Você está na página 1de 12

View Online / Journal Homepage / Table of Contents for this issue

FEATURE ARTICLE www.rsc.org/materials | Journal of Materials Chemistry

Cation ordering in perovskites


Graham King and Patrick M. Woodward*
Received 21st December 2009, Accepted 4th March 2010
First published as an Advance Article on the web 15th April 2010
DOI: 10.1039/b926757c

Although both A- and B-site cations have the same simple cubic topology in the perovskite structure
they typically adopt different patterns of chemical order. As a general rule B-site cations order more
readily than A-site cations. When cation ordering does occur, rock salt ordering of B/B0 cations is
favored in A2BB0 X6 perovskites, whereas layered ordering of A/A0 cations is favored in AA0 B2X6 and
AA0 BB0 X6 perovskites. The unexpected tendency for A-site cations to order into layers stems from the
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C

bond strains that would result at the anion site if A and A0 cations of different size were to order with
a rock salt arrangement. The bonding instabilities that are created by layered ordering are generally
Downloaded by University of Illinois at Chicago on 15 May 2012

offset either by anion vacancies or second order Jahn–Teller distortions of a B-site cation. Novel types
of A-site cation ordering can be stabilized by a+a+a+ or a+a+c tilting of the octahedra.

1. Introduction Distortions from the aristotype structure can significantly


impact the physical properties. Octahedral tilting distortions,
The aristotype perovskite structure has ABX3 stoichiometry and which are present in 80–90% of all perovskites, occur when the
Pm 3m space group symmetry (Fig. 1). The smaller B-site cations A-site cation is too small for the cubo-octahedral cavities.1–5
are octahedrally coordinated by the X anions. The octahedra Octahedral tilting distortions alter the conduction bandwidth6
share corners to form a three dimensional network, while the and the strength of the magnetic superexchange interactions.7,8
larger A-site cations sit in the 12-coordinate cubo-octahedral As such they provide a mechanism for fine tuning the electrical,
cavities within this network. Perovskites display a wide range of magnetic and optical properties. Cooperative Jahn–Teller
properties including superconductivity (e.g. Ba1xKxBiO3), distortions can occur in perovskites containing cations with
colossal magnetoresistance (e.g. La1xCaxMnO3), itinerant appropriate electron counts (e.g. Mn3+, Cu2+).9–11 Cooperative
electron ferromagnetism (e.g. SrRuO3), multiferroic behavior Jahn–Teller distortions are intimately tied to orbital ordering and
(e.g. TbMnO3), ferroelectricity (e.g. BaTiO3), piezoelectricity hence the magnetic behavior of perovskites. Second order Jahn–
(e.g. PbZr1xTixO3), and ionic conductivity (e.g. La0.67xLi3x- Teller (SOJT) distortions occur in perovskites where high-valent
TiO3, BaCeO3x), among others. Perovskites arguably represent d0 cations on the B-site (e.g. Ti4+, Nb5+), and/or lone-pair cations
the most important family of complex oxides. on the A-site (e.g. Pb2+, Bi3+) shift out of the center of their
coordination polyhedra.12–14 SOJT distortions are central to the
realization of ferroelectricity and piezoelectricity in perovskites.
Department of Chemistry, Ohio State University, 100 West 18th Ave.,
Columbus, OH, 43210-1185, USA. E-mail: woodward@chemistry. In addition to the aforementioned structural flexibility,
ohio-state.edu perovskites possess considerable compositional flexibility.

Graham King received his BS Patrick Woodward received BS


degree in Chemistry from degrees in Chemistry and
SUNY Buffalo in 2005 and his General Engineering from Idaho
PhD in Chemistry from Ohio State University in 1991, an MS
State University in 2010, degree in Materials Science and
working in the Woodward a PhD in Chemistry from Ore-
research group. He is currently gon State University in 1996. He
a postdoc at Los Alamos spent two years as a postdoc at
National Laboratory. His Brookhaven National Lab
research interests involve explo- before accepting a faculty posi-
rations of magnetic, dielectric tion at Ohio State University in
and structural aspects of 1998, where he currently holds
complex perovskites, as well as the rank of Professor in the
Graham King cutting-edge powder diffraction Patrick M: Woodward Department of Chemistry. His
studies of extended and molec- research interests involve
ular solids. understanding the links between
structure, composition and bonding in oxide and pseudo-oxide
materials.

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 5785–5796 | 5785
View Online
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C
Downloaded by University of Illinois at Chicago on 15 May 2012

Fig. 1 The cubic ABX3 perovskite structure. Grey spheres represent the
A-site cations, blue spheres the B-site cations, and smaller red spheres are
the anions.

Chemical substitutions can be made onto all three sites of the


aristotype structure. The anion site can accommodate both
chemical substitutions (e.g. oxynitrides, oxyfluorides, etc.) and
high concentrations of vacancies. Cation substitutions, which
can take on either ordered or random arrangements, have been
used with much success to expand the perovskite family and
engineer new materials. Cation ordering can have profound
effects on the properties. Consider the following representative
examples:
 The dielectric properties of Pb-based double perovskites, Fig. 2 Cation ordering schemes in perovskites. From top to bottom 0D
such as Pb2ScTaO6, are sensitive to changes in the degree of B/B0 (rock salt), 1D (columnar ordering) and 2D (layered ordering) are shown
ordering.15 While highly ordered samples behave as normal for B-site ordering in A2BB0 X6 perovskites (left) and for A-site ordering in
AA0 B2X6 (right) perovskites.
ferroelectrics, relaxor behavior emerges as the B/B0 order
decreases.
 The magnetotransport properties of half-metallic ferrimag- ordered (e.g. Sr2YNbO6) arrangements can result, depending on
nets, such as Sr2FeMoO6, are highly sensitive to B/B0 disorder. differences in size and/or bonding preference of the B and B0
The spin polarization (half-metallicity) of this compound cations.22–24
degrades rapidly as ordering between the cations decreases.16 There have been several extensive reviews of B-site cation
 The rare combination of ferromagnetism and insulating order in perovskites, and the forces that drive B-site cation
behavior cannot be realized in ABO3 perovskites, but is attained ordering are generally understood. The interested reader is
in ordered perovskites such as La2MnMO6 and Bi2MnMO6 directed to earlier works by Anderson et al.,22 Davies et al.,25
(M ¼ Ni, Co).17,18 and Howard et al.21 for a closer look. In comparison relatively
 The Li3xLa2/3xTiO3 family of compounds has the highest little attention has been paid to the topic of A-site cation
known Li-ion conductivity of any material at room temperature. ordering, which is the primary focus of this review. We begin
The ionic conductivity is strongly influenced by the degree of with a look at the differences between ordering of A- and B-site
A-site cation ordering.19 cations.
While cation ordering can be realized with either A- or B-site
cations there are important differences in the way they order. In
2. Rock salt, columnar and layered ordering
the vast majority of A2BB0 X6 perovskites the B and B0 cations
take on an ordered pattern that is analogous to cation and anion There are three simple patterns of ordering that can be envi-
positions in the rock salt structure (Fig. 2). There are over 400 sioned for either the A- or B-site cations. The most symmetric is
documented examples of rock salt ordering in A2BB0 X6 perov- called rock salt ordering because the pattern of B and B0 (or A
skites.20,21 As a general rule when the oxidation states of B and B0 and A0 ) cations is equivalent to the anion and cation positions in
differ by less than two a disordered arrangement is observed the rock salt structure. In addition to the rock salt ordering,
(e.g. La2CrFeO6), whereas, a difference greater than two nearly cations can order into columns, or layers as shown in Fig. 2. The
always produces an ordered arrangement (e.g. Sr2NiWO6). rock salt arrangement can be thought of as the 0D case because
When the difference in oxidation states is exactly two, disordered each B0 X6 octahedron is isolated from all other B0 X6 octahedra
(e.g. Sr2FeRuO6), partially ordered (e.g. Sr2AlTaO6), or fully by B cations. Columnar ordering corresponds to 1D ordering

