Sie sind auf Seite 1von 19

J. Am. Ceram. Soc.

, 97 [7] 19932011 (2014)


DOI: 10.1111/jace.12982
2014 The American Ceramic Society

Journal

BiFeO3 Ceramics: Processing, Electrical, and Electromechanical Properties


Tadej Rojac,, Andreja Bencan, Barbara Malic,,* Goknur Tutuncu,,* Jacob L. Jones,,*
John E. Daniels,k and Dragan Damjanovic,*

Electronic Ceramics Department, Jozef Stefan Institute, Ljubljana 1000, Slovenia

Department of Materials Science and Engineering, University of Florida, Gainesville 32611, Florida

Department of Materials Science and Engineering, North Carolina State University, Raleigh 27695, North Carolina
k

School of Materials Science and Engineering, University of New South Wales, NSW 2052, Australia

Ceramics Laboratory, Swiss Federal Institute of Technology EPFL, Lausanne 1015, Switzerland

Dedicated to Prof. Dr. Marija Kosec, our Maricka, who left us after a long struggle in her tenacious spirit in December 2012.

Bismuth ferrite (BiFeO3), a perovskite material, rich in properties and with wide functionality, has had a marked impact on
the eld of multiferroics, as evidenced by the hundreds of articles published annually over the past 10 years. Studies from
the very early stages and particularly those on polycrystalline
BiFeO3 ceramics have been faced with diculties in the preparation of the perovskite free of secondary phases. In this
review, we begin by summarizing the major processing issues
and clarifying the thermodynamic and kinetic origins of the
formation and stabilization of the frequently observed secondary, nonperovskite phases, such as Bi25FeO39 and Bi2Fe4O9.
The second part then focuses on the electrical and electromechanical properties of BiFeO3, including the electrical conductivity, dielectric permittivity, high-eld polarization, and strain
response, as well as the weak-eld piezoelectric properties. We
attempt to establish a link between these properties and
address, in particular, the macroscopic response of the ceramics under an external eld in terms of the dynamic interaction
between the pinning centers (e.g., charged defects) and the
ferroelectric/ferroelastic domain walls.

I.

Introduction

ISMUTH ferrite (BiFeO3) has been subject of intensive


research with a large number of articles (over 3000)
published in the past 10 years. Possessing a rare combination
of both (anti)ferromagnetic and ferroelectric properties, coex-

D. J. Greencontributing editor

Manuscript No. 34489. Received January 30, 2014; approved April 4, 2014.
*Member, The American Ceramic Society.

Author to whom correspondence should be addressed. e-mail: tadej.rojac@ijs.si

isting at room temperature, and intercoupling between these


properties, the ferrite has had an important impact in the
eld of multiferroics.13 The great interest in BiFeO3 was initiated by the study of Wang et al. in 2003.1 The authors
reported a large remanent polarization in epitaxial BiFeO3
thin lms, that is, 5060 lC/cm2, which was an order of
magnitude higher than the best value reported for bulk BiFeO3 (6.1 lC/cm2) at that time and, thus naturally, though
wrongly, associated with the epitaxial strain. Soon after,
another group measured a comparable remanent polarization
in single crystals of BiFeO3, that is, ~100 lC/cm2 along the
pseudocubic (pc) [111]pc direction, revealing the true intrinsic
origin of the large spontaneous polarization of the ferrite.4
The very large spontaneous polarization coupled with the
possibility of manipulating the magnetic ordering with an
electric eld have triggered a series of studies on ferroelectric/ferroelastic switching in BiFeO3 thin lms.57 Parallel
studies on polycrystalline bulk BiFeO3 ceramics were, however, faced with more severe problems related to the processing and high electrical conductivity of this material,
providing additional diculties in properly understanding
the switching behavior of the ferrite in bulk form.2
Another functional property of BiFeO3 that has been widely
explored is piezoelectricity. There are several reasons that
place the ferrite among potentially valuable piezoelectric ceramic materials. Firstly, BiFeO3 is a lead-free compound and it
contains bismuth, an element whose electronic structure is similar to Pb,8 but is surprisingly harmless when compared to
most heavy metals, many of which are toxic.9 Secondly, owing
to its high Curie temperature (TC = 825C),3,10 it has been
considered for high-temperature piezoelectric applications.1115
Thirdly, its rhombohedral R3c structure in solid solutions
with other perovskites allows the creation of morphotropic
phase boundaries (MPBs) at which the piezoelectric coecients exhibit a maximum. Recently, all of these reasons initiated a large number of studies on chemically modied

Feature

1994

Vol. 97, No. 7

Journal of the American Ceramic SocietyRojac et al.

BiFeO3 ceramics. Among the most interesting are the BiFeO3PbTiO3 (BFPT)14,16 and BiFeO3BaTiO3 (BFBT)15,17,18
systems, which provide both enhanced piezoelectricity and a
high TC at the MPB, the latter exceeding that of Pb(Zr,Ti)O3
(PZT) (TC ~ 650C for BFPT, TC ~ 600C for BFBT and
TC ~ 350C for PZT at the MPB). In addition, a number of
other BiFeO3-based lead-free compositions are presently the
subject of intensive studies, including BiFeO3REFeO3
(RE = La, Nd, Sm, Gd, Dy),1921 BiFeO3AETiO3
(AE = Mg, Ca, Sr),2226 BiFeO3Bi0.5K0.5TiO327, and BiFeO3Bi(Zn0.5Ti0.5)O3.28 The piezoelectric properties of many
of these ceramic systems have not yet been characterized systematically.
The processing of single-phase BiFeO3 ceramics is dicult; however, signicant progress has been made recently,
particularly in relation to the identication of the origins of
the frequently formed nonperovskite, secondary phases. In
addition, the complex relationship between processing and
defects, on one hand, and the high- and weak-eld electrical
and electromechanical properties, on the other, has been
addressed to some extent. Up to now, a lot of these new
ndings, in particular those relating to processing and
domain-switching behavior, have not been considered suciently or are even ignored in the literature. Along with the
aim of presenting a detailed overview of the past and recent
results on BiFeO3 ceramics, this absence or poor coverage
of some important topics was one of the motivations that
led us to prepare a comprehensive article, which also
includes new data.
The review comprises two topics on BiFeO3 that are the
most controversial and have encountered signicant research
barriers, that is, the processing and the electrical/electromechanical properties. The study does not cover magnetic properties and structural aspects. Whereas the crystal structure of
BiFeO3 has been reviewed in a study by Catalan and Scott,3
and referred to in earlier publications,10,29,30 the data on the
microstructure and the domain structure are only reported
occasionally, without general relations to, for example, the
piezoelectric properties of the undoped BiFeO3 bulk materials and ceramics in particular.
The section on processing covers the major issues
encountered in the synthesis of BiFeO3: (i) the thermodynamic instability and ternary phase diagrams with impurity
oxides; (ii) the reaction kinetics of the Bi2O3Fe2O3 system;
and (iii) the sublimation/evaporation of bismuth oxide during high-temperature treatments. By critically analyzing
these interrelated issues we reveal the multiple origins of the
often-formed secondary phases and their persistence in sintered BiFeO3 ceramics. The processing section ends by illustrating the major problems accompanying the conventional
solid-state processing of BiFeO3, along with general guidelines to be considered when undertaking the synthesis of the
ferrite.
BiFeO3 is known as a perovskite with a high electrical
conductivity, which is probably the major concern for the
applications of the ferrite and its chemical modications. The
rst part of the section on properties is devoted to this issue
where we discuss the impact of the electrical conductivity on
the low-frequency (mHz-to-MHz region) dielectric permittivity of BiFeO3 and address some of the inconsistencies
found in the literature. Based on the available literature data
and our own studies, the character of dielectric relaxation,
the type of conductivity, as well as the nature of point and
electronic defects are discussed.
Considering the unfortunate combination of the high electrical conductivity and the high coercive eld of BiFeO3, it is
not surprising that diculties in studying domain switching
and piezoelectricity persist. We address the switching behavior of the ferrite through an analysis of the polarization(PE) and strain-electric-eld (SE) hysteresis loops, supported
by synchrotron X-ray diraction analysis. In addition, we
review and discuss the direct longitudinal piezoelectric

response of BiFeO3 ceramics through its dependence on the


external driving variables, that is, the amplitude and the frequency of the stress.

II.

Processing Issues

(1) Thermodynamic Stability, Impurities, and Chemical


Compatibility
The Bi2O3Fe2O3 phase diagram, recently revised by Maitre
et al.31 and Palai et al.,32 indicates three equilibrium phases
(in order from Fe- to Bi-rich side): the orthorhombic
Bi2Fe4O9 [space group (s.g.) Pbam, ICSD #20067], the rhombohedral perovskite BiFeO3 (s.g. R3c, ICSD #15299), which
decomposes peritectically to Bi2Fe4O9 and a liquid phase at
~935C, and cubic Bi25FeO39 (s.g. I23, ICSD #62719), the
latter exhibiting a peritectic decomposition to Bi2O3 and a
liquid phase at ~790C.
The instability of BiFeO3 at elevated temperatures, which
was only indicated in one of the early Bi2O3Fe2O3 phase diagrams,33,34 was recognized as early as the 1960s when attempts
to determine its Curie temperature were compromised by the
reported slow decomposition of the ferrite at T > 700C, that
is, at temperatures well below its peritectic decomposition.3436
Dierent decomposition products were reported, including
Bi2O3, Fe2O3, and Bi2Fe4O9.35,37 It was suggested that the
decomposition reaction is irreversible,36 taking place at
temperatures before the onset of the Bi2O3 sublimation.35
Experimental evidence for the high-temperature instability
of BiFeO3 was later reported by several other authors. Owing
to the diculty in preparing single-phase perovskite by a thermal treatment of a Bi2O3Fe2O3 powder mixture, the authors
generally refer to the metastability of BiFeO3.32,3840 Morozov et al.38 showed that apparently pure BiFeO3 can be prepared by reacting a Bi2O3Fe2O3 mixture at 850C for short
time, that is, 510 min; however, upon further isothermal
annealing at the same temperature for 2 h, the BiFeO3 started
to decompose into Bi25FeO39 and Bi2Fe4O9, revealing more
directly the thermodynamic instability of the BiFeO3. The
same kind of decomposition was later conrmed by annealing
a solgel-derived BiFeO3 at 600C for 65 h.40 The decomposition was also observed in BiFeO3 single crystals, conrming
the intrinsic nature of the instability.32
The thermodynamic instability of BiFeO3 has been claried only recently as part of the experimental and theoretical
studies by Selbach et al.41 Using the thermodynamic data
published by Phapale et al.,42 they calculated the Gibbs free
energy for the equilibrium reaction:
1=49Bi25 FeO39 12=49Bi2 Fe4 O9

! BiFeO3

(1)

The results of the calculation for the temperature range


5001200 K are shown in Fig. 1(b). The calculations showed
that the Bi25FeO39 and Bi2Fe4O9 phases are, though only
weakly, more thermodynamically stable than the BiFeO3 in
the temperature range 447C767C, that is, the DrGm is
slightly positive in this temperature interval, implying that
the equilibrium reaction (1) is spontaneously shifted to the
left, that is, toward the Bi25FeO39 (Bi-rich) and Bi2Fe4O9
(Bi-poor) phases. This is in agreement with the experimental
results: a presynthesized BiFeO3 with a small initial amount
of the two Bi-rich and Bi-poor secondary phases [Fig. 1(a),
25C] decomposed into these phases within the temperature
instability range at 700C [Fig. 1(a), 700C], whereas the BiFeO3 phase reappeared once the sample was annealed above
the instability range [Fig. 1(a), 775C and 850C]. The same
behavior was observed when annealing a Bi2O3Fe2O3 mixture, instead of BiFeO3, in agreement with earlier studies.38,40
It should be noted that the decomposition reaction
described so far should not be confused with that related to
the formation of BiFeO3 from constituent oxides:

July 2014

BiFeO3 Ceramics: Processing and Properties

1995

(b)

(a)

Fig. 1. Thermodynamic instability of BiFeO3. (a) X-ray diraction (XRD) patterns of BiFeO3 at room temperature (25C) and after isothermal
annealing at 700C, 775C, and 850C, showing rstly the decomposition of BiFeO3 into Bi25FeO39 and Bi2Fe4O9 at 700C, followed by the
disappearance of these secondary phases (i.e., the reappearance of BiFeO3) at T 775C. (b) Calculated temperature dependence of the Gibbs
free energy (DrGm) of the equilibrium reaction between BiFeO3 and the Bi- and Fe-rich phases [data are plotted for DrHm = 1.98 kJ/mol and
DrSm = 3.62 J/molK; see Ref. (41)]. Note the slightly positive DrGm in the region 447C< T < 776C, implying that the Bi25FeO39 and
Bi2Fe4O9 are, though weakly, more thermodynamically stable than BiFeO3 in this temperature interval, in agreement with the experimental data
shown in panel (a). Reprinted with permission from Ref. [41] Copyright 2009 American Chemical Society.

