described by Park and Kreidler.21 To prepare doped single
crystals (1 and 5 mol% Sr21), the initial mixture was prepared
by mixing 1 g of lanthanum oxide (Alfa Aesar, 99.99%, Ward
Fig. 1. La2O3–P2O5 binary-phase diagram.21 Hill, MA) and the appropriate relative weight of strontium ox-
ide (Alfa Aesar, 99.99%) with 30 mL of 85% phosphoric acid
Morse potential, which is generally used to model the O–H in- and allowed to stand at 3001C for about 4 h until the oxide was
teraction in proton-conducting materials.24,25 Mathematically, completely dissolved and a transparent solution was formed.
the equation is represented as: The above solution was then transferred to a furnace preset at
3001C. After transferring the contents to the furnace, the tem-
perature was allowed to increase to 6501C at a ramp rate of
yðrÞ ¼ Do ð1  expðbðr  ro ÞÞ2 (3) 1001C/h and the solution was allowed to stand there for a period
of 7–10 days. All of the above-described reactions were carried
where r is the separation between the oxygen ion and the proton, out in covered vitreous carbon crucibles, which are resistant to
Do is the depth of the potential well, ro is the equilibrium proton- fused polyphosphoric acid. After cooling, the excess liquid (pol-
to-oxygen distance, and b is a constant that depends on the force yphosphoric acid) forms a gel-like transparent semisolid, which
constant of the proton–oxygen interaction. The numerical val- can be dissolved by placing the crucible in hot water. Upon dis-
ues of the constants required for the Morse function are adapted solution of the gel, rhombus-shaped LaP5O14 crystals in the
from Ruiz-Trejo et al.26 A detailed description of the parameters form of thin platelets were recovered from the bottom of the
can be found elsewhere.27 According to the interaction function, crucible. The obtained crystals were washed several times with
there is a potential well close to the oxygen ions while the value water and then dried in air and stored in a desiccator for further
of the potential stabilizes to a constant value at larger distances experimentation.
from the oxygen ion. Adding up the potentials for the proton To synthesize LaPO4 samples with the same dopant compo-
interaction with two neighboring oxygen ions results in a two- sition as LaP5O14, an initial mixture of lanthanum nitrate
potential well system as shown in Fig. 2. The proton which is (La(NO3)3  6H2O, 99.9%, Alfa Aesar) and strontium nitrate
initially located in one of these wells has to cross an energy bar- (99.9%, Alfa Aesar) was dissolved in deionized water to produce
a 1M solution. Following the dissolution process, diammonium
hydrogen phosphate ((NH4)2HPO4, 98%, Alfa Aesar) was
added to this solution drop-wise, which lead to the formation
of a white precipitate. The solution was allowed to stir for 2 h
after which the precipitate was removed from the solution by
filtration. The obtained precipitate was then dried at 1201C and
ground with mortar and pestle under ethanol. Following calci-
nation of the powder at 10001C for 2 h using ramp rate of
2001C/h, the samples were pressed into cylindrical pellets (using
polyvinyl alcohol as binder) and sintered at 14001C for 10 h
using a ramp rate of 2001C/h.

(2) Phase Analysis (X-Ray Diffraction (XRD), Scanning


Electron Microscopy (SEM), Electron-Dispersive
Spectroscopy (EDS))
A small part of the samples was crushed to powder with a mor-
tar and a pestle under acetone. These powders were then sieved
through a 212 mm sieve and taken for XRD analysis to verify
phase purity. In addition, XRD was also performed on single
crystals of the prepared samples. Previous to all further exper-
Fig. 2. The effect of reducing the oxygen–oxygen ion distance on the imentation, the crystals were dried in air at 6001C for 3 h. This
barrier for proton jump (or activation energy). was performed to remove any remaining phosphoric acid that
June 2011 Conductivity Enhancement in Lanthanum Phosphates 1819

could have been sticking to the surface sample because the crys-
tals were grown from a phosphoric acid melt. Initial samples
that were not subjected to this heat treatment showed a very
high P:La ratio as compared with the expected value in EDS
analysis.
SEM and EDS were performed on a JEOL 6400 microscope,
(JEOL, Peabody, MA) on samples that were sputter coated with
gold. An accelerating voltage of 10 kV was used during analysis.