5786 | J. Mater. Chem., 2010, 20, 5785–5796 This journal is ª The Royal Society of Chemistry 2010
View Online

Table 1 A summary of the different types of cation ordering observed in perovskites. A 1 : 1 ratio of cations is assumed unless otherwise stated. The
examples do not necessarily adopt the aristotype space group

Ordering type Aristotype space group Examples Comment

B-site
Rock salt Fm
3m Ba2MgWO6 Common, stabilized by differences in the sizes and/or oxidation
states of the B/B0 cations
Columnar P4/mmm NdSrMn3+Mn4+O6 Rare, stabilized via charge ordering and cooperative FOJT
distortions of the B cations
Layered P4/mmm La2CuSnO6 Rare, stabilized by a FOJT distortion of the B cations and the
appropriate degree of octahedral tilting
A-site
Rock salt Fm
3m NaBaLiNiF6 Rare, stabilized by A/A0 charge difference, but destabilized by A/A0
size difference
Columnar P42/nmc CaFeTi2O6 Rare, stabilized by large A/A0 size mismatch and a+a+c octahedral
tilting
Layered P4/mmm YBaMn2O5 La1/3NbO3 NaLaMgWO6 Common, stabilized by (a) large A/A0 size mismatch coupled with
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C

anion vacancies, (b) A-site vacancies coupled with SOJT


distortions of the B0 cations, (c) B/B0 rock salt ordering coupled
Downloaded by University of Illinois at Chicago on 15 May 2012

with SOJT distortions of the B0 cations


Body centered Im
3 CaCu3Ti4O12 Uncommon, stabilized by a 1 : 3 ratio of larger A and smaller A0
cations and a+a+a+ octahedral tilting

because it allows for connectivity of the B0 X6 octahedra in one crystal chemistry considerations associated with 1 : 2 and 1 : 3
dimension. Finally, layered ordering corresponds to the 2D case ordering in perovskites.25
because it allows for connectivity of the B0 X6 octahedra in two Although dwarfed by examples of rock salt ordering there are
dimensions. Yet another description is to use directions to label a few examples of columnar and layered ordering among
the different patterns of chemical order. Specifically, we can label A2BB0 X6 perovskites. Examples of columnar ordering arise
the planes that contain like atoms (B or B0 ). In this scheme rock from charge ordering in mixed valent perovskites (e.g.
salt, columnar and layered ordering become (111), (110) and LaCaMn3+Mn4+O6, NdSrMn3+Mn4+O6).29–31 This pattern of
(001) ordering, respectively. Identical patterns can be envisaged ordering is stabilized by the strains associated with accommo-
for A-site cation ordering as shown on the right hand side of dating first order Jahn–Teller distortions (FOJT) of the B cations
Fig. 2. The characteristics of the different cation ordering (e.g. Mn3+). There are only a handful of examples of layered
schemes are given in Table 1. B-site cation ordering: Ln2CuSnO6 (Ln ¼ La, Pr, Nd, Sm),
If we assume that one of the two cations (B0 or A0 ) is more La2xSrxCuSnO6, and La2CuZrO6.32–34 In these compounds the
highly charged than the other (B or A), as is often the case, Jahn–Teller distortion of the Cu2+ ions plays an important role in
electrostatic considerations become important. From an elec- stabilizing the unusual layered ordering.
trostatic point of view rock salt ordering is the most favorable A-site cation ordering is much less common than B-site cation
because it maximizes the separation of the more highly charged ordering. While there is no shortage of AA0 B2X6 and AA0 BB0 X6
B0 cations. Each B0 cation has six B cations as its nearest perovskites, in the vast majority of cases the A and A0 ions do not
neighbors. This helps to explain the preference for rock salt order. This raises an interesting question, why are B-site cations
ordering among A2BB0 X6 perovskites. Columnar ordering, so much more prone to ordering than A-site cations? Differences
where each B0 cation has four B and two B0 nearest neighbors, is in oxidation state are one contributing factor. While the B/B0
the next best arrangement. Layered ordering presents the least oxidation state differential can be as large as seven (e.g.
favorable electrostatic situation, because each B0 cation has two Sr2LiReO6), outside of a few rare cases the A/A0 oxidation state
B and four B0 nearest neighbors. differential is limited to two or less. While this helps to explain
While rock salt ordering is dominant, other types of B-site why A-site cation ordering is less common, it does not explain
cation ordering are known. In A3BB0 2X9 (e.g. Ba3MgNb2O9)26 why A-site cation ordering, when it does occur, nearly always
and A4BB0 3X12 perovskites (e.g. Ba4LiSb3O12)27 the B-site prefers layered ordering over the expected rock salt arrangement.
cations order into layers that run perpendicular to (111), but Examples of long-range rock salt ordering of A-site cations are
because the B : B0 ratios are 1 : 2 and 1 : 3, respectively, the B and limited to two examples: partial ordering of Na+ and Ba2+ in
B0 cations adopt layer repeat sequences of B–B0 –B0 – in A3BB0 2X9 NaBaLiNiF635 and ordering of Na+ with a site that is 50%
perovskites and B–B0 –B0 –B0 – in A4BB0 3X12 perovskites. These occupied by Ba2+ in Na2BaFe4F12.36 In one additional example,
types of ordering are essentially the closest analogs to rock salt PbCaTi2O6, evidence for short range rock salt ordering of Pb2+
ordering that can be achieved given the stoichiometry and Ca2+ ions, has been reported.37
constraints. It should be noted that 1 : 2 ordering in A3BB0 2X9 To understand why B-site cations favor rock salt ordering
perovskites is favored when a cation capable of a SOJT distor- whereas A-site cations favor layered ordering we must consider
tion occupies the B0 site in order to relieve the bonding instability how cation ordering impacts the anion environment. In an ABX3
created by the creation of two chemically different anion sites: perovskite each anion has six cation neighbors: two B-site cations
one that is coordinated by two B0 ions and one that is coordi- and four A-site cations. When B-site cations rock salt order
nated by one B and one B0 ion.28 A recent review examines the the anion sites remain chemically and crystallographically

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 5785–5796 | 5787
View Online
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C
Downloaded by University of Illinois at Chicago on 15 May 2012

Fig. 3 The anion environment in a rock salt ordered A2BB0 X6 perov- Fig. 5 The anion environment in a rock salt ordered AA0 B2X6 perov-
skite. The color scheme for the ions is grey (A), blue (B), green (B0 ), and skite. The anion sits on an inversion center with A–X and A0 –X bonds of
red (X). equal length. The color scheme for the ions is grey (A), orange (A0 ), blue
(B), and red (X).

equivalent, each anion is now surrounded by one B and one B0


cation (Fig. 3). In this environment the anion can shift toward Pauling bond valence sum analysis suggests these anions will be
the smaller B0 cation and away from the larger B cation to overbonded. Another one-sixth of the anions are coordinated by
optimize its bonding. In contrast, layered ordering of B-site two lower valent B cations. In the absence of distortions these
cations creates three chemically distinct anion environments cations will be underbonded. The remaining anions are coordi-
(Fig. 4). One-sixth of the anions reside in layers where they are nated by one B and one B0 cation, as they are in a rock salt
coordinated by two of the higher valent B0 cations. Simple

Fig. 4 The three anion environments created by layered ordering of Fig. 6 The three anion environments created by layered ordering of
B-site cations. O(1) is coordinated by two B cations, O(2) is coordinated A-site cations. O(1) is coordinated by four A cations, O(2) is coordinated
by one B and one B0 cation, O(3) is coordinated by two B0 cations. The by two A and two A0 cations in a cis configuration, O(3) is coordinated by
color scheme is the same as in Fig. 3. four A0 cations. The color scheme is the same as in Fig. 5.