Bi2 O3 Fe2 O3 ! 2BiFeO3

(2)

the calculated enthalpy of which is approximately 70 kJ/


mol at 25C.41 Thus, while reaction (2) will take place at elevated temperatures, resulting in the formation of BiFeO3, the
created BiFeO3 will tend to decompose according to reaction
(1) if the sample is annealed between  447C and  767C
[Fig. 1(b)].
In principle, the equilibrium reaction (1) implies that the
coexistence of the three phases would only be possible at
447C and 767C, where DrGm = 0 [Fig. 1(b)]. The oftenreported coexistence of these three phases in a broader temperature range [Fig. 1(a)] may be explained by the small driving force (DrGm) of the forward and backward reactions,
that is, DrGm is slightly positive within the instability region
[447C767C, Fig. 1(b)] and slightly negative in the stable
region at temperatures close to 767C. This small driving
force may slow down the reactions from BiFeO3 to the two
secondary phases and backwards, resulting in an apparent
three-phase coexistence.41
In addition to the above-mentioned reasons, Valant
et al.43 showed that the coexistence of the perovskite and
the two Bi-rich and Bi-poor phases may be explained thermodynamically by introducing a third component in the
Bi2O3Fe2O3 system. In terms of the Gibbs phase rule, this
(a)

eectively increases the number of independent variables.


They showed that the third component, which could be an
impurity added in a very small quantity (<1 wt%), could
result in a large amount (several tens of vol%) of thermodynamically stabilized Bi25FeO39-like and Bi2Fe4O9-like phases
in cases when the impurity is soluble in either Bi25FeO39
and/or Bi2Fe4O9.
For the sake of clarity, we explain in Fig. 2 the example
of SiO2 as the third component (impurity).43 In this case it
was assumed and subsequently conrmed by X-ray diraction (XRD) analysis that the SiO2 will react to form
Bi12SiO20, that is, a phase belonging to the family of socalled sillenite phases, which are isostructural with c-Bi2O3.44
The next reasonable assumption was that the Bi12SiO20 will
exhibit complete or limited solid solubility with the isostructural Bi25FeO39 phase [Fig. 2(b)]. According to the proposed
phase diagram [Fig. 2(b)], the incorporation of SiO2 into the
Bi2O3Fe2O3 system would shift the composition into a
three-phase region and, thus, three phases will result in equilibrium conditions: BiFeO3, Bi2Fe4O9, and the sillenite phase
[see the three-phase eld (triangle) marked with an arrow in
a small portion of the phase diagram in Fig. 2(b)]. Owing to
the exceptionally elongated and narrow BiFeO3Bi2Fe4O9
sillenite ternary phase eld, which is determined by the specic positions of the phases in the diagram, it was shown by
(b)

Fig. 2. Inuence of impurities on the phase composition of a reacted Bi2O3Fe2O3 mixture (example of SiO2 as impurity). (a) X-ray diraction
(XRD) patterns of Bi2O3Fe2O3 mixture, annealed at 800C for 5 h, without (0.0%) and with additions of 0.1wt% and 0.5wt% SiO2. (b)
Proposed phase relations in the ternary Bi2O3Fe2O3AOx system, where AOx is SiO2. According to the proposed phase diagram (panel b), a
small amount of SiO2 added to the Bi2O3Fe2O3 mixture should result in a large amount of secondary phases at equilibrium, as conrmed
experimentally (see panel a). Reprinted with permission from Ref. [43] Copyright 2007 American Chemical Society.

1996

Vol. 97, No. 7

Journal of the American Ceramic SocietyRojac et al.

calculations that a small amount of introduced SiO2 will generate a large amount of the Bi-rich and Bi-poor secondary
phases upon reacting the mixture. The calculations were conrmed experimentally by deliberately adding 0.1 and 0.5 wt
% of SiO2 to a stoichiometric Bi2O3Fe2O3 mixture; upon
annealing, the additions indeed resulted in large amounts of
the secondary phases, which were easily detected by XRD
[Fig. 2(a)].
In addition to SiO2, ternary phase diagrams were also proposed for other impurities: (i) Al2O3, which, according to
energy-dispersive X-ray spectroscopy analysis (EDXS), incorporates into the Bi2Fe4O9 phase; and (ii) TiO2, which, in
addition to incorporating into the Bi2Fe4O9, partially substitutes for the Fe3+ in BiFeO3.43 Both oxides, added in small
quantities (up to 0.5 wt%), led to the compositional degradation of the BiFeO3 upon annealing, just like in the case of
SiO2 [Fig. 2(a)]. Thus, the strong inuence of impurities on
the equilibrium phase composition has its origin in the
specic phase relations intrinsic to the Bi2O3Fe2O3AOx
ternary systems, where AOx represents an impurity oxide
that is more soluble in either Bi25FeO39 and/or Bi2Fe4O9
than in BiFeO3.
Though not directly discussed, the eect of impurities on
BiFeO3 should be considered with care, taking into account
the long list of oxides, including SiO2, ZnO, GeO2, PbO,
Ga2O3, ZrO2, Nb2O5, and MnO, that can form the sillenite
phase by reacting with Bi2O3,4447 possibly resulting in the
compositional degradation of the ferrite (analogous to the
case of SiO2, Fig. 2). Therefore, monitoring the type and
amount of impurities present in the starting Bi2O3 and Fe2O3
powders is extremely important; as a matter of fact, phasepure BiFeO3 can be successfully synthesized using starting
oxides with ultrahigh purity, that is, >99.999%.43 Another
source of contamination during the processing of BiFeO3,
which is rarely considered in the literature, is the milling procedure. For example, prolonged milling (>10 h) using conventional yttria-stabilized zirconia (YSZ) milling media and
vials may cause substantial wearing of the milling balls and
contamination of the powder with ZrO2;48 the ZrO2 may
then react upon annealing the contaminated powder to form
the sillenite phase.46
Even though all the impurity oxides may not necessarily
have such a strong eect as SiO2 or Al2O3, the sensitivity of
BiFeO3 to impurities should also be considered in view of
the doping. Namely, several authors report on the persistence
of the secondary phases in the ceramics even at low doping
levels (1 mol%).4951 For example, the study on TiO2,43 one
of the often-reported dopants for BiFeO3,5256 suggests that
once the maximum solubility of TiO2 in BiFeO3 is exceeded,
which may happen even locally in the ceramics due to an
inhomogeneous distribution of the dopant, the composition
of the system will fall within the narrow three-phase eld
[see, as an example, Fig. 2(b)], resulting in the appearance of
secondary phases in large quantities. Therefore, when choosing a dopant for BiFeO3, one has to be careful particularly
when the dopant tends to be incorporated into the Bi25FeO39
and/or Bi2Fe4O9 since this may stabilize these secondary
phases. Extended discussions on this topic were presented by
Bernardo et al.55 for the case of the B-site W6+, Nb5+, and
Ti4+ doping cations. Other problems related to doping BiFeO3 involve the segregation of the dopant at the grain boundaries and the suppression of grain growth by doping.56
Another problem that is the direct result of the eect of
impurities on the equilibrium phase composition is the chemical incompatibility between the BiFeO3 and substrates. SiO2
and Al2O3 are the main constituents of refractory materials
that are commonly in contact with the ferrite during sintering. This may lead to a compositional degradation of the
BiFeO3 at such contacts. In fact, it was proposed that the
often-observed decomposition of BiFeO3 in its paraelectric
phase is due to the chemical incompatibility between the ferrite and the materials in contact with the sample during

high-temperature treatments,57 rather than being intrinsic to


the BiFeO3.58
For illustrative purposes we show an example of the reaction between BiFeO3 and an Al2O3 substrate in Fig. 3. After
sintering a BiFeO3 pellet on Al2O3 at 900C an interface
reaction was observed, that is, a marked region appeared on
the crucible, below the original position of the pellet
[Fig. 3(b)]. According to XRD analysis, this interface reaction coincides with the formation of large quantities of
Bi25FeO39 and Bi2Fe4O9 secondary phases [Fig. 3(a), see
900C and compare with 760C, where the interface reaction
was not observed). The decomposition of the BiFeO3 is thus
activated at the contact with the Al2O3 and cannot be
explained by considering the intrinsic thermodynamic instability of BiFeO3 as 900C is well above the instability range
[see Fig. 1(b)]. The sublimation/evaporation of Bi2O3 itself is
not likely either since, in this case, one would expect the formation of only the Bi-poor phase, that is, Bi2Fe4O9, or eventually Fe2O3 (see next section and the Bi2O3Fe2O3 phase
diagram31,32). Similar reactions were reported between BiFeO3 and a quartz substrate,59 in agreement with the degrading
eect of SiO2 (Fig. 2).

(2) Reaction Kinetics, Sillenite Phase, and


Sublimation/Evaporation
In addition to the thermodynamics discussed in the previous
section, signicant progress has also been made in the reaction
kinetics of the Bi2O3Fe2O3 system. Studies on Bi2O3Fe2O3
diusion couples, performed at 650C, revealed that bismuth
diuses several micrometers inside the Fe2O3, while there was
little backward diusion of iron into the Bi2O3.60 Assuming
coupled diusion of the Bi3+ and O2 ions, which preserves
the electroneutrality, and considering that the oxygen diusion in complex oxides is generally faster than the cation diusion,61,62 we can reasonably assume that the formation of
BiFeO3 is probably controlled by the diusion of Bi3+ toward
the Fe2O3. The low diusion rate of iron is consistent with the
available diusion data, calculated for 700C: the tracer diusion coecient of Fe3+ in Fe2O3 (D700C ~ 2.8 9 10 25 m2/s)
is ve orders of magnitude lower than that of Bi3+ in Bi2O3
(D700C ~ 6.8 9 10 20 m2/s).63,64
The faster bismuth diusion explains the commonly
observed microstructural features in BiFeO3 ceramics, that is,
an Fe-rich region inside a BiFeO3 grain with the Bi-rich sillenite phase close to the BiFeO3 grain boundary [Fig 4(a)].60

(a)

(b)
(b1)

(b2)

Fig. 3. Chemical instability of BiFeO3 in contact with Al2O3


substrate. (a) X-ray diraction (XRD) patterns of Bi2O3Fe2O3
mixture annealed at 760C for 6 h and 900C for 10 h and (b)
corresponding photographs of the pellet annealed at 900C,
illustrating the interface reaction between BiFeO3 and the Al2O3
substrate. The pellet was placed on a curved Al2O3 substrate to
minimize the BiFeO3/Al2O3 contact. These contacts, at which the
interface reaction was initiated, are marked with arrows in the photo
(b2).