(3) Electrochemical Impedance Spectroscopy (EIS)


For EIS measurements, the LaP5O14 crystals were sputter
coated on parallel (100) surfaces with gold using sputter coat-
ing technique. To prevent the sputter coater from coating the
sides of the sample, a masking paper with a hole of appropriate
dimensions was used to cover the sample during the coating
procedure. For conductivity measurement on LaPO4, the sin-
tered samples were coated on parallel surfaces with platinum
paste and sintered at 9001C for 1 h. Platinum wires (99.9%,
0.127 mm diameter) were attached to the electroded surfaces of
the pellet or crystal using silver paste (part #5063, SPI Supplies,
West Chester, PA). Three samples with each of the undoped and
doped (1 and 5 mol% Sr21) compositions were used for con-
ductivity measurement. For impedance measurements at differ-
ent temperatures, the samples were placed inside a quartz
reactor itself placed inside a tube furnace. EIS measurements
were done in the temperature range of 3001–6001C using the
two-point probe technique (Solartron SI 1260, Solartron Ana-
lytical, Hampshire, U.K.). To minimize errors resulting from
temperature variation inside the furnace, a thermocouple placed
very close to the sample was used to monitor the temperature. It
is important to note here that all the samples were preheat
treated at 6001C for 3 h to get rid of any phosphoric acid on the
sample surface. In samples not subjected to this heat treatment
erroneously high values of conductivity were obtained, and the
conductivity showed a decreasing trend with increase of tem-
perature beyond 2001C (most likely due to evaporation of sur-
face phosphoric acid). The amplitude of the alternating signal
during the impedance measurement was 1000 mV. Frequency
sweeps were taken in the range of 32 MHz–0.1 Hz and fitting of Fig. 3. (a) Experimental powder and single-crystal diffraction pattern
of LaP5O14 compared with the theoretical powder pattern. (b) Schematic
the obtained spectra was carried out using ZViewt software showing direction of conductivity measurement with respect to orienta-
(Scribner Associates, Southern Pines, NC). tion of the phosphate anion in the structure. Also shown is the orien-
tation of crystal during single-crystal X-ray diffraction data collection.

III. Results and Discussion


(1) Phase Analysis in Fig. 3(b). Conductivity measurements were made along the
The experimental powder diffraction pattern for undoped [100] direction as well as in a transverse direction for the 1 mol%
LaP5O14 is compared with the theoretical pattern for the same Sr-doped sample to investigate the effect of structural anisot-
in Fig. 3(a). The theoretical pattern was generated (using Crystal ropy on the conductivity.
Diffract software) from previously reported structural data.28
All the diffraction peaks observed corresponded to the lantha-
num ultraphosphate phase, indicating that the prepared samples
were phase pure. No secondary phases could be detected from
the powder diffraction pattern of all undoped and doped
LaP5O14 crystals. Similarly, it was confirmed that synthesized
LaPO4 samples were also phase pure. The geometrical density
of the undoped crystals measured was 3.23 g/cm3, while the
Archimedes density was 3.20 g/cm3. These values are close to the
theoretical density, which is 3.25 g/cm3.
The single-crystal XRD pattern of undoped LaP5O14 crystals
is shown in Fig. 3(a). For XRD data collection, the crystal was
placed on the holder with its surface with maximum area facing
the surface of the holder. The relative orientation of the crystal
with respect to source and detector is shown in Fig. 3(b). In this
orientation, diffraction peaks corresponding to the (100) plane
(from several orders of reflection) could be seen in the XRD
pattern. Similarly, peaks from the (100) plane were also ob-
served for 1 and 5 mol% Sr-doped LaP5O14. Thus, it was con-
firmed that the main facet of the crystal was the (100) plane for
both the doped and the undoped crystals. The relative orienta-
tion of the phosphate anion ribbons with respect to the (100) Fig. 4. Scanning electron microscopic image of single crystal of
plane in the crystal can be seen in the schematic representation LaP5O14.
1820 Journal of the American Ceramic Society—Phadke and Nino Vol. 94, No. 6

SEM image of a typical sample of undoped LaP5O14 crystal is


shown in Fig. 4. Undoped LaP5O14 crystals grew in the form of
rhombus-shaped flat sheets. In the energy-dispersive spectra
(EDS) obtained from the undoped and the doped LaP5O14 sam-
ples, peaks corresponding to La, P, O, and Au (because samples
were sputter coated with Au previous to EDS) were detected.
The P/La ratio for both the doped and the undoped samples was
B5 as expected for LaP5O14. The signal from strontium atoms
could not be distinctly observed because their expected concen-
tration is 0.05 at.%, which is below the detection limit of the
instrument (B1 at.%).