5788 | J. Mater. Chem., 2010, 20, 5785–5796 This journal is ª The Royal Society of Chemistry 2010
View Online

ordered A2BB0 X6 perovskite. This environment allows them to ambient pressure41 most require high pressure–high temperature
attain their ideal valence. Therefore, layered ordering leads to synthetic conditions.
a clear violation of Pauling’s fifth rule, the rule of parsimony, Unlike nearly all other perovskites, the AA0 3B4X12 composi-
which states that if possible all ions of the same type should have tions can accommodate transition metal ions on both the A- and
the same environment.38 B-sites. This leads to interactions between the A- and B-site
Similar arguments apply to ordering of A-site cations. Rock
salt ordering results in a single environment for all anions, one
where two A and two A0 cations coordinate the anion in a trans
configuration (Fig. 5). However, unlike rock salt ordering of
B-site cations the anion now sits on a site with inversion
symmetry and there is no way for it to shift in response to an A/A0
size differential. Simply put, the anion is forced to make A–X and
A0 –X bonds of equal length, even if A and A0 have different radii.
This simple topological argument is an important part of the
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C

reason why A-site cations rarely rock salt order.


Just as with B-site ordering, layered ordering of the A-site
Downloaded by University of Illinois at Chicago on 15 May 2012

cations creates three distinct anion environments (Fig. 6). Two-


thirds of the anions are surrounded by two A and two A0 cations
arranged in a cis fashion. The anion is now free to displace if
there is a significant difference in the size of the A and A0 cations.
Although this is a step in the right direction we still encounter the
same problems that we encountered with layered ordering in
A2BB0 X6 perovskites. One-sixth of the anions are surrounded by
four A cations while another one-sixth are surrounded by four
A0 -cations. For layered A-site ordering to occur, some mecha-
nism must be present to relieve this bonding instability. As we
will see in Sections 5 and 6 the preferred mechanism is SOJT
distortions of d0 cations on the B-site.

3. A-site ordering driven by octahedral tilting


As mentioned in the introduction, octahedral tilting distortions are
common in perovskites. In most cases octahedral tilting does not
remove the crystallographic equivalence of the A-site cations.
However, there are four Glazer tilt systems that are an exception to
this rule: a+a+a+, a+a+c, a0b+b+, and a0b+b.3,39 This situation
presents a novel strategy for stabilizing A-site ordering, namely to
use octahedral tilting to create holes of different size within the
corner sharing BX3 network. Of the four tilt systems that can give
rise to different A-site environments a+a+a+ is easily the most
important. No compounds with a0b+b+ or a0b+b tilting have been
reported, and only one example of a+a+c tilting is known, CaFe-
Ti2O6. In this compound the octahedral tilting drives columnar
ordering of Ca2+ and Fe2+.40 The Fe2+ ions have alternating tetra-
hedral and square planar geometry going down the column.
There are approximately 40 AA0 3B4X12 perovskites in which
ordering of the A and A0 cations is stabilized by a+a+a+ tilting. In
these perovskites A is a larger cation such as an alkali, alkaline
earth, or rare earth cation while A0 is a much smaller cation,
typically Cu2+ or Mn3+. The ordering that results from a+a+a+
tilting leaves the cubic symmetry intact, but doubles the length of
the cell edge. The structure, shown in Fig. 7, possesses Im 3 space
group symmetry. Three-fourths of the A-site cations have
a nearly square planar environment with four short A0 –X bonds.
This site is exclusively occupied by the smaller A0 -cations. These Fig. 7 Three depictions of the AA0 3B4X12 perovskite structure showing
A0 X4 squares are oriented such that they are orthogonal to the the octahedral environment of the B cation (top), the nearly square
nearest neighbor A0 X4 squares. Meanwhile, the A-cations retain planar environment of the A0 cation (middle), and the icosohedral envi-
a much larger 12-coordinate icosohedral environment. While ronment of the A cation. The color scheme for the ions is grey (A), orange
some of these compounds, such as CaCu3Ti4O12, can be made at (A0 ), blue (B), and red (X).

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 5785–5796 | 5789
View Online

A fascinating example of valence reorganization is observed in


LaCu3Fe4O12.53 Above 393 K this compound is best described as
La(Cu2+)3(Fe3.75+)4O12. Below 393 K electron density is trans-
ferred from copper to iron giving La(Cu3+)3(Fe3+)4O12, which
contains the rare diamagnetic square planar Cu3+ ion. The
valence reorganization triggers a change from paramagnetic
metallic behavior (T > 393 K) to antiferromagnetic insulating
behavior (T < 393 K).
The magnetic interactions in AA0 3B4O12 perovskites are
discussed at length in a recent review by Shimakawa.47 The
perpendicular arrangement of the A0 O4 square planes prevents
direct superexchange interactions between the A0 -cations,
although it does allow for direct ferromagnetic exchange. In
compounds with d10 ions on the B-sites, such as CaCu3Ge4O12
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C

and CaCu3Sn4O12, the direct exchange dominates and low


temperature ferromagnetic behavior is observed (TC ¼ 13 and
Downloaded by University of Illinois at Chicago on 15 May 2012

10 K, respectively). When d0 cations like Ti4+ occupy the B-site


they mediate superexchange interactions between Cu2+ ions. As
Fig. 8 The distortion of the anion environment created by large a+a+a+
a result CaCu3Ti4O12 orders antiferromagnetically at 25 K.47
octahedral tilting in an AA0 3B4O12 perovskite. The bond distances given
are for CaCu3Ti4O12. The color scheme is the same as used in Fig. 7. When the B-site cations possess localized, unpaired electrons
relatively strong antiferromagnetic A0 –O–B interactions lead to
ferrimagnetic ordering. In CaCu3Mn4O12 the Cu2+ and Mn4+
sublattices that are much stronger than normal. This is perhaps ions couple ferrimagnetically to give a saturation magnetization
most clearly illustrated by looking at the anion environment in of 9 mB per f.u. and TC ¼ 355 K.54 The high temperature ferri-
the AA0 3B4X12 perovskite structure. Instead of making two short magnetic coupling is retained in BiCu3Mn4O12, but the Bi3+ for
bonds to B-site cations, each anion now makes 3 short bonds, two Ca2+ substitution introduces carriers that lead to half metallic
with B cations and one to the A0 cation (Fig. 8). The fact that both behavior.55–57
the A0 and B cations make strong covalent bonds to the anions There are a few examples of AA0 3B2B0 2O12 perovskites where
allows for a variety of interesting phenomena involving either body centered A/A0 ordering and rock salt B/B0 ordering occur
strong B–X–A0 superexchange coupling, B 4 A0 charge transfer, simultaneously. Ordering of the B-site cations reduces the
or electron delocalization involving both cation sublattices. symmetry to Pn 3 (Fig. 9). The compositions CaCu3Ga2Sb2O12
The AA0 3B4O12 perovskites have a number of interesting and CaCu3Cr2Sb2O12 both have a complete rock salt ordering of
physical properties. Subramanian et al. have surveyed the the B-site cations while CaCu3Ga2Ta2O12 shows partial
dielectric properties of a large number of AA0 3B4O12 perov- ordering.51,58 The magnetic properties of these phases are also of
skites.42 Of particular interest is CaCu3Ti4O12 which has been interest. AA0 3Ga2B0 2O12 (B0 ¼ Sb, Ta) compounds, where Cu2+ is
extensively studied since it was discovered to have a nearly the only magnetic ion, remain paramagnetic down to 5 K.59 In
temperature independent colossal dielectric constant on the
order of 105.43,44 NaMn7O12, with manganese on both the A0 and
B-sites, has been found to pass through a series of transitions
involving charge, orbital, and magnetic ordering. On cooling it
charge orders into (NaMn33+)(Mn3+2Mn4+2)O12, where the Mn3+
and Mn4+ on the B-site are rock salt ordered.45 CaCu3Fe4O12
undergoes a charge disproportionation on cooling below 200 K
driving a transition from the high temperature paramagnetic
metallic CaCu3(Fe4+)4O12 state to the low temperature insulating
ferrimagnetic CaCu3(Fe3+)2(Fe5+)2O12 state, with rock salt
ordering of Fe3+ and Fe5+.46
In ACu3B4O12 perovskites oxygen 2p mediated hybridization
between the d-orbitals of Cu and the B-site ions is responsible for
a variety of intriguing electronic and magnetic properties. For
example ACu3B4O12 compounds with B ¼ Ru, Cr, or V are
metallic. The observation of Pauli paramagnetism rather than
Curie–Weiss paramagnetism implies delocalization of the
unpaired electrons on both Cu2+ and B cations.47–49 One of these
compounds, CaCu3Ru4O12 has been reported to show heavy-
fermion-like behavior with a Kondo temperature of 200 K.50
Itinerant electron behavior is also seen in the quintinary perov- Fig. 9 An AA0 3B2B0 2O12 perovskite with a+a+a+ tilting, ordering of the
skites CaCu3Ga2Ru2O12 and CaCu3Cr2Ru2O12, where the B-site A-site cations and rock salt ordering of the B-site cations. The color scheme
cations are disordered.51,52 for the ions is grey (A), orange (A0 ), blue (B), green (B0 ), and red (X).