July 2014

BiFeO3 Ceramics: Processing and Properties


(a)

1997

(b)

Fig. 4. Reaction kinetics in the Bi2O3Fe2O3 system. (a) Scanning electron microscopy (SEM) image of a characteristic distribution of phases
with dierent Bi/Fe molar ratios observed in the sintered ceramics, that is, a BiFeO3 grain (light-gray phase) with the Bi-rich sillenite phase at
the BiFeO3 grain boundary (bright phase) and an Fe-rich phase (Bi2Fe4O9 or Fe2O3; dark-gray phase) in the interior of the BiFeO3 grain. The
sample was prepared by a conventional mixed-oxide route at 750C for 2 h. (b) Proposed reaction pathway mechanism for the solid-state
synthesis of BiFeO3 from Bi2O3 and Fe2O3, based upon diusion-couple studies and experimental observations (see panel a). Reprinted with
permission from Ref. [60] Copyright 2011 Elsevier Ltd.

This picture may be understood as a kinetically stabilized


phase coexistence, in which incomplete diusion of the bismuth through the BiFeO3 grain resulted in an unreacted Ferich core. Thus, the initial particle size of the Fe2O3 powder,
which determines the diusion distance, is an important
parameter that should be controlled to avoid unreacted
phases in this system. For comparison, a similar diusion
mechanism governs the formation of BaTiO3 from a mixture
of BaCO3 and TiO2, whereby the growth of BaTiO3 is controlled by the diusion of barium ions through the perovskite
layer.65,66
Based on the observed formation of the sillenite, BiFeO3
and Bi2Fe4O9 phases at the interface of the annealed Bi2O3
Fe2O3 diusion couple, Bernardo et al.60 proposed a reaction
mechanism, which is sketched in Fig. 4(b). At the beginning
of the reaction, the sillenite is the rst phase formed at the
Bi2O3Fe2O3 contact surface, which is consistent with earlier
studies,38,67 and the BiFeO3 phase develops toward the interior of the Fe2O3 particles. In parallel with the BiFeO3, the
Bi2Fe4O9 phase is also formed, resulting in a diusion/reaction phase sequence of the type: Bi25FeO39/BiFeO3/Bi2Fe4O9/
Fe2O3 (Fig. 4b). As discussed by the authors, it is often the
case that large, regularly shaped Bi2Fe4O9 grains persist in
the nal ceramics [see also schematics in Fig. 4(b)]. This was
explained by the stable growth of these crystals once formed;
however, we note that a similar eect could also occur due
to the presence of impurities (see previous section). In addition, one has to consider both the thermodynamics and the
kinetics, for example, the Bi2Fe4O9 phase that is kinetically
stabilized can grow if the reaction is driven at a temperature
within the intrinsic thermodynamic instability region of BiFeO3 (Fig. 1).
One of the characteristics of the synthesis of BiFeO3 is the
easy formation of the sillenite Bi25FeO39 phase or its isostructural phases stabilized by the impurity oxides (see the
example of Bi12SiO20 in Fig. 2). Several authors showed independently that the Bi25FeO39 is the rst reaction product
formed upon reacting the Bi2OFe2O3 1:1 mixture at temperatures as low as 400C500C.38,60,67 In terms of processing
the BiFeO3, this represents a serious problem that is rarely
considered in the literature. For example, if this phase does
not react completely with the Fe-rich counterparts (Fe2O3 or
Bi2Fe4O9), which may be due to the thermodynamic or
kinetic reasons discussed earlier, it can melt at temperatures
even lower than that of pure Bi2O3 (Tm(Bi2O3) ~ 830C and
Tp(Bi25FeO39) ~ 790C31,32, where Tm and Tp refer to melting and peritectic temperature, respectively). The incongruent
melting through peritectic decomposition of the sillenite
phase may then lead to (i) uncontrolled Bi2O3 losses associated with the increased vapor pressure of the bismuth oxide
above the melt compared to the expected lower vapor pressure above solid BiFeO3,43,57 and (ii) to segregation of the
liquid phase.68

A typical example of the Bi-rich liquid-phase segregation


is illustrated in Fig. 5(b). After sintering the ceramics at
760C, that is, below the Bi25FeO39 peritectic point (T < Tp),
the sillenite phase appeared in the form of lm-sized inclusions, well distributed in the BiFeO3 matrix [see the bright
spots indicated by an arrow in Fig. 5(a)]. Once the ceramics
were annealed at 820C, that is, at T > Tp, the small sillenite
inclusions started to segregate [see the inset of Fig. 5(b)],
forming larger inclusions of up to ~10 lm in size [Fig. 5(b),
see larger bright regions]. Such de-mixing, which is driven by
the liquid-phase formation, spatially prevents the reaction
between the Bi-rich phase and the Fe-rich counterparts,
resulting unavoidably in multiphase ceramics. We note that a
large number of studies found in the literature report sintering temperatures above 820C; such ceramics may thus
contain nonnegligible concentrations of secondary phases.68
Another important issue is the sublimation or evaporation
of the volatile Bi2O3 at elevated temperatures, which may
result into substoichiometric BiFeO3. Our recent experiments,
performed by annealing BiFeO3 ceramics for extended periods (>10 h), suggest Bi2O3 loss at temperatures 820C.68
The results are shown in Fig. 5 and are consistent with the
experiments and conclusions made by Thrall et al.67 To
enrich the atmosphere with bismuth oxide vapors and, thus,
minimize the Bi2O3 loss from the BiFeO3 pellet, we embedded a presintered BiFeO3 pellet (820C, 10 h) into a BiFeO3
packing powder. After postannealing at 820C in the packing
powder, the sillenite phase was clearly detected [Fig. 5(b)]. In
contrast, this phase almost completely disappeared when the
pellet was annealed at the same temperature (820C), but in
open air [Fig. 5(c)], that is, without using the packing
powder. Such a result could be explained by the evaporation
of Bi2O3 from the liquid phase that results from the melting
of the sillenite.43,68
A further increase of the annealing temperature to 880C
resulted in the formation of large (up to 50 lm) regularly
shaped Bi2Fe4O9 crystals [Fig. 5(d)]. Assuming equilibrium
conditions and considering the Bi2O3Fe2O3 phase diagram,31,32 the system appears as though it shifted to a twophase BiFeO3Bi2Fe4O9 region, corresponding to a Bi/Fe
molar ratio <1. Thus, the appearance of the Fe-rich phase is
due to the Bi2O3 loss, as was also observed by other
authors,67,69 and can be represented by a reaction that was
proposed based on thermodynamic calculations:57
8BiFeO3 s ! 2Bi2 Fe4 O9 s 4BiOg O2 g

(3)

where s and g refer to the solid and vapor phases,


respectively. We nally note that the formation of the
Bi2Fe4O9 crystals, like those shown in Fig. 5(d), may be easily confused with those stabilized kinetically and/or by impurities (see this and the previous section). Finding the origin

1998

Journal of the American Ceramic SocietyRojac et al.


(a)

(b)

(c)

(d)

Vol. 97, No. 7

Fig. 5. Segregation of Bi-rich liquid phase and Bi2O3 sublimation/evaporation. Scanning electron microscopy (SEM) backscattered-electron
(BE) images of BiFeO3 ceramics sintered at (a) 760C for 6 h and (b, c, d) at 820C for 10 h with an additional 10 h of postannealing at (b)
820C with the sample immersed in the packing powder, (c) 820C in open air and (d) 880C in open air (both (c) and (d) are denoted as
nonimmersed). The inset of (b) shows the segregation of a Bi-rich liquid phase at 820C as a result of the melting of the Bi25FeO39 phase with
the onset at the peritectic temperature (Tp~790C). The large, regularly shaped Bi2Fe4O9 crystals, identied after annealing at 880C (panel d),
are consistent with the Bi2O3 sublimation loss (see text for details). Reprinted with permission from Ref. [68] Copyright 2010 American Institute
of Physics.

of this secondary phase in BiFeO3 is likely to be a nontrivial


exercise.

(3) Processing Methods


Owing to the particular thermodynamics and kinetics, the
preparation of single-phase BiFeO3 is notoriously dicult.
Numerous and often unsuccessful attempts have been made
using a variety of dierent processing methods. In addition
to the conventional solid-state method, there are many
reports on hydrothermal synthesis, solgel, rapid liquid-phase
sintering, mechanochemical synthesis, precipitation method,
combustion synthesis, and high-pressure synthesis (due to the
large number of publications for each method, references are
not given). It is beyond the scope of this study to review all
of the proposed synthesis methods; instead, we will address
some key problems related to the conventional solid-state
(a)

synthesis of BiFeO3, which may be general and applicable to


other cases. In the second part of this section, we will then
introduce our approach, that is, mechanochemical activation.
There are several problems related to the conventional
multicalcination processing of BiFeO3. The following example from our own work refers to BiFeO3 prepared using the
standard ceramic procedure, including the premilling of
Bi2O3 and Fe2O3 powders in isopropanol using planetary
milling and YSZ milling media, the mixing of the two powders and calcination, followed by a milling step. A typical
sintering curve of a homogenized Bi2O3Fe2O3 powder mixture with a micrometer particle size (median particle size
d50 = 1.1 lm) is shown in Fig. 6(a) (see mixture). To avoid
particle coarsening, which is a general approach, we calcined
the mixture at three dierent temperatures between 650C
and 700C, that is, before the onset of the densication at
800C [Fig. 6(a), mixture]. For all the calcinations the
(b)

Fig. 6. (a) Sintering curve of Bi2O3Fe2O3 mixture before (mixture) and after calcination at 700C for 5 h with additional planetary milling
(calcined + milled). The sintering curve of the mechanochemically activated Bi2O3Fe2O3 mixture (mixture (activated)) is added for
comparison. (b) X-ray diraction (XRD) patterns of Bi2O3Fe2O3 mixture calcined at 650C (6 h), 680C (4 h), and 700C (4 h). The pattern of
the mechanochemically activated mixture, which was reactive sintered at 760C for 6 h (A+RS), is also added.