(2) Impedance Spectroscopy


Single crystals with three different compositions, undoped
LaP5O14, 1 and 5 mol% Sr-doped LaP5O14, were grown for con-
ductivity measurements. Typical Nyquist plots for 5 mol%
Sr-doped LaP5O14 samples at two different temperatures are
shown in Fig. 5. All the plots obtained for all samples of
LaP5O14 exhibited two impedance arcs corresponding to the Fig. 6. Conductivity of undoped, 1 and 5 mol% Sr-doped LaPO4 (open
bulk impedance (at higher frequency) and the electrode imped- symbols) and LaP5O14 (closed symbols).
ance (lower frequency). In general, ceramic materials exhibit a
three-arc response, where each arc corresponds to bulk, grain
boundary, and electrode impedance in decreasing order of For the 1 mol% Sr-doped LaP5O14 samples, measurements
frequency.29 However, due to the absence of grain boundaries were made in two different crystal orientations; that is along the
in the single crystals of LaP5O14, we observe a two-arc response phosphate anion ribbon direction and along the transverse di-
in the Nyquist plot. The capacitance associated with an imped- rection (indicated in Fig. 8(b)). The activation energy calculated
ance arc can be calculated by knowing the peak frequency of the for protonic conduction along the transverse direction is slightly
arc (shown in Fig. 5) and the resistance associated with it higher (0.7670.02 and 0.8970.02 eV) as shown in Table I.
(obtained by fitting data with an equivalent circuit). The capac- However, the activation energy values for measurements in ei-
itance associated with the bulk impedance was found to be in- ther orientation are still much lower as compared with the ac-
dependent of the temperature and the calculated value was tivation energy obtained for LaPO4 with the same dopant
3.4170.11 pF. The value is similar to the capacitance for bulk concentration (1.0470.01 eV).
impedance for other ionic conductors.29,30 All the Nyquist plots It can also be observed from Fig. 6 that a greater enhance-
for all the LaPO4 samples exhibited an additional arc at inter- ment in conductivity is observed in LaP5O14 as compared with
mediate frequency representing grain-boundary impedance in LaPO4 in moving from 1 to 5 mol% dopant concentration. This
conformity with the polycrystalline nature of the samples. may be due to differences in the homogeneity of dopant distri-
As seen in Fig. 6, the conductivity increased as the dopant bution within the single crystals of LaP5O14 as compared with
content was increased up to 5 mol%. The highest conduc- the sintered samples of LaPO4. Considering the synthesis pro-
tivity measured for the 5 mol% Sr-doped LaP5O14 samples cedure for LaPO4 and given the polycrystalline nature of the
was 1.01  104 S/cm at 6001C. Conductivity measurements sample, dopant segregation at the grain boundary is probable.
on LaPO4 exhibited a similar trend of increasing conductivity This would occur at the cost of the dopant concentration in the
with increasing dopant concentration. However, for all compo- bulk grain leading to a poorer enhancement in bulk conductivity
sitions synthesized, the conductivity of LaP5O14 samples was upon acceptor doping. Indeed, dopant segregation and second-
higher than the respective LaPO4 samples. The maximum con- ary-phase formation has been observed by Norby and col-
ductivity measured for the 5 mol% Sr-doped LaPO4 sample was leagues previously in doped samples of LaPO4.31 On the
7.00  106 S/cm at 6001C. For all the synthesized composi- contrary, the LaP5O14 crystals were grown slowly from a phos-
tions, the measured activation energies were comparatively phoric acid melt containing dissolved oxides of the host and the
much higher for LaPO4 as summarized in Table I. These results dopant cation, which should have led to a more homogeneous
will be discussed in more detail in the next section. distribution of the dopant. Another possibility to consider is the
relative solubility limit for dopant incorporation, which may be
higher for LaP5O14 as compared with LaPO4. Higher dopant
incorporation would lead to a higher protonic defect concen-
tration in the bulk and hence a higher conductivity.