5790 | J. Mater. Chem., 2010, 20, 5785–5796 This journal is ª The Royal Society of Chemistry 2010
View Online

CaCu3Cr2Sb2O12 Cr–O–Cu superexchange interactions domi-


nate and ferrimagnetic ordering results (TC ¼ 160 K).52

4. A-site ordering driven by anion vacancies


A-site cation ordering is often found in combination with anion
vacancy ordering. In these compounds the anion vacancies and
A/A0 size mismatch work cooperatively to stabilize layered
ordering. Many examples can be found in the RBaB2O5+x family
where the large size difference between Ba2+ and the rare-earth
cation, R3+, favors layered ordering. In these compounds the
oxygen vacancies tend to lie exclusively within the R layers. This
type of ordering doubles the c-axis lowering the symmetry to
tetragonal. For the fully oxygen deficient phases (x ¼ 0) the
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C

B-site coordination environment is reduced to square pyramidal


and the rare earth coordination number from 12 to 8 (Fig. 10).
Downloaded by University of Illinois at Chicago on 15 May 2012

Thus oxygen vacancy ordering provides a mechanism to reduce


the coordination number of the smaller R3+ ion without reducing
the coordination number of the larger Ba2+ ion. The fact that
A2B2O5 perovskites with a single A-site cation favor patterns of
oxygen vacancy ordering (e.g. Ca2Fe2O5) that differ from the
RBaB2O5+x perovskites illustrates the close coupling between
A-site ordering and oxygen vacancy ordering.60
Most RBaB2O5+x compounds can incorporate additional Fig. 11 The RBaB2O5.5 structure type showing the ordering of the
oxygen into the R-layers. The amount of additional oxygen that oxygen vacancies into columns. The color scheme is the same as for
can be incorporated is highly dependent upon the size of the rare Fig. 10.
earth cation. For example, within the RBaFe2O5 series those
compounds containing smaller R ions, such as Y3+, Ho3+, and
can be as high as 0.65 and 0.79, respectively.64 When R is even
Tb3+, can incorporate only very small amounts of additional
larger, such as with La3+, the oxygen content tends toward 1 and
oxygen.61–63 For larger R, such as Sm3+ and Nd3+, the value of x
the layered cation ordering is lost. In RBaB2O5.5 compounds,
such as NdBaFe2O5.5,64 SmBaFe2O5.5,64 and LaBaMn2O5.5,65 the
oxygen vacancies order so that half of the B-site cations are five
coordinate (square pyramids) and the other half are six coordi-
nate (heavily distorted octahedra), as shown in Fig. 11.
One of the distinctive features of the RBaB2O5 compounds is
the +2.5 oxidation state of the B-site cation, because half-integer
oxidation states are particularly favorable for charge ordering.
When B ¼ Mn, as in YBaMn2O5, charge and dx y orbital 2 2

ordering work cooperatively to stabilize rock salt ordering of


Mn2+/Mn3+ well above room temperature.66 When B ¼ Fe the
charge ordering temperatures are reduced. At high temperature
the iron sites are equivalent and mixed valent (all Fe2.5+). Upon
cooling antiferromagnetic ordering sets in below TN z 450 K,
followed by two charge ordering transitions. A premonitory
charge ordering, which can be detected by differential scanning
calorimetry and M€ ossbauer spectroscopy, is first seen (320 K < T
< 275 K). Further cooling leads to a complete long range
Fe2+/Fe3+ charge ordering (300 K < T < 215 K).61–63,67 The charge
ordering is not the rock salt pattern seen for YBaMn2O5 but
rather a columnar ordering. The columnar charge ordering is
thought to be favored over rock salt ordering as a result of orbital
ordering involving dxz orbitals. The RBaCo2O5 perovskites
adopt a similar pattern of charge ordering at somewhat lower
temperatures.68,69
In some cases it is possible to obtain AA0 B2X6 perovskites with
A/A0 layered ordering through a topotactic oxidation of an
Fig. 10 The RBaB2O5 perovskite structure type. The color scheme for oxygen deficient phase. When LaBaMn2O6 is prepared through
the ions is grey (Ba), orange (R), blue (B), and red (X). high temperature solid state routes the La3+ and Ba2+ ions do not

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 5785–5796 | 5791
View Online

order. However, low temperature oxidation of the oxygen defi-


cient phase, LaBaMn2O5, produces a metastable phase where the
layered cation ordering is retained.70 This strategy has been used
to prepare a number of other ordered RBaMn2O6 perovskites.71–74
Changing the A-site cation distribution from disordered to
ordered has a profound effect on the properties. The layered
compounds have considerably higher magnetic and charge
ordering temperatures.
Layered ordering of A-site cations is also observed in so-called
triple perovskites such as YBa2Fe3O8+x.75 In these compounds
the oxygen vacancies and Y3+ ions occupy every third layer along
the c-axis. The structures of many high TC superconductors, such
as YBa2Cu3O7x, can be derived from the triple perovskite
structure by removal of additional oxygen ions.76 An ordered
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C

arrangement of three A-site cations has been observed in the


perovskite LaYBa2Cu2Ti2O11.77 The B-site cations in this
Downloaded by University of Illinois at Chicago on 15 May 2012

compound are arranged into bilayers consisting of 2 layers of


TiO6 octahedra alternating with 2 layers of CuO5 square pyra-
mids. This creates three distinct A-sites. The A-site cations form
a repeating Y–Ba–La–Ba layer sequence, with Y sandwiched
between two Cu layers, La between two Ti layers, and Ba
between the Cu and Ti layers.