July 2014

BiFeO3 Ceramics: Processing and Properties

resulting powders contained a signicant amount of secondary phases, which were easily detected by XRD [Fig. 6(b)].
The result is not surprising, taking into account that the
reaction is driven within the temperature region of the thermodynamic instability of BiFeO3, that is, between 447C and
767C [see Fig. 1(b)].
To achieve a higher reaction yield with a given powder
mixture, one of the common approaches is to increase the
diusion rate of the reacting species by increasing the calcination temperature. In the case of BiFeO3, temperatures
higher than 767C would be required to enter the temperature region in which BiFeO3 is stable [see Fig. 1(b)]. There
are, however, two key problems related to this approach.
Firstly, the Bi25FeO39 phase starts to melt at Tp ~ 790C,
which may cause the uncontrolled loss of Bi2O3 through
evaporation and/or segregation of the resulting liquid phase,
as discussed in the previous section. In fact, the slight expansion of the pellet at ~790C, observed on the sintering curve
of the Bi2O3Fe2O3 mixture [Fig. 6(a), mixture], coincides
with the peritectic point of the Bi25FeO39. Secondly, as seen
from the sintering curve [Fig. 6(a), mixture], the densication of the powder mixture sets in at 800C. In principle, this
limits the calcination to below this temperature if we intend
to avoid particle coarsening and, consequently, additional
milling steps, which may increase the possibility of the contamination and stabilization of secondary phases by impurities. We also note that the onset of the densication of the
powder after the rst calcination and milling appears at even
lower temperatures, that is, 600C [Fig. 6(a), calcined +
milled]. Thus, by considering these two arguments, the rst
and subsequent calcination steps would necessarily require
temperatures that fall within the range of the thermodynamic
instability of BiFeO3 or are close to the top-end of this
temperature interval.
To overcome the apparent incompatibility between the
reaction and the densication of conventionally processed BiFeO3, we recently proposed the use of mechanochemical activation (high-energy milling). This method is capable of
providing highly sinterable and reactive precursors, allowing
us to merge the two processes into a single step, known as
reactive sintering.70 We believe that reducing the number of
processing steps, including milling, is indeed important for
the synthesis of BiFeO3, considering the degrading eect of
impurities. The mechanochemical activation of the Bi2O3
Fe2O3 powder mixture resulted in an increased sinterability,
that is, the sintering curve of the activated mixture [Fig. 6(a),
mixture (activated)] is comparable to that of the calcined
and milled powder [Fig. 6(a), calcined + milled]; at the
same time, it also provided increased reactivity between the
Bi2O3 and the Fe2O3. By applying reactive sintering, we were
able to obtain BiFeO3 ceramics in one processing step with a
minimum content of secondary phases [<1%; Fig. 6(b),
A+RS] and a low level of impurities, originating from the
milling media (W: 130 ppm, Co: 390 ppm). Figure 6(b)
(A+RS) shows the case of 760C, however, similar results
were obtained for higher sintering temperatures.68 The
obtained ceramics were able to withstand electric elds as
high as 180 kV/cm, allowing us to study the domain-switching behavior and piezoelectricity of BiFeO3. Those results
are presented in Sections IV and V, while studies on the
dielectric permittivity are presented in the following section.

III.

Dielectric Permittivity, Electrical Conductivity, and


Defects

A comprehensive review, including the most important references on the dielectric permittivity of BiFeO3 across a wide
frequency range, has recently been published by Catalan and
Scott.3 The intrinsic GHz dielectric permittivity of BiFeO3 is
actually small, that is, er~30, however, room-temperature
values as high as er~10000 were reported in the Hz-to-MHz
frequency range, both in ceramics71 and single crystals.72

1999

This is due to the contribution of the electrical conductivity,


which, in addition to directly aecting e (true DC conductivity), it may contribute both to low-frequency e and e
either through hopping or MaxwellWagner (interfacial)
mechanisms.73,74 The latter was invoked to explain the frequency dispersions and high permittivity values of BiFeO3
ceramics and single crystals.3,71,72,75,76 The domain-wall
contribution to the permittivity may be equally important,
but it is generally less often discussed.
It is tempting to assign the dissipation mechanism, for
example, hopping or MaxwellWagner, based on the type of
the observed dielectric relaxation, even when the covered frequency range is limited and two or more mechanisms behave
similarly.74 Since this is often done for BiFeO3, we shall discuss here an example where, qualitatively, dierent types of
relaxations were observed in BiFeO3 that was processed
using the same procedure, but under slightly dierent processing conditions; in this case, the premilling of the starting
Bi2O3 and Fe2O3 powders (Fig. 7).
In the case shown in Fig. 7(a) both e and tand increase
abruptly below 1 Hz with no clear loss peaks. Considering
the classication of Jonscher,73 we could assign this kind of
behavior to the hopping conductivity. As in most experiments, however, only a part of the spectrum is measured and
possible loss peaks and divergence of the loss toward the DC
limit may become apparent only below the experimental frequency limit. Thus, other mechanisms, such as the Maxwell
Wagner, cannot be easily ruled out.74 A similar dispersion to
that in Fig. 7(a) was measured in the case shown in
Fig. 7(b), however, a clear peak in tand was observed at
1 Hz. Even though a step-like increase in e with decreasing
frequency, characteristic for the MaxwellWagner relaxation
(e.g., in CaCu3Ti4O1277) and often reported for BiFeO3,71,72,75,76,78 was not observed, the overall behavior is compatible with this kind of dispersion mechanism. Finally, in
the case of Fig. 7(c), the tand increases with decreasing frequency signicantly more than in the other two cases, reaching huge values (tand = 1100) in the low-frequency range [see
also Fig. 7(d)]. In this sample, the phase angle (d) between
the polarization and the electric eld reached ~90 in the
low-frequency limit, suggesting purely conductive behavior
(the current in phase with the voltage), in which case e is
expected to diverge as the frequency approaches zero.73,74
The true DC conductivity apparently dominates the permittivity response of this sample.
Similar types of relaxations to those shown in Figs. 7(a)
and (b) are frequently reported for BiFeO3, typically in the
Hz-to-MHz range (see, e.g., Refs. [76,7884]). The proposed
dispersion mechanisms in those cases are often speculative.
Although the origin of the diverse dielectric behaviors
shown in Fig. 7 is not clear and appears rather complex, the
data clearly suggest that the low-frequency dielectric dispersion of BiFeO3 ceramics is strongly inuenced by the processing conditions. Thus, signicant qualitative and
quantitative inconsistencies in the literature regarding the
dielectric dispersions in BiFeO3 are not surprising. In most
cases, the identication of the dominant mechanism is not
straightforward, as we illustrate here, and the dielectric
response of BiFeO3 may be case sensitive and dicult to
control.
In a fashion similar to the low-frequency permittivity, the
DC conductivity of BiFeO3 ceramics varies signicantly from
case to case. The as-reported specic DC conductivity of BiFeO3 ceramics and single crystals at room temperature are
spread over several orders of magnitude, typically between
~102 and ~1010 (Ohm m)1.3,56,72,8587 The high conductivity of BiFeO3 is commonly attributed either to the presence
of: (i) reduced Fe3+ (Fe2+ sites), and/or oxygen vacancies
(see, e.g., Refs. [11,76,82,8893]), (ii) secondary phases85,88 or
a combination of these, whereby mechanism (i) found a consensus over the largest number of studies (not all are referenced).

2000

Vol. 97, No. 7

Journal of the American Ceramic SocietyRojac et al.


(a)

(c)

(b)

(d)

Fig. 7. Dielectric permittivity (e ) and dielectric losses (tand) versus eld frequency for three types of BiFeO3 ceramics prepared by the reactive
sintering of a mechanochemically activated Bi2O3Fe2O3 mixture. The three batches dier in the conditions of premilling of the initial Bi2O3 and
Fe2O3 powders: (a) no premilling and (b,c) premilling in a planetary mill at 200 min1 of rotational frequency for 4 h using (b) isopropanol and
(c) acetone. (d) e and tand of the three BiFeO3 ceramics plotted on a logarithmic scale. The median particle sizes (d50) of the starting Bi2O3/
Fe2O3 powders were 14.6/11.3 lm (no premilling), 1.6/0.6 lm (isopropanol milling) and 0.8/0.5 lm (acetone milling). All three ceramics were
reactively sintered at 760C for 6 h [details are given in Ref. (68)], resulting in relative densities in the range 92%95%. e in panel (c) is plotted
in the frequency range 2000.6 Hz; below 0.6 Hz, e decreases and approaches zero as the phase angle reaches ~90 (not shown on the plot). The
real part of the permittivity coecient (e ) can be expressed as e = (Q0/V0)(L/Ae0)cos(d) where Q0, V0, L, A, e0, and d indicate the charge
amplitude, voltage amplitude, sample thickness, sample area, vacuum permittivity, and phase angle, respectively. According to this equation, if
d = 90, e becomes zero, while tand diverges (pure DC conductive behavior).

Recent theoretical studies using density-functional theory


(DFT) suggested that BiFeO3 is actually a p-type semiconductor.94,95 Under oxygen-rich conditions, typical for normal
ceramic processing (i.e., in air atmosphere), the creation of
Bi and Fe vacancies is energetically favored over oxygen
vacancies. Moreover, the cation vacancies form shallow
acceptor levels within the bandgap; for example, the rst ionization transition for Bi vacancies (VBiX?VBi+h) and Fe
vacancies (VFeX?VFe+h) occurs 0.13 and 0.21 eV above the
valence-band edge, respectively. These defects are thus easily
ionized, releasing holes and potentially leading to p-type conduction. On the other hand and as expected, calculations
made for oxygen-poor conditions predicted oxygen vacancies
as the most favorable intrinsic defects in BiFeO3 with, interestingly, the ionization transition (VOX?VO+2e) well or
moderately below the conduction-band edge, that is, 1.994 or
0.6 eV.96 Thus, in comparison with the shallow acceptor levels (cation vacancies), the donor levels (oxygen vacancies) are
rather deep, so that BiFeO3 samples treated in an oxygenpoor environment (leading to n-type conductivity) are, in
principle, not expected to have as high conductivities as those
processed under ordinary oxygen-rich conditions (p-type conductivity). For comparison, a reversed situation was found
experimentally in BaTiO3, which is an oxygen-decient
n-type semiconductor and an insulator under oxygen-excess
conditions with a weak p-type character.97 In the absence
of systematic experimental data on BiFeO3 it is useful to
look at other Fe-containing complex oxides. Interestingly,
Sr-doped LaFeO3 was found experimentally to exhibit analogous behavior to that predicted for BiFeO3, with conductivity that is orders of magnitude lower at low oxygen partial
pressures [r(789C) = 10 (Ohm m)1 at p(O2) = 1020] than
at high partial pressures [r(789C) = 4000 (Ohm m)1 at p
(O2) = 1].98
The theoretical predictions discussed above for the conductivity of BiFeO3 are supported by recent experimental
results. In addition to some sporadic studies on BiFeO3 that
suggest p-type conductivity,78,99102 it was shown more

directly that the conductivity of Ca-doped BiFeO3 could be


reduced by several orders of magnitude if the ceramics were
sintered in an oxygen-poor atmosphere (e.g., in N2), consistent with the p-type conduction behavior.103 Thus, Ca-doped
BiFeO3 is a p-type semiconductor with an activation energy
(Ea) of ~0.270.4 eV when sintered in oxygen or air, whereas
it is an ionic conductor with Ea~0.821.04 eV when sintered
and cooled in nitrogen (Fig. 8). The semiconductivity was
attributed to the mixed Fe3+/Fe4+ valance state. This is in
agreement with other studies, which indicate the absence of
Fe2+ in BiFeO3-based ceramics,79,91 favoring p-type conductivity.