Table I. Obtained Activation Energies for Three Different


Compositions of LaPO4 and LaP5O14
LaPO4 LaP5O14

Activation
Material energy (eV) Material Activation energy (eV)

Undoped 1.2770.03 Undoped 0.8570.01


1 mol% Sr 1.0470.01 1 mol% Sr 0.7670.02/
0.8970.02 (t)
5 mol% Sr 1.0970.01 5 mol% Sr 0.8070.01
Fig. 5. Typical Nyquist plots for 5 mol% Sr-doped LaP5O14 samples at ‘‘t’’ indicates conductivity measurement along the transverse direction as
3001 and 4001C. indicated in Fig. 8(b).
June 2011 Conductivity Enhancement in Lanthanum Phosphates 1821

Fig. 7. View of LaPO4 along the c-axis. La31 ions shown in orange and
PO4 tetrahedra shown in green.32

(3) Model for Conductivity Enhancement


In this section, an attempt has been made to explain the ob-
served differences in the conduction behavior of LaPO4 and
LaP5O14 on the basis of a crystallochemical criterion by closely
examining the crystal structure of the two materials. The rele-
vant ion–ion distances for LaPO4 were measured from the crys-
tal structure, which was prepared in CrystalMakert using
previously reported structural data.32 The crystal structure of
LaPO4 as seen along the c-axis is shown in Fig. 7. It consists of Fig. 8. (a) View of LaP5O14 along the [001] direction. La31 ions shown
in orange, PO4 tetrahedra shown in green and oxygen ions are shown in
isolated tetrahedra of PO4 that are held together by La31 cat-
red. (b) Segment of ribbon composed of PO4 tetrahedra as viewed along
ions. Each La31 cation is coordinated with nine neighboring the [010] direction. Arrow indicates direction of transverse conductivity
oxygen ions. Within a single tetrahedron, the average oxygen– measurement.
oxygen ion distances is 2.5113 Å, while the smallest intertetra-
hedron being 2.8410 Å as summarized in Table II.
The structure of LaP5O14 is isomorphic with other ultraphos-
phates such as EuP5O14 and GdP5O14, and crystallizes in a between adjacent oxygen ions, can be expected to be similar.
monoclinic crystal system (space group: P21/c (No. 14), However, in LaPO4, the jump distance is much higher when the
a 5 8.8206 Å, b 5 9.1196 Å, c 5 13.1714 Å, and b 5 90.6611).28 proton has to hop between oxygen ions belonging to two neigh-
In LaP5O14, the phosphate anion structure is much more com- boring tetrahedra. On the contrary, in LaP5O14, the condensed
plex compared with LaPO4. Although the basic building blocks nature of the tetrahedra in the phosphate anion allow for an
composing the phosphate anion are still PO4 tetrahedra, they are easier pathway for conduction (see Table II). Thus, a migrating
arranged in a regular ribbon-like structure by corner sharing an proton will encounter a smaller jump distance and hence a
oxygen ion. The crystal structure of LaP5O14 is shown in smaller energy barrier (activation energy) for proton conduc-
Fig. 8(a). The ribbon-like strands of PO4 tetrahedra extend tion, which is in agreement with our experimental results.
along the [100] direction and are held together by La31 ions However, one must treat the above analysis with some cau-
as in LaPO4. A small segment of the anion shown in Fig. 8(b) tion, as there could be some obvious problems with treating
shows the manner in which the tetrahedra are interconnected. the structure as a static rigid lattice. For example, it has been
The La31 ions are coordinated with eight neighboring oxygen observed previously in the case of perovskite oxide materials
ions. Within a single tetrahedron, the average oxygen–oxygen that the picture at the atomic level is significantly distorted by
ion distances are 2.5150 Å, which is very similar to the values the local alterations caused in the material due to atomic vibra-
obtained for LaPO4. tions.22 Compared with the perovskite oxides, the situation in
The basic difference between the two structures lies in the these complex phosphate materials could be much more intri-
isolated versus condensed structure of the phosphate anion, cate. Nonetheless, our current analysis, which relies on the
which presents vastly different avenues for proton transport average spatial arrangement of ions in a lattice (as can be
within the material. obtained from XRD measurements), can be used as the funda-
The various possibilities for proton migration in LaPO4 and mental basis for more detailed investigations in the future.
LaP5O14 have been graphically depicted in Fig. 9. In both struc- In addition to the justification presented above, there could
tures, La31 cations are not shown for the purpose of clarity. The be other factors that also affect the mobility of protonic defects
jump distances for the migrating proton are also depicted in the (and hence the activation energy) such as protonic defect–
images for both the materials. Owing to the close dimensional dopant site association (OHo    Sr0 La ). There have been com-
resemblance of the PO4 tetrahedra in both the phosphates, the putational investigations on such defect interactions for perovs-
energetics of proton migration within a tetrahedron, which in- kite-type oxide conductors, where it has been mentioned that
cludes rotational motion around a single oxygen ion or a jump higher interaction energy could lead to a lower conductivity by
defect trapping.24,33 In case of lanthanum phosphates, higher
interaction energy for the proton–dopant association in LaPO4
Table II. Comparison of Mean Inter- and Intratetrahedral could be a potential explanation for the lower conductivity and
Distances in Oxygen–Oxygen Ion Distances in LaPO4 and higher activation energy observed as compared with LaP5O14.
LaP5O14 However, to the best of our knowledge, no such studies have
been reported for proton-conducting phosphate materials for us
Mean oxygen–oxygen ion distance (Å) to present such a comparison.
According to DFT studies by Yu and De Jonghe34 on ener-
System Intratetrahedra Intertetrahedron getics of proton migration in LaPO4, both intertetrahedral and
intratetrahedral proton transfer processes have almost similar
LaPO4 2.511370.06 2.841070.04
but relatively high activation energies (B0.8 eV). However, the
LaP5O14 2.515070.05 — activation energy for rotational motion around an oxygen ion is
1822 Journal of the American Ceramic Society—Phadke and Nino Vol. 94, No. 6