5. A-site ordering in perovskites with A-site


vacancies
Another place to look for layered ordering of the A-site cations is Fig. 12 The crystal structure of the Li3xR2/3xTiO3 perovskites. The
color scheme is blue (Ti4+), red (O2), and grey (R3+). The large three
in perovskites that have partially occupied A-sites. In compounds
colored spheres are randomly occupied by R3+, Li+, and vacancies. The
such as La1/3NbO3, La1/3TaO3, and Nd1/3NbO3 there is an
arrows show the displacements of the Ti4+ ions towards the anions in the
ordering of layers where the A-sites order into layers that are 2/3 layer containing the vacancies.
occupied and layers where the A-sites are completely vacant.78–80
In compounds such as La2/3TiO3 a related ordering occurs where
fully occupied layers alternate with layers where the A-sites are windows on the face of the unit cell (Fig. 13) where they can
1/3 occupied.81 The formulas of these compounds might be obtain a fourfold coordination.86 The same square windows are
more appropriately written as (R0.67M0.33)(M)B2O6 and the ones that Li+ ions must pass through to achieve ionic
(R)(R0.33M0.67)B2O6, where M is a vacancy. conductivity. To get high Li+ ion conductivity A-site vacancies
The presence of a highly charged d0 cation on the B-site is must be present and the octahedral tilting cannot be too large,
a common feature of compounds in this family. Such cations are
able to displace out of the centers of their octahedra by means of
a second order Jahn–Teller distortion. The displacements are
always towards the A-site cation layer that contains the higher
concentration of vacancies (Fig. 12). By moving toward the
underbonded anions, those contained in the layer that has the
larger concentration of A-site vacancies, the bonding instability
caused by the layered ordering is relieved.
There has been extensive work done on Li containing
compounds of the formula Li3xR2/3xTiO3 owing to their high
Li+ ionic conductivities.19,82,83 Such compounds have layers fully
occupied by R3+ which alternate with layers containing R3+, Li+,
and vacancies. These compounds often have complex micro-
structures which depend on both the composition and
the thermal treatment.84,85 The ionic conductivity of these
compounds depends on the concentration of vacancies in the
structure as well as the nature of the R3+ and B-site cations. The
Li+ ions are too small to fully occupy cubo-octahedral A-site
environment of an undistorted perovskite. In principle this
situation might be addressed by severe octahedral tilting, but the Fig. 13 A Li ion (orange sphere) residing in the four-coordinate square
presence of much larger R cations limits the octahedral tilting. window site of a Li3xR2/3xTiO3 perovskite. The blue spheres are tita-
The solution is for the Li+ ions to displace towards the square nium and red spheres are oxygen.

5792 | J. Mater. Chem., 2010, 20, 5785–5796 This journal is ª The Royal Society of Chemistry 2010
View Online

because octahedral tilting reduces the size of the square window


thereby limiting Li+ ion hops.
Recently, a fascinating type of phase segregation has been
observed for some Nd2/3xLi3xTiO3 compositions.87 TEM
images of these samples have revealed nanoscale chessboard like
contrast differences which have been attributed to segregation of
the sample into regions with different compositions. The
two regions have nominal compositions of Nd2/3TiO3 and
Nd1/2Li1/2TiO3. The driving force for the segregation has been
attributed to the Li+ ions preference for the square window site.
In the absence of vacancies a Li+ ion in the square window site
would lie much too close to a Nd3+ occupying a neighboring
A-site. This is thought to be the driving force for separation into
Nd2/3TiO3 and Nd1/2Li1/2TiO3 regions. By controlling the Li+ to
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C

Nd3+ ratio the relative sizes of these compositional regions can be


tuned. In addition there is a periodic twinning of the octahedral
Downloaded by University of Illinois at Chicago on 15 May 2012

tilt system.88
Several related phases have also been prepared which have
a phase segregation of both the A and B-site cations. If Ti4+ is
partially replaced by a trivalent cation (Al3+, Cr3+, or Mn3+) in the
proper proportion it will preferentially substitute into the
Nd2/3TiO3 region. These regions then become vacancy free
Fig. 14 An ideal AA0 BB0 O6 perovskite with rock salt ordering of B/B0
NdBO3 regions to maintain charge balance. The B3+ cations will and layered ordering of A/A0 The color scheme for the ions is grey (A),
prefer this region because it is possible to compensate for the orange (A0 ), blue (B), green (B0 ), red (X).
charge balance by removal of the vacancies and also because the
now vacancy free regions do not require a d0 cation to
compensate for layered ordering. In this way several compounds
comprised of Nd1/2Li1/2TiO3 and NdBO3 domains have been
obtained. These compounds display an intriguing variety of
nanoscale morphologies.89

6. Doubly ordered AA0 BB0 O6 perovskites


Both rock salt ordering of the B/B0 cations and layered ordering
of the A/A0 cations are found in the doubly ordered AA0 BB0 O6
perovskites (Fig. 14). They represent the largest class of perov-
skite that show simultaneous A- and B-site cation ordering. As
with the R1xBO3 perovskites discussed in Section 5 layered
ordering is stabilized by SOJT displacements of B-site cations.
The B0 cations displace along the c-axis towards the under-
bonded anion and away from the overbonded anion (Fig. 15).
Interestingly the layered ordering in these compounds appears to
be linked to the ordering of the B-site cations. When B-site order
is absent A-site order also disappears, even when SOJT cations
are present on the B-sites. For example, NaLaTi2O6 and
NaLaZr2O6 do not show any long range order of the Na+ and Fig. 15 The displacements of the SOJT cations (blue) in an AA0 BB0 O6
La3+ cations, whereas the A-site cations are highly ordered in perovskite. The color scheme is the same as for Fig. 14. In this figure the
NaLaMgWO6 and NaLaMnWO6.90,91 grey A-cations are of lower charge than the orange A0 -cations.
The first example of a doubly ordered AA0 BB0 O6 perovskite
was NaLaMgWO6, reported in 1984.92 Subsequent examples
include KLaMgWO6,93,94 NaLaCoWO6, NaLaNiWO6,95 There are a few compounds which do not follow the compo-
NaLaScNbO6,90 NaLiMgWO6,96 NaNdCoWO6,91 KLaMn- sitional trends that hold for the majority of doubly ordered
WO6,91 NaLnMnWO6 (Ln ¼ La, Ce, Pr, Nd, Sm, Gd, Dy, AA0 BB0 O6 perovskites. The compound NaLaMg2/3Nb4/3O6 is
Ho),91,97 and NaLnMgWO6 (Ln ¼ Ce, Pr, Nd, Sm, Eu, Gd, Tb, unique in that it has a 2 : 1 ordering of the B-site cations in
Dy, and Ho).91,97 To our knowledge this list represents all of the combination with layered ordering of the A-site cations.98 There
doubly order perovskites that have been discovered to date. The are a few compositions that show layered A-site cation ordering
common features among these compounds are monovalent and despite the absence of a SOJT cation on the B0 site. The Na+ and
trivalent cations as A and A0 together with the presence of La3+ cations in NaLaMgTeO6 have been reported to show
a highly charged d0 cation on the B0 site, but not the B site. layered ordering,93 although a subsequent study has shown the