Fig. 8. DC electric conductivity versus inverse of temperature


(Arrhenius plot) for Ca-doped BiFeO3 ceramics sintered in air,
oxygen (O2), and nitrogen (N2). When sintered in air or O2 the
ceramics behave as a p-type semiconductor with activation energy
(Ea) close to 0.3 eV. Sintering in N2 reduces signicantly the
conductivity with Ea~0.8 eV, showing ionic conduction behavior.
Reprinted with permission from Ref. [103] Copyright 2012 American
Chemical Society.

July 2014

It appears, from recent studies, that the conductivity in BiFeO3-related systems can be signicantly reduced by thermal
treatments in an oxygen-poor atmosphere; however, this
leaves an open question as to whether the Bi2O3 loss through
sublimation and/or evaporation may be controlled under
such conditions. In fact, a lack of oxygen during annealing
may enhance the Bi2O3 loss [see Section II(2) and reaction
(3)], an issue that has been initiated,69,104 but so far not
studied in detail. A related problem found in argon-annealed
BiFeO3 is a poor microstructure with a high level of porosity, which has its origin, apparently, in the loss of Bi2O3.67

IV.

Ferroelectric and Ferroelastic Domain-Switching


Behavior

(1) High-Field Polarization Hysteresis Loops


Studies on the ferroelectric switching behavior of BiFeO3
have encountered considerable diculties, which have their
origin in the unfortunate combination of the high electrical
conductivity and the high coercive eld of this material, the
latter already predicted by Michel et al. in 1969.29 Due to
the high leakage current and low breakdown eld, it is often
the case, even in very recent studies, that authors report
subcoercive (unsaturated) polarization-electric-eld (PE)
loops, which are often confused with the switching (saturated) PE loops. This problem has recently been addressed
by J. F. Scott,105 but has a tendency to persist. We note,
however, that both PE and strain-electric-eld (SE) loops
measured under subcoercive eld conditions, if done properly, may contain very rich information about the domainwall dynamics, dielectric, and piezoelectric behavior of the
material, including nonlinearity, hysteresis, conductivity contributions, or coupled eects.106 These issues are discussed
in Section V.
Some representative PE loops of BiFeO3 from the literature, reminiscent of those resulting from switching in other
ferroelectric ceramics, are shown in Fig. 9.85,90,107 In some
cases, such as those shown in Figs. 9(a) and (b), the loops
are pinched (see arrows), whereas in other cases [Fig. 9(c)
and Refs. (12,108,109)] they appear more open. The three
loops, shown in Fig. 9, exhibit dierent degree of internal
bias, manifested as a shift along the eld axis. Despite the
comparable electric-eld amplitude (150 kV/cm), the apparent remanent polarization (2Pr) in these cases varies by a factor of three, that is, ~20 lC/cm2 < 2Pr < ~60 lC/cm2,
whereas all the loops show a coercive eld of Ec ~ 75 kV/
cm. The signicant qualitative and quantitative dierences in
the switching behavior among dierent samples call for a
systematic study of the processingproperties relationship in
BiFeO3.

(a)

2001

BiFeO3 Ceramics: Processing and Properties

(b)

Pinched or double ferroelectric hysteresis loops, like the


ones shown in Fig. 9(a) and (b) and also observed by other
authors,110112 indicate the presence of domain-wall pinning
centers, which aect the switching behavior and induce aging
of the materials properties.110 Pinched loops are commonly
observed in acceptor-doped, often called hard ferroelectric
materials, such as Mn and Fe-doped Pb(Zr,Ti)O3
(PZT)113,114 and Mn-doped BaTiO3.115,116 They were also
measured in undoped materials, for example, PZT,106, BaTiO3,117 and (K,Na)NbO3 (KNN).118 In all these cases, the
loop pinching is usually attributed to the pinning of the
domain walls by charged defects, which are shown or
assumed to be acceptoroxygenvacancy (VO) defect complexes (e.g., FeZr,TiVO in Fe-doped PZT or TiTiVO in
reduced undoped BaTiO3119).
Considering the accepted view of hardening in these
acceptor-doped perovskites, one may assume that similar
domain-wall pinning centers, that is, acceptorVO complexes, probably also operate in BiFeO3. We note, however,
that this assumption, which was made previously in the literature,68,110 was based on the macroscopic behavior of BiFeO3, such as, loop pinching, depinching and aging, which
appears similar to that in other hard ferroelectrics (see
later in this section). In the absence of more direct proofs,
other pinning mechanisms and/or pinning defects that may
be specic to BiFeO3 cannot be excluded and are discussed
at the end of this section.
In terms of the processingdefects relationship, we rst
examine the cooling-rate eect, which may inuence the
order/disorder state of the charged defects with respect to the
polarization within the domains, and, consequently, the state
of pinching of the PE loop.
Several microscopic mechanisms have been proposed to
explain the pinning eect and aging behavior in PZT and BaTiO3.113,115 Given sucient time, the acceptorVO defect
complexes may arrange through the diusion of oxygen
vacancies into an ordered conguration by aligning along the
spontaneous polarization within a domain;113,115117,119125 in
this ordered state, the complexes clamp the domain walls,
giving rise to the observed pinched PE loops. The domain
walls may be depinned and, thus, the loop may be depinched, by converting the arrangement of the defects from the
ordered to a disordered state. One of the ways to do this is
to rst heat the material above TC, that is, into the paraelectric phase where the defects are relaxed (disordered) due to
the absence of the spontaneous polarization. This is followed
by rapid cooling, which provides freezing of the disordered
defect state at room temperature. The quenching experiments
can give indirect, but valuable information about the interaction between domain walls and pinning defects.106,114

(c)

Fig. 9. Selected polarizationelectric-eld (PE) hysteresis loops of BiFeO3 ceramics from the literature. The ceramics were prepared using (a)
solgel powder processing [sintered at 800C in O2 with high heating (30C/min) and cooling rate (not specied)],90 (b) rapid liquid-phase
sintering [sintered at 855C with high heating (100C/min) and cooling rate (~170C/min)],85 and (c) solid-state reaction method [sintered at
860C with high cooling rate (quenching medium not reported)].107 Arrows in panels (a) and (b) indicate the pinched state of the loops.
Reprinted with permission from Ref. [90] (Copyright 2006 American Institute of Physics), Ref. [85] (Copyright 2006 Elsevier Ltd) and Ref. [107]
(Copyright 2005 American Institute of Physics).

2002

Journal of the American Ceramic SocietyRojac et al.

Figure 10(a) shows the PE loops of BiFeO3 ceramics


quenched into water from various temperatures.68 The opening of the loop is only achieved by rapid cooling from
T > TC, that is, from 900C (TC = 825C). In contrast, minor
changes in the PE loops are observed upon quenching from
temperatures T < TC, that is, 450C and 760C, as compared
to the initial, nonquenched state (non-Q). Note the signicant reduction in the apparent coercive eld and the increase
in the remanent polarization after quenching from 900C. As
described above, these changes suggest disordering of the
pinning centers through quenching, which results in a
reduced pinning eect and facilitated polarization switching
and domain-wall motion. The fact that the PE loops open
after quenching from T > TC is consistent with the assumption that the material has to be brought into the paraelectric
phase, where the defects may be disordered due to the
absence of the spontaneous polarization.
The strong inuence of the defects on the switching behavior of BiFeO3 is further conrmed by experiments conducted
with dierent cooling rates [Fig. 10(b)]. Reducing the cooling
rate, from quenching in air to slow cooling within a furnace,
while keeping the same target temperature (900C), resulted
in a progressive reappearance of the loop pinching and the
associated reduction of Pr, like that observed in Fe-doped
hard PZT.106 The results are thus consistent with the gradual reordering of the defects as the cooling rate is decreased.
It is interesting to note that the loop of the sample quenched
in air [Fig. 10(b), Q-air] exhibited a stronger internal bias
eld in comparison with the sample slowly cooled within the
furnace [Fig. 10(b), slow]; this might suggest that the
defects in the air-quenched sample partially ordered along a
preferred direction, which is, in this case, the direction of the
applied electric eld during the PE measurements.113 We
also note that the initial state before quenching [Fig. 10(a),
non-Q] could not be completely recovered, even with 1C/
min of cooling rate [Fig. 10(b), slow], suggesting that the
defects encountered diculties in rearranging back into the
ordered pinning positions or the presence of other irreversible contributions to the polarization. The discrepancies in
the switching behavior of dierent BiFeO3 samples reported
in the literature, such as the dierent level of loop pinching
(see Fig. 9), may be in part explained by the dierent cooling
rates employed, which are often not reported and controlled.
Based on the pinched PE loops and the associated depinching or de-aging by quenching, indicative of the presence of
domain-wall pinning centers, an obvious question opens up
as to which defects cause the pinning of domain walls in BiFeO3. Various possibilities, which we next discuss, have been
considered,68 however, the exact origin is still not clear.
(a)

Vol. 97, No. 7

Assuming the scenario of defect complexes, which is


known to play a major role in, for example, acceptor-doped
PZT and BaTiO3,119 possible pinning centers for the domain
walls in BiFeO3 would be the VBiVO complexes, created
due to the loss of Bi2O3 at elevated temperatures. These
types of complexes, that is, A-site-vacancyoxygen-vacancy
pairs, are believed to be responsible for the hardening
behavior of undoped PZT.125
To evaluate the role of the defects created in BiFeO3 by
Bi2O3 loss, we performed a set of experiments in which we
controlled the atmosphere during annealing by using a packing powder. The BiFeO3 was thus annealed with or without
immersing the ceramics into a packing powder, and the loss
of Bi2O3 was estimated through the analysis of the phase
composition of the ceramics by means of SEM. These results
were discussed in Section II(2) (Fig. 5). In contrast to the
ceramics annealed at 820C in a packing powder [Fig. 5(b)],
the BiFeO3 annealed at temperatures 820C in open air
(without the packing powder) showed evidence of the Bi2O3
loss [Figs. 5(c) and (d); see also discussion in Section II(2)].
The PE loops of the ceramics annealed with and without
the packing powder are shown in Fig. 11. A pinched hysteresis loop is observed after annealing the BiFeO3 in a packing
powder [Fig. 11(a), full line], whereas the loop opens up,
showing a decreased coercive eld and increased remanent
polarization, after annealing in air, that is, without covering
the pellet with the packing powder [Fig. 11(b)]. Note also the
signicant dierence between the loops of the two samples
when compared with the same electric-eld amplitude
(90 kV/cm): the polarization response of the ceramics treated
in the packing powder [Fig. 11(a), dashed line] is much
weaker than that of the sample annealed in air [Fig. 11(b),
full red line]. This depinching of the loop by annealing the
ceramics in the absence of the packing powder is qualitatively similar to that observed by quenching the ceramics
[Fig. 10(a)] and it is thus indicative of a domain-wall depinning (details are reported in Ref. [68]). Although without
direct evidence, we could infer from these results that the
defects created in BiFeO3 as a result of the Bi2O3 loss, which,
considering the analogous case of undoped PZT, could be
the VBiVO complexes, do not cause domain-wall pinning.
Instead, a depinning eect is seen as the Bi2O3 is lost
through evaporation/sublimation. This behavior of BiFeO3
thus appears qualitatively dierent from that of PZT.
In analogy with the TiTiVO complexes in BaTiO3, which
may be formed during processing at low oxygen partial pressures,119 and considering the role of acceptorVO complexes
in the hardening behavior of PZT and BaTiO3,113,115117,119
the next possibility that might be considered for BiFeO3 are
(b)

Fig. 10. Inuence of quenching temperature and cooling rate on the domain-switching behavior of BiFeO3. (a) Polarizationelectric-eld (PE)
hysteresis loops of as-sintered (nonquenched) ceramics (non-Q) and after rapid quenching in water from 450C, 760C, and 900C.68 The holding
time at each temperature was 5 min. (b) PE loops of ceramics annealed at 900C for 5 min, followed by quenching in water (Q-water) and air
(Q-air), and by slow cooling within the furnace with 1C/min of cooling rate (slow). Reprinted with permission from Ref. [68] Copyright 2010
American Institute of Physics.