doping can be expected to lead to additional local condensation


in the ultraphosphate structure. The locally condensed phos-
phate group can then by hydrolyzed in the presence of ambient
water vapor to form protonic defects in the material.
In this study, it has been observed that the condensation of
the phosphate anion facilitates the proton transfer process in the
material leading to the enhanced proton conductivity. Although
the obtained conductivity is otherwise modest, it shows a proof
of concept which can be applied to other phosphate systems
(Ln2O3–P2O5, Ln 5 lanthanides) and other material systems
such as germanates, which also exhibit a rich and diverse anion
chemistry similar to the phosphates.38–41

IV. Conclusion
Synthesis of lanthanum ultraphosphate (LaP5O14) single crystals
of three different compositions (undoped, 1 and 5 mol% Sr21
doped) was carried out by precipitation from concentrated
phosphoric acid solutions. Phase purity of the prepared crystals
was established by comparing the experimental powder diffrac-
tion pattern with the theoretical pattern and also from quan-
titative EDS analysis. Conductivity measurements were per-
formed on crystal samples of all the above-prepared composi-
tions along the [100] direction in the temperature range of 3001–
6001C using EIS. Within the 1 mol% Sr-doped LaP5O14 sam-
ples, conductivity was also measured along the transverse direc-
tion. The obtained results were critically compared and
contrasted with the measured conductivity of identically doped
lanthanum orthophosphate (LaPO4) samples to better under-
stand crystal structure–property relationships in lanthanum
phosphate materials. For all the compositions characterized,
the LaP5O14 samples showed a higher conductivity and lower
activation energy as compared with the respective LaPO4
samples. The conductivity of 5 mol% Sr21-doped LaP5O14
Fig. 9. Proton jump distances in (a) LaPO4 and (b) LaP5O14. (1.01  104 S/cm, 6001C) was an order of magnitude higher
than identically doped LaPO4 (7.00  106 S/cm, 6001C), while
comparatively much lower (0.15 eV). Such low-energy values for the activation energy for protonic conduction was lower
rotational motion are consistent with other computational (0.8070.01 eV) as compared with (1.0970.01 eV). To explain
investigations on perovskite oxides.35–37 The rotational motion the differences in observed conduction behavior of the two ma-
of a proton does not contribute to proton transport directly but terials, a hypothesis relating the proton jump distance in a ma-
can substantially assist in the process in LaP5O14. Essentially, terial to the activation energy was presented. Such a correlation
this means that an intertetrahedral proton jump, which has a has been reported previously in the perovskite-type oxide mate-
high-energy barrier in LaPO4, can be substituted for by an rials, which was extended to phosphate materials. The obtained
energetically inexpensive low-barrier rotation around an oxygen results were critically analyzed on the basis of the proposed