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 5785–5796 | 5793
View Online

ordered domains to be very small.90 In the related compound a compositional modulation of the A-site cations. A model
KLaMgTeO6 the ordering of K+ and La3+ is nearly complete, of alternating regions with nominal stoichiometries of
presumably driven by the large size mismatch of the A-site NaLaMgWO6 and Na13xLa1+xMgWO6 has been proposed to
cations.96 The A-site cations in NdAgTi2O6 are partially ordered explain the compositional modulation.
despite the fact that d0 cations occupy both B and B0 sites.39 NPD patterns of NaLaMgWO6 exhibit superstructure reflec-
Large size and charge differences are also known to drive tions that cannot be explained by either simple octahedral tilting
a partial layered ordering of Na+ and Th4+ in Na2/3Th1/3TiO3. In or the compositional modulation found in the electron micro-
this compound Na0.88Th0.12 layers alternate with Na0.45Th0.55 scopy studies. The NPD patterns can only be modeled by
layers, and the Ti4+ cation is displaced towards the Na rich a periodic 1D twinning (Fig. 17) of the aac0 octahedral tilt
layer.99 system.101 The resulting unit cell dimensions are 12ap  2ap 
The combination of layered ordering of A-site cations, rock 2ap (46.84 Å  7.82 Å  7.90 Å with a ¼ 90 , b ¼ 90.1 , g ¼ 90 ),
salt ordering of B-site cations and octahedral tilting, results in showing a complexity that is unusual for perovskites and not
space group symmetries that differ from other perovskites.90 It’s evident from XRD studies. Since the compositional modulation
notable that the most common pattern of tilting in perovskites, and tilt twinning occur with the same periodicity it is thought
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C

aac+, leads to P21 space group symmetry in a doubly ordered that the two are intimately related.
AA0 BB0 O6 perovskites, because this is a polar space group that Complex superstructures have been observed in other
Downloaded by University of Illinois at Chicago on 15 May 2012

allows for ferroelectricity, piezoelectricity and second harmonic AA0 BB0 O6 perovskites. The HRTEM image of KLaMnWO6
generation. Octahedral tilting does not create polar space group shows a 2D pattern.102 Indexing of the ED pattern reveals a 10ap
symmetries in ABX3 and rock salt ordered A2BB0 X6 perovskites.  10ap supercell in the ab plane. A compositional modulation of
Neutron powder diffraction (NPD) has confirmed aac+ tilting the lanthanum content has also been confirmed for this
and P21 space group symmetry in NaLaMnWO6, compound. Modeling of the NPD pattern reveals two sets of
NaNdMnWO6, and NaTbMnWO6 (Fig. 16).91 perpendicular twin boundaries spaced 5ap apart, each also
The crystal structures of several doubly ordered AA0 BB0 O6 perpendicular to the layers of A-site cations. Satellite reflections
perovskites have proven to be more complex than originally have also been observed in the NPD patterns of several other
thought. The true nature of these structures only becomes compounds: NaLaScNbO6, KLaMgWO6, NaNdCoWO6, and
apparent when probed by NPD and electron microscopy. High NaNdMgWO6.91,96 These have not been fully analyzed, although
resolution transmission electron microscopy (HRTEM) studies it can be assumed that they also have a compositional modul-
of NaLaMgWO6 reveal a strong contrast of light and dark ations accompanied by tilt twin boundaries.
stripes that run perpendicular to the ordered layers of A-site The AA0 BB0 O6 perovskite family has been found to possess
cations.100 The corresponding electron diffraction (ED) pattern interesting magnetic properties. A number of AA0 MnWO6
showed supercell reflections corresponding to a periodicity of compounds order antiferromagnetically, with Neel temperatures
12ap. An analysis by electron energy loss spectroscopy (EELS) ranging from 6 to 15 K.91,97 This ordering appears to be driven by
showed a variation in the lanthanum content that repeats with long range Mn–O–W–O–Mn superexchange interactions. The
the same periodicity. The combined experimental probes reveal corresponding AA0 MgWO6 compounds remain paramagnetic
down to 2 K despite the presence of magnetic rare earth ions on
the A0 site.97 In AA0 MnWO6 compounds containing magnetic A0

Fig. 16 The crystal structure of NaTbMnWO6 showing the cation


ordering combined with aac+ octahedral tilting. Blue spheres are W6+, Fig. 17 One dimensional octahedral tilt twining (top) as in
green spheres are Mn2+, large grey spheres are Na+, large orange spheres NaLaMgWO6 and two dimensional octahedral tilt twinning (bottom) as
are Tb3+, and small red spheres are O2. in KLaMnWO6.

5794 | J. Mater. Chem., 2010, 20, 5785–5796 This journal is ª The Royal Society of Chemistry 2010
View Online

ions both the rare earth and the Mn2+ ions participate in and their impact on the magnetic and dielectric properties of
the magnetic ordering. Furthermore, several phases exhibit AA0 BB0 O6 perovskites.
multiple magnetic phase transitions as a result of the interplay
between magnetic ordering of the A0 and Mn2+ sublattices.103 Acknowledgements
Support for this research was provided by the National Science
Foundation (Award Number DMR-0907356). Partial funding
7. Conclusions for this research was provided by the Center for Emergent
A- and B-site ordering is both widespread and important for Materials at the Ohio State University, an NSF MRSEC (Award
understanding and controlling the properties of complex Number DMR-0820414).
perovskites. As a general rule B-site cation ordering is more
prevalent than A-site cation ordering, although as shown in this References
review, there are a variety of mechanisms for stabilizing A-site
1 A. M. Glazer, Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst.
cation ordering. Both A- and B-site cations can take on rock salt, Chem., 1972, 28, 3384–3392.
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C

columnar or layered patterns of ordering. Electrostatic consid- 2 P. M. Woodward, Acta Crystallogr., Sect. B: Struct. Sci., 1997, 53,
erations suggest that rock salt ordering should be the most 32–43.
Downloaded by University of Illinois at Chicago on 15 May 2012