July 2014

BiFeO3 Ceramics: Processing and Properties


(b)

(a)

Fig. 11. Inuence of Bi2O3 loss upon annealing on the domainswitching behavior of BiFeO3. Polarizationelectric-eld (PE)
hysteresis loops of BiFeO3 sintered at 820C for 10 h with an
additional 10 h of postannealing at 820C or 840C with the ceramic
sample (a) immersed in a BiFeO3 packing powder or (b) treated in
air (without the packing powder). Note that the loops shown in
panels (a) and (b) were measured at dierent electric-eld amplitudes
(140 and 90 kV/cm, respectively). To allow a comparison under the
same electric-eld conditions, the PE loop of the sample annealed
in the packing powder and measured at 90 kV/cm is added in panel
(a) (see dashed line).

the FeFeVO complexes, where FeFe is Fe2+.68 We note


that these complexes have been proposed to explain the
switching behavior of N2-annealed BiFeO3 thin lms.126
Based on recent calculations and conductivity measurements
(see Section III), however, Fe2+ does not appear likely in BiFeO3 processed under regular oxygen-rich sintering conditions (air atmosphere).79,91 Those studies rather suggest the
presence of Fe4+. The complex nature of the defects that creates such a strong pinning eect in the undoped BiFeO3 is
still to be claried and requires further studies.
It is well accepted that the defect complexes play an
important role in (deliberately or non-) acceptor-doped perovskites, such as PZT and BaTiO3.115117,119123,127,128 Even
though the macroscopic behavior of BiFeO3 (e.g., pinched
PE loop) is consistent with the behavior of these perovskites, the defect structure may be dierent in BiFeO3, and
other domain-wall pinning mechanisms must be considered,

(a)

2003

for example, the pinning defects may be located within the


domain walls. In addition, one should probably also consider
other pinning defects, such as dislocations.129,130
In the case of BiFeO3, the defects located within the
domain walls might play a role in the domain-wall pinning.
In fact, several studies on BiFeO3 thin lms131134 show evidence of conductive 180, 109, and 71 domain walls, indicating the presence of mobile charge carriers at the walls, as
assumed originally by Carl and H
ardtl113 and Postnikov
et al.135,136 in their microscopic models. In addition, headto-head domain congurations were directly identied by
piezoresponse-force microscopy (PFM) in fatigued (111)pc
oriented BiFeO3 thin lm.7 The associated compensating
charges at these domain walls may cause pinning, as was
proposed for the BiFeO3 thin lms,7 and/or, in principle,
may reduce the wall motion under an applied eld.113,135,136
On the other hand, the model of Li et al.137 suggests that the
concentration of the neutralizing charged defects within the
domain walls may lead to either an enhancement or reduction of the piezoelectric eect, depending on whether the
charges in the walls are under or overcompensated. These
results and the observations for thin lms are consistent with
the observed macroscopic behavior of the BiFeO3 ceramics:
the domain-switching strain [see Section IV(2)] and the irreversible domain-wall contribution to d33 (see Section V) are
considerably enhanced once the domain walls are depinned,
which occurs by driving the BiFeO3 ceramics with low driving-eld frequencies, that is, <1 Hz.86,138

(2) High-Field Strain Hysteresis Loops


In contrast to the numerous polarization-switching studies,
data on the electric-eld-induced strain in BiFeO3 ceramics
under switching conditions are limited.68,86,112,139 We recently
measured a large bipolar strain of up to ~0.4% peak-to-peak
at a low frequency (0.1 Hz) and high amplitude (140 kV/cm)
of the driving eld.86 The strainelectric-eld (SE) loops of
BiFeO3, measured with increasing eld amplitude, are shown
in Fig. 12(a). The loop at 160 kV/cm shows a clear buttery-like shape [see the inset of Fig. 12(a)], suggesting switching and movement, particularly of the non-180 domain
walls, which, depending on the switchable lattice strain,
involves a signicant change in the dimensions of the grains
in ferroelectric ceramics.140142 We note that the pinched-like
strain loops with small remanence, measured at low-eld
amplitudes [see the loop measured at 110 kV/cm denoted
with an arrow in Fig. 12(a)], are consistent with the pinning
of domain walls by defects, in agreement with the PE data
(see previous section). The loops evolve progressively with

(b)

Fig. 12. (a) Strainelectric-eld (SE) hysteresis loops of BiFeO3 ceramics measured with increasing eld amplitudes at 0.1 Hz (the inset shows
the loop at 160 kV/cm). (b) Peak-to-peak strain as a function of driving-eld frequency for 110, 120, and 130 kV/cm. The data shown in panel
(b) were compiled from measurements reported in Ref. [86]. As indicated with arrows in panel (a), the SE loops evolve from a pinched-like
state with little remanent strain at low driving-eld amplitudes (110 kV/cm) to a de-pinched state at high amplitudes (160 kV/cm).

2004

Journal of the American Ceramic SocietyRojac et al.

increasing eld, that is, from a pinched state at low elds to


an opened or de-pinched state at high elds [see the loop
evolution with eld in Fig. 12(a)]. In agreement with the
polarization-eld studies,68 this loop evolution suggests
progressive wall depinning possibly by the reorientation (disordering) of defects.
A particular aspect of the high-eld strain response of BiFeO3 ceramics is its strong dependence on the frequency of
the driving eld [Fig. 12(b)]. Especially at lower eld amplitudes [Fig. 12(b), see 110 and 120 kV/cm], the strain exhibits
an abrupt increase below 1 Hz. Two origins were proposed
for the observed release of the strain at low driving frequencies: (i) multistep non-180 domain switching, which was predicted by Kubel and Schmidt30 and conrmed in BiFeO3 thin
lms;6,7 and (ii) rearrangement (disordering) of the pinning
centers.86 The latter scenario is consistent with the domainwall depinning eect conrmed by quenching (see the previous section) and eld cycling [see Fig. 12(a) and Ref. (68)].
Interestingly, the same strong frequency dependence is also
seen in the weak-to-moderate eld piezoelectric d33 coecient
(see next section), suggesting that both the high- and
weak-to-moderate eld responses of BiFeO3 at low driving
frequencies are strongly aected by the dynamic interaction
between the domain walls and the pinning centers.
Some authors have reported that the large electric-eldinduced strain in BiFeO3-based thin lms is related to the
switching between tetragonal-like and rhombohedral-like
phases.143145 This does not seem to be the dominant mechanism in polycrystalline bulk BiFeO3. The role of non-180
domain switching in the large strain response of the BiFeO3
ceramics was conrmed by synchrotron X-ray diraction
analysis. Details of the measurement and analysis methods as
applied to other compositions have been reported previously,
respectively, in Refs. [146,147] and [148,149].
We next discuss the results of our own observations of
domain reorientation and its potential contribution to the
switching strain of BiFeO3. The studies were performed using
X-ray diraction analysis and were carried out ex situ on
poled and unpoled samples. The samples were prepared by
the authors using the mechanochemical method [details of
the synthesis procedure are reported in Ref. (68)].
In rhombohedral ceramics, such as BiFeO3, the extent of
non-180 domain switching is determined by measuring the
relative intensities of (111)pc and (111)pc diraction peaks
along a particular specimen direction. Figure 13(a) illustrates
the intensity changes of {111}pc reections as a function of
the orientation of the poled BiFeO3 sample relative to the
electric-eld direction. Parallel to the electric-eld direction,
that is, at the angle of 0, the intensity of the (111)pc peak,
(a)

Vol. 97, No. 7

relative to the (11


1)pc, is stronger than what would be
expected for randomly oriented rhombohedral ceramics (in
which case the (111) pc:(11
1)pc intensity ratio would be close
to 1:3). The dierence in the present measurement is due to
domain switching that occurred during the poling process,
which preferentially oriented the rhombohedral polar [111]pc
direction more closely to the electric-eld direction. In addition, a decrease in the intensity of the (111)pc peak is evident
with increasing angle relative to the electric-eld direction,
that is, toward the angle of 90, which shows the distribution
of the preferred orientation, or domain texture, present in
the poled ceramics.
Using the diracted intensities, the extent of domain orientation induced by poling can be quantied. For this purpose,
the {111}pc peaks were tted with two Gaussian proles and
the integrated intensities were extracted. A degree of domain
orientation was then calculated using the ratios of the integrated intensities for {111}pc peaks in the poled (oriented)
state relative to those in the unpoled (nonoriented) state in
all sample directions, dened with respect to the electriceld direction. The degree of orientation can be represented
as either a multiple of a random distribution (MRD or f111)
or as a fraction of domains reoriented (g111). These values
quantify the same preferred orientations or orientation distributions, but use dierent units.148
Figure 13(b) shows the MRD (f111) or, equivalently, the
fraction of [111]pc-oriented domains that were reoriented
(g111) as a function of the angle to the electric eld in the
poled ceramic BiFeO3. For nonoriented ceramics, we would
expect f111 = 1 and g111 = 0 [as indicated by the dashed line
in Fig. 13(b)], whereas lower or higher values, respectively,
correspond to ceramics with a decreased or increased volume
fraction of [111]pc-oriented domains in a particular direction.149 Thus, with respect to the nonoriented (unpoled)
state, the electric eld applied to the BiFeO3 during poling
lowered the fraction of the [111]pc domains (g111 < 0) at
angles >45 away from the eld direction, whereas it
increased the fraction (g111 > 0) along directions more closely
oriented to the eld direction (<45). This can be interpreted
as a eld-induced [111]pc preferred orientation in the ceramic,
which occurs through non-180 domain switching, that is,
through 71 and 109 in the case of rhombohedral BiFeO3.
The angle where g111 = 0 [Fig. 13(b)] corresponds to the orientation along which the average of all [111]pc domains lie in
an energetically equivalent state with respect to the eld
direction.
Figure 14 illustrates a simplied schematic of a hypothetical non-180 domain-wall movement or switching process
under an applied eld with the corresponding increase in the
(b)

Fig. 13. (a) Portion of X-ray diraction (XRD) pattern and (b) calculated degree of domain alignment in poled BiFeO3 as a function of the
angle to the applied electric eld. Before XRD analysis, the BiFeO3 ceramic sample was poled with electric elds of increasing amplitude from 10
to 160 kV/cm in steps of 10 kV/cm. The corresponding SE loops of this sample are shown in Fig. 12(a). The degree of domain alignment can
be represented as either multiples of random distribution (MRD) for the polar [111]pc direction (f111) or the domain switching fraction (g111);
both of these values are shown as separate y-axes in panel b. An initially unpoled state would be represented as f111 = 1 MRD and g111=0 in all
directions of the polycrystalline ceramic; such a state is shown on panel b as a dashed line.