ion in LaP5O14, thus leading to an overall lower activation hypothesis and other relevant literature available. From the anal-
energy in LaP5O14. ysis it was shown that the condensed nature of the phosphate
Because the corner-sharing PO4 tetrahedra in LaP5O14 form a anion, which forms a three-dimensional networked structure in
three-dimensional networked structure (as seen in Figs. 8(a) and LaP5O14 renders, is more favorable for achieving enhanced pro-
(b)), multiple paths are available for the protons to travel along ton conduction. Some other plausible explanations for the same
different directions in the material. This can probably explain have also been discussed along with possible limitations of the
the similar activation energy calculated for protonic conduction currently proposed model for proton conduction.
along two different directions in 1 mol% Sr-doped LaP5O14
(0.7670.02 and 0.8970.02 eV).
With respect to the mechanism of proton conduction in un- References
1
doped ultraphosphates, there is an alternate model according to S. M. Haile, D. A. Boysen, C.R. I. Chisholm, and R. B. Merle, ‘‘Solid Acids as
which proton conduction occurs along chemically adsorbed wa- Fuel Cell Electrolytes,’’ Nature, 410 [6831] 910–3 (2001).
2
S. Ishimaru, M. Togawa, R. Ikeda, T. Shimizu, E. Shinohara, and Y. Umemura,
ter molecules which lie all along the condensed phosphate anion ‘‘Complex Impedance and Solid-State NMR Studies on a Novel High Proton
in the material structure rather than along the phosphate anion Conductor Na-RUB-18,’’ Bull. Chem. Soc. Jpn, 79 [4] 656–9 (2006).
itself.20 This could probably explain the measured conductivity 3
I. Brach, D. J. Jones, and J. Roziere, ‘‘The Crystal Structure of RbHSeO4: A
in the undoped material reported at low temperatures, but it Neutron Diffraction Study of the Paraelectric Phase,’’ J. Solid State Chem., 48,
401–6 (1983).
does not explain how protonic conduction can be observed 4
T. Fukui, S. Ohara, and S. Kawatsu, ‘‘Conductivity of BaPrO3 Based Perovs-
in phosphates at much higher temperatures (B7001C), where kite Oxides,’’ J. Power Sources, 71 [1–2] 164–8 (1998).
5
chemically adsorbed water does not exist in the material. Also, M. Fetchtelkord, ‘‘Proton Dynamics in Letovicite, (NH4)3H(SO4)2: A 1H and
because the material measured by Hammas and colleagues is 14N NMR Spectroscopic Study,’’ Solid State Nucl. Magn. Reson., 17, 76–88
(2000).
undoped (and it does not contain any structural protons), it is 6
B. Gross, S. Marion, R. Hempelmann, D. Grambole, and F. Herrmann, ‘‘Pro-
unclear how a proton would be incorporated in the material ton Conducting Ba3Ca1.18Nb1.82O8.73/H2O: Sol–Gel Preparation and Pressure/
leading to the formation of a hydronium ion from a water Composition Isotherms,’’ Solid State Ionics, 109 [1–2] 13–23 (1998).
7
molecule. H. H. Huang, M. Ishigame, and S. Shin, ‘‘Protonic Conduction in the Single-
Crystals of Y-Doped SrZrO3,’’ Solid State Ionics, 47 [3–4] 251–5 (1991).