favorable and layered ordering the least favorable. In practice 3 P. M. Woodward, Acta Crystallogr., Sect. B: Struct. Sci., 1997, 53,
44–66.
B-site cations prefer rock salt ordering whereas A-site cations 4 C. J. Howard and H. T. Stokes, Acta Crystallogr., Sect. B: Struct.
prefer layered ordering. Columnar ordering of B-site cations, Sci., 1998, 54, 782–789.
when observed, is generally associated with charge ordering 5 M. W. Lufaso and P. M. Woodward, Acta Crystallogr., Sect. B:
transitions and is likely stabilized by bond strains imposed by Struct. Sci., 2001, 57, 725–738.
6 H. W. Eng, P. W. Barnes, B. M. Auer and P. M. Woodward, J. Solid
simultaneous orbital ordering. State Chem., 2003, 175, 94–109.
The unexpected tendency for A-site cations to order into layers 7 J. S. Zhou and J. B. Goodenough, Phys. Rev. Lett., 2006, 96, 247202/
stems from the bond strains that are created at the anion site 247201–247202/247204.
8 B. H. Kim and B. I. Min, Phys. Rev. B: Condens. Matter Mater.
when A and A0 cations of different size order with a rock salt Phys., 2009, 80, 064416/064411–064416/064415.
arrangement. These strains destabilize rock salt order and allow 9 M. W. Lufaso and P. M. Woodward, Acta Crystallogr., Sect. B:
layered ordering to be realized. The bonding instabilities that are Struct. Sci., 2004, 60, 10–20.
created by layered ordering can be offset by (a) simultaneous 10 M. A. Carpenter and C. J. Howard, Acta Crystallogr., Sect. B:
Struct. Sci., 2009, 65, 134–146.
layered ordering of anion vacancies, as in the AA0 B2X6x 11 M. A. Carpenter and C. J. Howard, Acta Crystallogr., Sect. B:
perovskites, (b) layered ordering of A-site cation vacancies Struct. Sci., 2009, 65, 147–159.
coupled with second order Jahn–Teller (SOJT) distortions of d0 12 H. T. Stokes, E. H. Kisi, D. M. Hatch and C. J. Howard, Acta
Crystallogr., Sect. B: Struct. Sci., 2002, 58, 934–938.
cations on the B-site, as in the A1xBX3 perovskites, or (c) rock
13 K. M. Ok, P. S. Halasyamani, D. Casanova, M. Llunell, P. Alemany
salt ordering of B and B0 cations coupled with SOJT distortions and S. Alvarez, Chem. Mater., 2006, 18, 3176–3183.
of d0 cations of the B0 cations, as in the AA0 BB0 X6 perovskites. 14 M. Kunz and I. D. Brown, J. Solid State Chem., 1995, 115, 395–406.
The links between cation ordering, octahedral tilting, vacancy 15 B. P. Burton and E. Cockayne, Phys. Rev. B: Condens. Matter
Mater. Phys., 1999, 60, R12542–R12545.
ordering, first order Jahn–Teller (FOJT) distortions, and/or 16 D. D. Sarma, Curr. Opin. Solid State Mater. Sci., 2001, 5, 261–268.
SOJT distortions should not be underestimated. 17 K. Takata, M. Azuma, Y. Shimakawa and M. Takano, Funtai oyobi
A novel type of A-site cation ordering is stabilized in Funmatsu Yakin, 2005, 52, 913–917.
AA0 3B4X12 perovskites by a+a+a+ octahedral tilting. This pattern 18 S. F. Matar, M. A. Subramanian, A. Villesuzanne, V. Eyert and
M. H. Whangbo, J. Magn. Magn. Mater., 2007, 308, 116–119.
of octahedral tilting creates A-site environments of distinctly 19 S. Stramare, V. Thangadurai and W. Weppner, Chem. Mater., 2003,
different size within the corner sharing octahedral network: four 15, 3974–3990.
coordinate sites for the A0 cations and twelve coordinate sites for 20 M. W. Lufaso, P. W. Barnes and P. M. Woodward, Acta
Crystallogr., Sect. B: Struct. Sci., 2006, 62, 397–410.
the A cations. The A0 cations, usually either Cu2+ or Mn3+, make 21 C. J. Howard, B. J. Kennedy and P. M. Woodward, Acta
short covalent bonds with the anions that enable strong elec- Crystallogr., Sect. B: Struct. Sci., 2003, 59, 463–471.
tronic coupling between the A- and B-site sublattices. The 22 M. T. Anderson, K. B. Greenwood, G. A. Taylor and
resulting bond network allows for strong hybridization between K. R. Poeppelmeier, Prog. Solid State Chem., 1993, 22, 197–233.
23 P. W. Barnes, M. W. Lufaso and P. M. Woodward, Acta
the A0 and B cations. Metallic transport, including half-metal- Crystallogr., Sect. B: Struct. Sci., 2006, 62, 384–396.
licity and heavy-Fermion behavior, can result. As shown by 24 P. Woodward, R. D. Hoffmann and A. W. Sleight, J. Mater. Res.,
LaCu3Fe4O12, subtle changes in external conditions (tempera- 1994, 9, 2118–2127.
25 P. K. Davies, H. Wu, A. Y. Borisevich, I. E. Molodetsky and
ture, pressure) can trigger electron transfer between A0 and B L. Farber, Annu. Rev. Mater. Res., 2008, 38, 369–401.
cations. Finally, strong antiferromagnetic superexchange inter- 26 M. W. Lufaso, Chem. Mater., 2004, 16, 2148–2156.
actions between the A0 and B cations lead to high temperature 27 C. J. Howard and H. T. Stokes, Acta Crystallogr., Sect. B: Struct.
ferrimagnetic ordering in insulating AA0 3B4O12 perovskites. Sci., 2004, 60, 674–684.
28 J.-H. Park and P. M. Woodward, Int. J. Inorg. Mater., 2000, 2,
Although examples of doubly ordered AA0 BB0 O6 perovskites 153–166.
have been known since 1984 detailed structural studies have only 29 V. Caignaert, F. Millange, M. Hervieu, E. Suard and B. Raveau,
been carried out in recent years. These studies reveal a previously Solid State Commun., 1996, 99, 173–177.
30 P. M. Woodward, T. Vogt, D. E. Cox, A. Arulraj, C. N. R. Rao,
unknown complexity involving compositional modulation of the
P. Karen and A. K. Cheetham, Chem. Mater., 1998, 10, 3652–3665.
A-site cations and twinning of the octahedral tilt systems. There 31 P. G. Radaelli, D. E. Cox, M. Marezio and S. W. Cheong, Phys. Rev.
is still much work to be done to understand the structural details B: Condens. Matter, 1997, 55, 3015–3023.

This journal is ª The Royal Society of Chemistry 2010 J. Mater. Chem., 2010, 20, 5785–5796 | 5795
View Online

32 M. T. Anderson and K. R. Poeppelmeier, Chem. Mater., 1991, 3, 68 T. Vogt, P. M. Woodward, P. Karen, B. A. Hunter, P. Henning and
476–482. A. R. Moodenbaugh, Phys. Rev. Lett., 2000, 84, 2969–2972.
33 M. T. Anderson, K. R. Poeppelmeier, S. A. Gramsch and 69 J. C. Burley, J. F. Mitchell, S. Short, D. Miller and Y. Tang, J. Solid
J. K. Burdett, J. Solid State Chem., 1993, 102, 164–174. State Chem., 2003, 170, 339–350.
34 M. Azuma, S. Kaimori and M. Takano, Chem. Mater., 1998, 10, 70 F. Millange, V. Caignaert, B. Domenges, B. Raveau and E. Suard,
3124–3130. Chem. Mater., 1998, 10, 1974–1983.
35 M. Ducau, K. S. Suh, J. Senegas and J. Darriet, Mater. Res. Bull., 71 S. V. Trukhanov, I. O. Troyanchuk, M. Hervieu, H. Szymczak and
1992, 27, 1115–1123. K. Barner, Phys. Rev. B: Condens. Matter Mater. Phys., 2002, 66,
36 J. Darriet, S. G. Mayorga and A. Tressaud, Eur. J. Solid State Inorg. 184424/184421–184424/184410.
Chem., 1990, 27, 783–790. 72 T. Nakajima, H. Kageyama, H. Yoshizawa and Y. Ueda, J. Phys.
37 G. King, E. Goo, T. Yamamoto and K. Okazaki, J. Am. Ceram. Soc. Jpn., 2002, 71, 2843–2846.
Soc., 1988, 71, 454–460. 73 T. Nakajima, T. Tsuchiya, K. Daoudi, M. Ichihara, Y. Ueda and
38 L. Pauling, J. Am. Chem. Soc., 1929, 51, 1010–1026. T. Kumagai, Chem. Mater., 2007, 19, 5355–5362.
39 J. H. Park, P. M. Woodward and J. B. Parise, Chem. Mater., 1998, 74 Y. Ueda and T. Nakajima, Prog. Solid State Chem., 2007, 35,
10, 3092–3100. 397–406.
40 K. Leinenweber and J. Parise, J. Solid State Chem., 1995, 114, 75 P. Karen, A. Kjekshus, Q. Huang, V. L. Karen, J. W. Lynn,
277–281. N. Rosov, I. N. Sora and A. Santoro, J. Solid State Chem., 2003,
41 A. Deschanvres, B. Raveau and F. Tollemer, Bull. Soc. Chim. Fr., 174, 87–95.
Published on 15 April 2010 on http://pubs.rsc.org | doi:10.1039/B926757C

1967, 11, 4077–4078. 76 C. Chaillout, M. A. Alario-Franco, J. J. Capponi, J. Chenavas,


42 M. A. Subramanian, D. Li, N. Duan, B. A. Reisner and J. L. Hodeau and M. Marezio, Phys. Rev. B: Condens. Matter,
Downloaded by University of Illinois at Chicago on 15 May 2012