July 2014

BiFeO3 Ceramics: Processing and Properties

Fig. 14. Schematics of non-180 domain-wall movement or


switching process under a eld (E) and the resulting increase in the
fraction of [111]pc-oriented domains (g111) along the direction of the
applied electric eld where [111]pc is the rhombohedral polar axis.
The deformation (DL) of a hypothetical grain that accompanies the
switching process is also illustrated. For simplicity, the angles
between the polarization vectors are drawn as 90.

fraction of [111]pc-oriented domains along the eld axis and


the decrease of this fraction along the direction perpendicular
to the applied eld. Note the accompanying deformation of
the grain as switching occurs (denoted as DL on the schematics).
Being coupled to strain, the non-180 domain-wall motion
contributes to the macroscopic strain realized in ferroelectric
materials by application of the electric eld. In Jones
et al.,150 a relationship is given through which the non-180
domain-wall orientation distribution, such as that shown in
Fig. 13(b), and the spontaneous ferroelastic strain can be
used to determine the macroscopic strain resulting from such
domain reorientation. Using this approach, we found that
the degree of the domain orientation for the poled BiFeO3
ceramics [Fig. 13(b)] yields a longitudinal strain contribution
of 0.21% relative to the unpoled state. This large strain is
similar to the magnitude of the measured remanent strain,
shown in Fig. 12(a) (see strain at zero eld), reinforcing the
conclusion that the signicant contribution to the macroscopic electric-eld-induced strain of BiFeO3 comes from the
non-180 domain-wall motion and ferroelectric/ferroelastic
domain-switching processes (see also schematics in Fig. 14)
and not from phase transformation, as reported for epitaxial
thin lms.
Finally, we note that the magnitude of the switching
strain of BiFeO3 ceramics is comparable to that measured
in lead-based ferroelectric ceramics with morphotropic phase
boundary composition, such as PZT and PMN-PT.86 This
is exceptional for a simple oxide such as BiFeO3 and opens
up opportunities in the search for potentially ecient leadfree systems based on the ferrite. A good, recent example,
which conrms this, is the BiFeO3BaTiO3 MPB system
with an excellent piezoelectric response and a high TC
(~600C).15,17,18

V.

Direct Piezoelectric Response

Despite the intensive research on BiFeO3, an understanding


of the macroscopic piezoelectric response of BiFeO3 ceramics
has been progressing slowly. Some of our recent studies,
however, reveal a peculiar piezoelectric response of the
ferrite, which is discussed in detail in this section.138 Crosschecking the values reported for the piezoelectric d33 coecient,12,13,78,85,151153 we found a very broad range of
reported d33 values, that is, from 2 to 60 pm/V. The problem

2005

is primarily the high electric conductivity of the BiFeO3,


which prevents the application of suciently high electric
elds to the material for an ecient and reproducible poling.
A characterization of the direct piezoelectric response of
BiFeO3 ceramics, including the eld and frequency dependence of the d33 coecient and the piezoelectric phase angle,
was recently carried out by the authors. The details are
reported in Rojac et al.138 Here, we focus on the behavior at
low driving stress frequencies, where the piezoelectric
response of the BiFeO3 exhibits a signicant enhancement.
As shown in Fig. 15(a), the ferrite exhibits a strong frequency dispersion of d33 and tand below ~1 Hz. In the lowfrequency range (0.021 Hz), the d33 increases from 26 pC/N
(1 Hz) to 35 pC/N (0.02 Hz), that is, by 35%; likewise, the
piezoelectric tand exhibits an abrupt increase from 0.03
(1 Hz) to 0.19 (0.02 Hz). The lossy low-frequency piezoelectric response of the BiFeO3 is reected in the strongly
hysteretic charge density (D) stress (r) loops measured at
0.1 and 0.01 Hz [Fig. 15(c)]. In contrast to these two loops,
a small hysteresis with tand<0.05 was measured at frequencies 1 Hz [see the loops measured at 1 Hz and 10 Hz in
Fig. 15(c)].
The low-frequency d33 dispersion [Fig. 15(a), <1 Hz] was
found to be linked to the piezoelectric nonlinearity, that is, it
correlates with the amplitude of the driving eld.138 This is
seen in Fig. 15(b), which shows the plots of d33 versus stress
amplitude for dierent stress frequencies. An increase of d33
with increasing amplitude is observed for all the measured
frequencies; however, there are signicant quantitative dierences. While the maximum relative increase in d33 with stress
is similar at 10 and 1 Hz [between 12% and 16%; see the relative d33 at 3.2 MPa in Fig. 15(b)], d33 increases with eld up
to 40% and 66% when the material is cycled with 0.1 and
0.01 Hz, respectively. The piezoelectric nonlinear response is
therefore strongly enhanced at frequencies below 1 Hz, in
agreement with the d33 frequency dispersion [Fig. 15(a)]. In
relative terms, the maximum increase of d33 with the driving
stress at the lowest measured frequency (66% at 0.01 Hz) is
comparable to that measured in Nb-doped soft PZT and
coarse-grained BaTiO3.154
Macroscopic piezoelectric nonlinearity and hysteresis in
ferroelectric materials are usually attributed to extrinsic origins, most commonly to the irreversible movement of ferroelectric-ferroelastic non-180 domain walls under an applied
eld.149,154157 This implies a signicant contribution of the
irreversible domain-wall motion in BiFeO3 at low driving frequencies [Fig. 15(b)].138 The same process also gives rise to
energy dissipation in the material and a hysteretic response.
Figure 15(d) shows the Dr loops measured at 0.01 Hz with
increasing stress amplitude. As the amplitude is increased, in
addition to an increase in the slope of the loop, which is proportional to d33, the piezoelectric response of the BiFeO3
becomes increasingly hysteretic. The piezoelectric tand was
also found to increase with increasing stress amplitude (not
shown, see Rojac et al.138), which is expected for the irreversible contribution from non-180 domain-wall displacements.
We note, however, that an additional linear (stress-independent) and frequency-dispersive mechanism operates in the
background and contributes to the overall d33 and piezoelectric dispersion.138
What distinguishes BiFeO3 from other piezoelectrically
nonlinear ceramics, such as the morphotropic PZT, is its
low-frequency nonlinearity [Fig. 15(b)] and the associated
dispersion of d33 at low driving frequencies [Fig. 15(a)]. This
behavior is very dierent from the rather general logarithmic
frequency dependence of the piezoelectric coecient and the
nonlinearity observed in soft morphotropic PZT and some
other materials, in which d33 increases linearly with decreasing log(x) where x is the frequency of the driving eld.156,158
Considering that the piezoelectric nonlinearity and hysteresis
in BiFeO3 originates from the irreversible non-180 domainwall motion, the low-frequency dispersion of d33 suggests

2006

Journal of the American Ceramic SocietyRojac et al.


(a)

Vol. 97, No. 7

(b)

(c)

(d)

Fig. 15. Direct piezoelectric response of BiFeO3 ceramics. (a) Piezoelectric d33 coecient and piezoelectric losses (tand) as a function of the
frequency of the driving stress (measured at 1.1 MPa of peak-to-peak stress amplitude); (b) relative piezoelectric coecient (d33 relative) versus
amplitude of the driving stress, measured at 10, 1, 0.1, and 0.01 Hz;138 (c, d) charge densitystress loops measured with increasing (c) frequency
and (d) amplitude of driving stress. Note the considerable increase of the slope and the hysteresis of the loops as the frequency of the driving
eld is reduced (panel c) or the amplitude is increased (panel d). The static stress was set to either 2.7 MPa (panel a) or 3.7 MPa (panels b, c,
and d).

that the domain-wall motion in BiFeO3 is dierent than in


these other ferroelectric materials.
A possible explanation for the particular behavior of BiFeO3 is the coupling between the domain-wall motion and the
electric conductivity, which is reected in the lossy loops
measured at high amplitudes and low frequencies of the driving eld [Figs. 15(c) and (d)]. Assuming that mobile charge
carriers are accumulated at the domain-wall area [see discussion in Section IV(1)], the domain walls will move under an
applied external eld, providing there is a suciently long
eld exposure time (equivalent to low frequencies). This is
because domain-wall motion requires charge migration, that
is, electrical conduction in the material. The hypothesis is
consistent with the measured nonlinearity and hysteresis at
low driving frequencies [Figs. 15(b),(c),(d)]. In addition, possible local conductivity at domain walls, as found in BiFeO3
thin lms,131,133 could aect the macroscopic piezoelectric
response, its frequency dispersion and hysteresis through the
Maxwell-Wagner piezoelectric eect.159 A strong indication
of this mechanism is the negative piezoelectric phase angle,
which was actually measured in BiFeO3 ceramics.138 All these
possibilities have recently been considered to explain the
macroscopic piezoelectric behavior of the ferrite;138 however,
more focused studies are needed to conrm them.

VI.

Summary

In ceramic form, BiFeO3 exhibits a number of interesting


functional properties, however, it presents diculties when it
comes to processing, both of which are systematically
reviewed in this contribution. An analysis of the thermodynamics and kinetics of this system unveiled the multiple origins of the frequently observed formation and stabilization
of the Bi-rich and Fe-rich secondary phases. These nonperovskite phases can be stabilized by:

1. spontaneous decomposition of the BiFeO3 in the temperature range 447C767C according to the reaction
[see Eq. (1)]:.
BiFeO3 ! 1=49Bi25 FeO39 12=49Bi2 Fe4 O9

(4)

2. Contamination with impurities that are likely to


incorporate into the Bi25FeO39 and/or Bi2Fe4O9
phases, which leads to the stabilization of these phases
according to the Bi2O3Fe2O3AOx ternary phase relations (AOx is an impurity oxide and also includes
refractory oxides, such Al2O3 (AOx = AlO1.5) and
SiO2, which are in often contact with the BiFeO3 during thermal treatments),
3. kinetic reasons, whereby, due to solid-state diusionlimited processes, the reaction between Bi2O3 and
Fe2O3 remains incomplete with the two secondary
phases in the ceramics,
4. melting of the easily formed Bi25FeO39 phase, leading
to segregation and de-mixing of the Bi-rich phase and
Fe-rich reaction counterparts,
5. sublimation and/or evaporation of Bi2O3 at elevated
temperatures, which leads to the formation of typically
large (tenths of lm) Bi2Fe4O9 crystals in the ceramics.
All these second-phase sources may operate simultaneously, rendering the synthesis of BiFeO3 and the identication of the origin of the secondary phases rather dicult. An
ideal processing method for BiFeO3 should take into
consideration all of these issues.
In contrast to the general belief that Fe2+ and/or oxygen
vacancies are responsible for the high electric conductivity
of BiFeO3, recent theoretical and experimental studies suggest that it is p-type conductivity for ceramics processed

July 2014

BiFeO3 Ceramics: Processing and Properties

under ordinary oxygen-rich conditions (air atmosphere).