Although not focus of the current study, it can be envisioned 8
S. Shin, H. H. Huang, M. Ishigame, and H. Iwahara, ‘‘Protonic Conduction in
that the defect chemistry in LaP5O14 is quite similar to LaPO4.18 the Single-Crystals of SrZrO3 and SrCeO3 Doped with Y2O3,’’ Solid State Ionics,
Analogous to LaPO4, oxygen deficiencies created from acceptor 40–1, 910–3 (1990).
June 2011 Conductivity Enhancement in Lanthanum Phosphates 1823
9 25
K. C. Liang and A. S. Nowick, ‘‘High-Temperature Protonic Conduction in K. D. Kreuer, W. Munch, M. Ise, T. He, A. Fuchs, U. Traub, and J. Maier,
Mixed Perovskite Ceramics,’’ Solid State Ionics, 61 [1–3] 77–81 (1993). ‘‘Defect Interactions in Proton Conducting Perovskite-Type Oxides,’’ Ber. Bunsen
10
P. Murugaraj, K. D. Kreuer, T. He, T. Schober, and J. Maier, ‘‘High Proton Ges. Phys. Chem. Chem. Phys., 101 [9] 1344–50 (1997).
26
Conductivity in Barium Yttrium Stannate Ba2YSnO5.5,’’ Solid State Ionics, 98 [1– E. Ruiz-Trejo, M. S. Islam, and J. A. Kilner, ‘‘Atomistic Simulation of Defects
2] 1–6 (1997). and Ion Migration in LaYO3,’’ Solid State Ionics, 123 [1–4] 121–9 (1999).
11 27
R. Haugsrud and T. Norby, ‘‘High-Temperature Proton Conductivity in Ac- P. Saul, C. R. A. Catlow, and J. Kendrick, ‘‘Theoretical-Studies of Protons in
ceptor-Substituted Rare-Earth Ortho-Tantalates, LnTaO4,’’ J. Am. Ceram. Soc., Sodium-Hydroxide,’’ Philos. Mag. B, 51 [2] 107–17 (1985).
28
90 [4] 1116–21 (2007). J. M. Cole, M. R. Lees, J. A. K. Howared, R. J. Newport, G. A. Saunders, and
12
R. Haugsrud and T. Norby, ‘‘High-Temperature Proton Conductivity in Ac- E. Schonherr, ‘‘Crystal Structures and Magnetic Properties of Rare-Earth Ultra-
ceptor-Doped LaNbO4,’’ Solid State Ionics, 177 [13–14] 1129–35 (2006). phosphates, RP5O14 (R 5 La, Nd, Sm, Eu, Gd),’’ J. Solid State Chem., 150, 377–82
13
N. Kitamura, K. Amezawa, Y. Tomii, and N. Yamamoto, ‘‘Protonic Con- (2000).
29
duction in Sr-Doped (La1xSmx)PO4,’’ Solid State Ionics, 175 [1–4] 563–7 (2004). J. E. Bauerle, ‘‘Study of Solid Electrolyte Polarization by a Complex Admit-
14
H. Onoda, Y. Inagaki, A. Kuwabara, N. Kitamura, K. Amezawa, A. tance Method,’’ J. Phys. Chem. Solids, 30 [12] 2657–70 (1969).
30
Nakahira, and I. Tanaka, ‘‘Synthesis and Electrical Conductivity of Bulk Tetra- E. Lilley and J. E. Strutt, ‘‘Bulk and Grain-Boundary Ionic-Conductivity in
Valent Cerium Pyrophosphate,’’ J. Ceram. Process. Res., 11 [3] 344–7 (2010). Polycrystalline Beta’’-Alumina,’’ Phys. Status Solidi A, 54 [2] 639–50 (1979).
15 31
S. Jorgensen, J. A. Horst, O. Dyrlie, Y. Larring, H. Raeder, and T. Norby, F. Tyholdt, J. A. Horst, S. Jorgensen, T. Ostvold, and T. Norby, ‘‘Segregation
‘‘XPS Surface Analyses of LaPO4 Ceramics Prepared by Precipitation with or of Sr in Sr-doped LaPO4 Ceramics,’’ Surf. Interface Anal., 30 [1] 95–7 (2000).
without Excess of PO3 4 ,’’ Surf. Interface Anal., 34 [1] 306–10 (2002).
32
Y. X. Ni, J. M. Hughes, and A. N. Mariano, ‘‘Crystal-Chemistry of the Mo-
16