A. W. Sleight, J. Solid State Chem., 2000, 151, 323–325. 1987, 36, 7118–7120.
43 C. C. Homes, T. Vogt, S. M. Shapiro, S. Wakimoto and 77 K. B. Greenwood, G. M. Sarjeant, K. R. Poeppelmeier,
A. P. Ramirez, Science, 2001, 293, 673–676. P. A. Salvador, T. O. Mason, B. Dabrowski, K. Rogacki and
44 A. P. Ramirez, M. A. Subramanian, M. Gardel, G. Blumberg, D. Li, Z. Chen, Chem. Mater., 1995, 7, 1355–1360.
T. Vogt and S. M. Shapiro, Solid State Commun., 2000, 115, 78 B. J. Kennedy, C. J. Howard, Y. Kubota and K. Kato, J. Solid State
217–220. Chem., 2004, 177, 4552–4556.
45 A. Prodi, E. Gilioli, A. Gauzzi, F. Licci, M. Marezio, F. Bolzoni, 79 A. N. Salak, N. P. Vyshatko, D. D. Khalyavin, O. Prokhnenko and
Q. Huang, A. Santoro and J. W. Lynn, Nat. Mater., 2004, 3, V. M. Ferreira, Appl. Phys. Lett., 2008, 93, 162903.
48–52. 80 Z. Zhang, C. J. Howard, B. J. Kennedy, K. S. Knight and Q. Zhou,
46 I. Yamada, K. Takata, N. Hayashi, S. Shinohara, M. Azuma, J. Solid State Chem., 2007, 180, 1846–1851.
S. Mori, S. Muranaka, Y. Shimakawa and M. Takano, Angew. 81 C. J. Howard and Z. Zhang, J. Phys.: Condens. Matter, 2003, 15,
Chem., Int. Ed., 2008, 47, 7032–7035. 4543–4553.
47 Y. Shimakawa, Inorg. Chem., 2008, 47, 8562–8570. 82 Y. Inaguma, L. Chen, M. Itoh, T. Nakamura, T. Uchida, H. Ikuta
48 M. Labeau, B. Bochu, J. C. Joubert and J. Chenavas, J. Solid State and M. Wakihara, Solid State Commun., 1993, 86, 689–693.
Chem., 1980, 33, 257–261. 83 Y. Harada, T. Ishigaki, H. Kawai and J. Kuwano, Solid State Ionics,
49 M. A. Subramanian, W. J. Marshall, T. G. Calvarese and 1998, 108, 407–413.
A. W. Sleight, J. Phys. Chem. Solids, 2003, 64, 1569–1571. 84 S. Garcia-Martin and M. A. Alario-Franco, J. Solid State Chem.,
50 W. Kobayashi, I. Terasaki, J.-i. Takeya, I. Tsukada and Y. Ando, 1999, 148, 93–99.
J. Phys. Soc. Jpn., 2004, 73, 2373–2376. 85 S. Garcia-Martin, M. A. Alario-Franco, H. Ehrenberg,
51 S.-H. Byeon, S.-S. Lee, J. B. Parise, P. M. Woodward and N. H. Hur, J. Rodriguez-Carvajal and U. Amador, J. Am. Chem. Soc., 2004,
Chem. Mater., 2004, 16, 3697–3701. 126, 3587–3596.
52 S.-H. Byeon, S.-S. Lee, J. B. Parise, P. M. Woodward and N. H. Hur, 86 S. Garcia-Martin, A. Morata-Orrantia, M. A. Alario-Franco,
Chem. Mater., 2005, 17, 3552–3557. J. Rodriguez-Carvajal and U. Amador, Chem.–Eur. J., 2007, 13,
53 Y. W. Long, N. Hayashi, T. Saito, M. Azuma, S. Muranaka and 5607–5616, S5607/5601–S5607/5606.
Y. Shimakawa, Nature, 2009, 458, 60–63. 87 B. S. Guiton and P. K. Davies, Nat. Mater., 2007, 6, 586–591.
54 Z. Zeng, M. Greenblatt, M. A. Subramanian and M. Croft, Phys. 88 B. S. Guiton, H. Wu and P. K. Davies, Chem. Mater., 2008, 20,
Rev. Lett., 1999, 82, 3164–3167. 2860–2862.
55 J. A. Alonso, J. Sanchez-Benitez, A. De Andres, M. J. Martinez- 89 B. S. Guiton and P. K. Davies, J. Am. Chem. Soc., 2008, 130,
Lope, M. T. Casais and J. L. Martinez, Appl. Phys. Lett., 2003, 17168–17173.
83, 2623–2625. 90 M. C. Knapp and P. M. Woodward, J. Solid State Chem., 2006, 179,
56 X.-J. Liu, H.-P. Xiang, P. Cai, X.-F. Hao, Z.-J. Wu and J. Meng, 1076–1085.
J. Mater. Chem., 2006, 16, 4243–4248. 91 G. King, S. Thimmaiah, A. Dwivedi and P. M. Woodward, Chem.
57 K. Takata, I. Yamada, M. Azuma, M. Takano and Y. Shimakawa, Mater., 2007, 19, 6451–6458.
Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 024429/ 92 T. Sekiya, T. Yamamoto and Y. Torii, Bull. Chem. Soc. Jpn., 1984,
024421–024429/024424. 57, 1859–1862.
58 S.-H. Byeon, M. W. Lufaso, J. B. Parise, P. M. Woodward and 93 M. L. Lopez, M. L. Veiga and C. Pico, J. Mater. Chem., 1994, 4,
T. Hansen, Chem. Mater., 2003, 15, 3798–3804. 547–550.
59 S.-H. Byeon, S.-S. Lee, J. B. Parise and P. M. Woodward, Chem. 94 M. A. Arillo, J. Gomez, M. L. Lopez, C. Pico and M. L. Veiga, Solid
Mater., 2006, 18, 3873–3877. State Ionics, 1997, 95, 241–248.
60 M. T. Anderson, J. T. Vaughey and K. R. Poeppelmeier, Chem. 95 M. A. Arillo, J. Gomez, M. L. Lopez, C. Pico and M. L. Veiga,
Mater., 1993, 5, 151–165. J. Mater. Chem., 1997, 7, 801–806.
61 P. M. Woodward and P. Karen, Inorg. Chem., 2003, 42, 1121–1129. 96 M. C. Knapp, PhD dissertation, The Ohio State University,
62 P. M. Woodward, E. Suard and P. Karen, J. Am. Chem. Soc., 2003, Columbus, OH, USA, 2006.
125, 8889–8899. 97 G. King, L. M. Wayman and P. M. Woodward, J. Solid State Chem.,
63 P. Karen, P. M. Woodward, J. Linden, T. Vogt, A. Studer and 2009, 182, 1319–1325.
P. Fischer, Phys. Rev. B: Condens. Matter Mater. Phys., 2001, 64, 98 R. Tarvin and P. K. Davies, J. Am. Ceram. Soc., 2004, 87, 859–863.
214405/214401–214405/214414. 99 R. H. Mitchell and A. R. Chakhmouradian, J. Solid State Chem.,
64 P. Karen and P. M. Woodward, J. Mater. Chem., 1999, 9, 789–797. 1998, 138, 307–312.
65 V. Caignaert, F. Millange, B. Domenges, B. Raveau and E. Suard, 100 S. Garcia-Martin, E. Urones-Garrote, M. C. Knapp, G. King and
Chem. Mater., 1999, 11, 930–938. P. M. Woodward, J. Am. Chem. Soc., 2008, 130, 15028–15037.
66 J. P. Chapman, J. P. Attfield, M. Molgg, C. M. Friend and 101 G. King, S. Garcia-Martin and P. M. Woodward, Acta Crystallogr.,
T. P. Beales, Angew. Chem., Int. Ed. Engl., 1996, 35, 2482–2484. Sect. B: Struct. Sci., 2009, 65, 676–683.
67 P. Karen, P. M. Woodward, P. N. Santhosh, T. Vogt, 102 S. Garcia-Martin, G. King and P. M. Woodward, in press.
P. W. Stephens and S. Pagola, J. Solid State Chem., 2002, 167, 103 G. King, A. S. Wills and P. M. Woodward, Phys. Rev. B: Condens.
480–493. Matter Mater. Phys., 2009, 79, 224428/224421–224428/224429.

5796 | J. Mater. Chem., 2010, 20, 5785–5796 This journal is ª The Royal Society of Chemistry 2010

Você também pode gostar