V. Nalini, R. Haugsrud, and T. Norby, ‘‘High-Temperature Proton Conduc- nazite and Xenotime Structures,’’ Am. Mineral., 80 [1–2] 21–6 (1995).
33
tivity and Defect Structure of TiP2O7,’’ Solid State Ionics, 181 [11–12] 510–6 (2010). S. J. Stokes and M. S. Islam, ‘‘Defect Chemistry and Proton-Dopant Asso-
17
S. W. Tao, ‘‘Conductivity of SnP2O7 and In-Doped SnP2O7 Prepared by an ciation in BaZrO3 and BaPrO3,’’ J. Mater. Chem., 20 [30] 6258–64 (2010).
34
Aqueous Solution Method,’’ Solid State Ionics, 180 [2–3] 148–53 (2009). R. Yu and L. C. De Jonghe, ‘‘Proton-Transfer Mechanism in LaPO4,’’
18
K. Amezawa, H. Maekawa, Y. Tomii, and N. Yamamoto, ‘‘Protonic Con- J. Phys. Chem. C, 111 [29] 11003–7 (2007).
35
duction and Defect Structures in Sr-Doped LaPO4,’’ Solid State Ionics, 145 [1–4] E. Matsushita and T. Sasaki, ‘‘Theoretical Approach for Protonic Conduction
233–40 (2001). in Perovskite-Type Oxides,’’ Solid State Ionics, 125 [1–4] 31–7 (1999).
19 36
K. Amezawa, Y. Kitajima, Y. Tomii, N. Yamamoto, M. Wideroe, and T. K. D. Kreuer, ‘‘Proton-Conducting Oxides,’’ Annu. Rev. Mater. Res., 33,
Norby, ‘‘Protonic Conduction in Acceptor-Doped LaP3O9,’’ Solid State Ionics, 333–59 (2003).
37
176 [39–40] 2867–70 (2005). T. Norby and Y. Larring, ‘‘Concentration and Transport of Protons in
20
I. Hammas, K. Horchani-Naifer, and M. Ferid, ‘‘Conduction Properties of Oxides,’’ Curr. Opin. Solid State Mater. Sci., 2 [5] 593–9 (1997).
38
Condensed Lanthanum Phosphates: La(PO3)3 and LaP5O14,’’ J. Rare Earths, 28 G. Z. Liu, H. X. Zhang, Z. E. Lin, S. T. Zheng, J. Zhang, J. T. Zhao, G. M.
[3] 321–8 (2010). Wang, and G. Y. Yang, ‘‘Germanates of 1D Chains, 2D Layers, and 3D Frame-
21
H. D. Park and E. R. Kreidler, ‘‘Phase Equilibria in the System La2O3–P2O5,’’ works Built from Ge–O Clusters by using Metal-Complex Templates: Host-Guest
J. Am. Ceram. Soc., 67 [1] 23–6 (1984). Symmetry and Chirality Transfer,’’ Chemistry, 2 [10] 1230–9 (2007).
22 39
M. Cherry, M. S. Islam, J. D. Gale, and C. R. A. Catlow, ‘‘Computational A. Jouini, M. Ferid, J. C. Gacon, L. Grosvalet, A. Thozet, and M. Trabelsi -
Studies of Protons in Perovskite-Structured Oxides,’’ J. Phys. Chem., 99 [40] Ayadi, ‘‘Crystal Structure and Optical Study of Praseodymium Polyphosphate
14614–8 (1995). Pr(PO3)3,’’ Mater. Res. Bull., 38 [11–12] 1613–22 (2003).
23 40
E. Ruiz-Trejo and R. A. De Souza, ‘‘Dopant Substitution and Oxygen Mi- N. N. Chudinova, A. V. Lavrov, and I. V. Tananaev, ‘‘Reaction of Bismuth
gration in the Complex Perovskite Oxide Ba3CaNb2O9: A Computational Study,’’ Oxide with Phosphoric Acid During Heating,’’ Izv. Akad. Nauk SSSR, Neorg.
J. Solid State Chem., 178 [6] 1959–67 (2005). Mater., 8 [11] 1971–6 (1972).
24 41
R. Glockner, M. S. Islam, and T. Norby, in Protons and Other Defects in H. Y. P. Hong, ‘‘Crystal Structures of Neodymium Metaphosphate (NdP3O9)
BaCeO3: A Computational Study,’’ Solid State Ionics, 122 [1–4] 145–56 (1999). and Ultraphosphate (NdP5O14),’’ Acta Crystallogr. B, B30, 468–74 (1974). &

Das könnte Ihnen auch gefallen