Thus, it appears that the conductivity of the ferrite could be
controlled and lowered by a heat treatment in an oxygendecient environment, such as in N2, as shown recently for
Ca-doped BiFeO3. Systematic experiments, such as conductivity versus the partial pressure of oxygen in a wide pressure
range (e.g., p(O2) = 10201), would be an important step
toward an understanding of the defect chemistry of the ferrite and its electrical conductivity.
Macroscopic evidences exist to suggest that BiFeO3
behaves as a hard ferroelectric material with typically
pinched or constricted PE and SE loops. We showed that
the overall behavior of BiFeO3, including de-aging or depinching by quenching and electric-eld cycling, is consistent
with the accepted view of hardening in other perovskites,
such as the acceptor-doped PZT and BaTiO3. In these cases
the domain walls are pinned by charged defects, involving
oxygen vacancies, and the wall movement and switching in a
eld is thus reduced. It has to be emphasized, however, that
the nature of the pinning centers in BiFeO3 may be dierent
from those in acceptor-doped PZT or BaTiO3, despite many
similarities in the macroscopic behavior.
Once the domain walls are depinned from the defects,
which occurs by application of elds of high amplitude and
low frequency, the BiFeO3 produces large eld-induced
strains, comparable to those measured in morphotropic leadbased ferroelectric ceramics, such as PZT and PMN-PT. Synchrotron XRD analysis conrmed that the large strains in
BiFeO3 ceramics are due to the switching and movement of
non-180 domain walls, and not due to the motion of phase
boundaries, as reported for thin lms.
The piezoelectric response of BiFeO3 is characterized by a
strong nonlinearity and hysteresis, which we attribute to irreversible domain-wall displacements. At high driving stress
amplitudes and low stress frequencies, this domain-wall contribution reaches levels comparable to those measured in
donor-doped PZT and BaTiO3. This is rather unexpected,
considering the hard nature of the BiFeO3 and the strong
pinning of domain walls as seen through the domain-switching behavior. This inconsistency appears to be, in part, reconciled by the particular dependence of the domain-wall
contribution on the frequency of the applied stress. Indeed,
the irreversible non-180 domain-wall motion in BiFeO3 is
strongly restricted to low driving frequencies (<1 Hz). This
low-frequency nonlinear piezoelectric behavior suggests coupling between the domain-wall motion and the conductivity
and has not been reported so far for other ferroelectrics. The
behavior could have multiple origins, but might be, in part,
caused by the conductive domain walls, which have been
extensively studied in BiFeO3 thin lm, but not yet shown to
exist in polycrystalline bulk BiFeO3.

Acknowledgments
This work was supported by the Slovenian Research Agency (programme P20105 and project J2-5483). TR would like to thank Prof. Dr. Nava Setter for
her nancial and technical support related to parts of this work. DD acknowledges the nancial support of FNS-PNR62. JJ and GT acknowledge the U.S.
Department of the Army for support under contract number W911NF-09-10435. JD acknowledges nancial support from an AINSE research fellowship
and ARC DP120103968. Use of the Advanced Photon Source was supported
by the US Department of Energy, Oce of Science, Oce of Basic Energy
Sciences, under Contract No. DE-AC02-06CH11357.

References
1

J. Wang, J. B. Neaton, H. Zheng, V. Nagarajan, S. B. Ogale, B. Liu, D.


Viehland, V. Vaithyanathan, D. G. Schlom, U. V. Waghmare, N. A. Spaldin,
K. M. Rabe, M. Wuttig, and R. Ramesh, Epitaxial BiFeO3 Multiferroic Thin
Film Heterostructures, Science, 299 [5613] 171922 (2003).
2
J. Yu and J. Chu, Progress and Prospect for High Temperature SinglePhased Magnetic Ferroelectrics, Chi. Sci. Bull., 53 [14] 2097112 (2008).
3
G. Catalan and J. F. Scott, Physics and Applications of Bismuth Ferrite,
Adv. Mater., 21 [24] 246385 (2009).
4
D. Lebeugle, D. Colson, A. Forget, and M. Viret, Very Large Spontaneous Electric Polarization in BiFeO3 Single Crystals at Room Temperature and

2007

Its Evolution Under Cycling Fields, Appl. Phys. Lett., 91 [2] 022907, 3pp
(2007).
5
N. Balke, S. Choudhury, S. Jesse, M. Huijben, Y. H. Chu, A. P. Baddorf,
L. Q. Chen, R. Ramesh, and S. V. Kalinin, Deterministic Control of Ferroelastic Switching in Multiferroic Materials, Nat. Nanotechnol., 4 [12] 86875
(2009).
6
S. H. Baek, H. W. Jang, C. M. Folkman, Y. L. Li, B. Winchester, J. X.
Zhang, Q. He, Y. H. Chu, C. T. Nelson, M. S. Rzchowski, X. Q. Pan, R. Ramesh, L. Q. Chen, and C. B. Eom, Ferroelastic Switching for Nanoscale
Non-Volatile Magnetoelectric Devices, Nat. Mater., 9 [4] 30914 (2010).
7
S. H. Baek, C. M. Folkman, J. W. Park, S. Lee, C. W. Bark, T. Tybell,
and C. B. Eom, The Nature of Polarization Fatigue in BiFeO3, Adv. Mater.,
23 [14] 16215 (2011).
8
J. R
odel, W. Jo, K. T. P. Seifert, E. M. Anton, T. Granzow, and D. Damjanovic, Perspective on the Development of Lead-Free Piezoceramics, J.
Am. Ceram. Soc., 92 [6] 115377 (2009).
9
R. Mohan, Green Bismuth, Nat. Chem, 2 [4] 336 (2010).
10
J. M. Moreau, C. Michel, R. Gerson, and W. J. James, Ferroelectric BiFeO3 X-Ray and Neutron Diraction Study, J. Phys. Chem. Solids, 32 [6]
131520 (1971).
11
G. L. Yuan and S. W. Or, Enhanced Piezoelectric and Pyroelectric
Eects in Single-Phase Multiferroic Bi1xNdxFeO3 (x = 00.15) Ceramics,
Appl. Phys. Lett., 88 [6] 062905, 3pp (2006).
12
C. Sun, X. Chen, J. Wang, G. Yuan, J. Yin, and Z. Liu, Structure and
Piezoelectric Properties of BiFeO3 and Bi0.92Dy0.08FeO3 Multiferroics at High
Temperature, Solid State Commun., 152 [14] 11948 (2012).
13
X. Chen, Y. Wang, Y. Yang, G. Yuan, J. Yin, and Z. Liu, Structure,
Ferroelectricity and Piezoelectricity Evolutions of Bi1xSmxFeO3 at Various
Temperatures, Solid State Commun., 152 [6] 497500 (2012).
14
T. P. Comyn, S. P. McBride, and A. J. Bell, Processing and Electrical
Properties of BiFeO3PbTiO3 Ceramics, Mater. Lett., 58 [30] 38446 (2004).
15
H. Yang, C. Zhou, X. Liu, Q. Zhou, G. Chen, H. Wang, and W. Li,
Structural, Microstructural and Electrical Properties of BiFeO3BaTiO3
Ceramics With High Thermal Stability, Mater. Res. Bull., 47 [12] 42339
(2012).
16
T. P. Comyn, T. Stevenson, and A. J. Bell, Piezoelectric Properties of BiFeO3PbTiO3 Ceramics, J. Phys. IV France, 128 [1] 37 (2005).
17
S. O. Leontsev and R. E. Eitel, Dielectric and Piezoelectric Properties in
Mn-Modied (1X)BiFeO3XBaTiO3 Ceramics, J. Am. Ceram. Soc., 92 [12]
295761 (2009).
18
S. O. Leontsev and R. E. Eitel, Origin and Magnitude of the Large Piezoelectric Response in the Lead-Free (1X)BiFeO3XBaTiO3 Solid Solution,
J. Mater. Res., 26 [1] 917 (2011).
19
S. Karimi, I. M. Reaney, Y. Han, J. Pokorny, and I. Sterianou, Crystal
Chemistry and Domain Structure of Rare-Earth Doped BiFeO3 Ceramics, J.
Mater. Sci., 44 [19] 510212 (2009).
20
S. Fujino M. Murakami, V. Anbusathaiah, S.-H. Lim, V. Nagarajan, C. J.
Fennie, M. Wuttig, L. Salamanca-Riba, and I. Takeuchi, Combinatorial Discovery of a Lead-Free Morphotropic Phase Boundary in a Thin-Film Piezoelectric Perovskite, Appl. Phys. Lett., 92 [20] 202904, 3pp (2008).
21
D. Kan, L. Palova, V. Anbusathaiah, C. J. Cheng, S. Fujino, V. Nagarajan, K. M. Rabe, and I. Takeuchi, Universal Behavior and Electric-FieldInduced Structural Transition in Rare-Earth-Substituted BiFeO3, Adv. Funct.
Mater., 20 [7] 110815 (2010).
22
T. L. Ivanova and V. V. Gagulin, Dielectric Properties in the Microwave
Range of Solid Solutions in the BiFeO3SrTiO3 System, Ferroelectrics, 265
[1] 2416 (2002).
23
N. Itoh, T. Shimura, W. Sakamoto, and T. Yogo, Eects of SrTiO3 Content and Mn Doping on Dielectric and Magnetic Properties of BiFeO3SrTiO3
Ceramics, J. Ceram. Soc. Jpn., 117 [9] 93943 (2009).
24
Z. Z. Ma, Z. M. Tian, J. Q. Li, C. H. Wang, S. X. Huo, H. N. Duan, and
S. L. Yuan, Enhanced Polarization and Magnetization in Multiferroic (1X)
BiFeO3XSrTiO3 Solid Solution, Solid State Sci., 13 [12] 2196200 (2011).
25
J. Wu and J. Wang, Multiferroic Behavior of BiFeO3RTiO3 (Mg, Sr,
Ca, Ba, and Pb) Thin Films, J. Appl. Phys., 108 [2] 026101, 3pp (2010).
26
Q. Q. Wang, Z. Wang, X. Q. Liu, X. M. Chen, and D. W. Johnson,
Improved Structure Stability and Multiferroic Characteristics in CaTiO3Modied BiFeO3 Ceramics, J. Am. Ceram. Soc., 95 [2] 6705 (2012).
27
M. I. Morozov, M. A. Einarsrud, and T. Grande, Polarization and Strain
Response in Bi0.5K0.5TiO3BiFeO3 Ceramics, Appl. Phys. Lett., 101 [25]
252904, 4pp (2012).
28
K. Yazawa, S. Yasui, H. Morioka, T. Yamada, H. Uchida, A. Gruverman, and H. Funakubo, Composition Dependence of Crystal Structure and
Electrical Properties for Epitaxial Films of Bi(Zn1/2Ti1/2)O3BiFeO3 Solid
Solution System, J. Ceram. Soc. Jpn., 118 [8] 65963 (2010).
29
C. Michel, J. M. Moreau, G. Achenbach, R. Gerson, and W. J. James,
The Atomic Structure of BiFeO3, Solid State Commun., 7 [9] 7014 (1969).
30
F. Kubel and H. Schmid, Structure of a Ferroelectric and Ferroelastic
Monodomain Crystal of the Perovskite BiFeO3, Acta Crystallogr. B, 46 [6]
698702 (1990).
31
A. Ma^tre, M. Francois, and J. C. Gachon, Experimental Study of the Bi2O3
Fe2O3 Pseudo-Binary System, J. Phase Equilib. Dius., 25 [1] 5967 (2004).
32
R. Palai, R. S. Katiyar, H. Schmid, P. Tissot, S. J. Clark, J. Robertson, S.
A. T. Redfern, G. Catalan, and J. F. Scott, Phase and c-b Metal-Insulator
Transition in Multiferroic BiFeO3, Phys. Rev. B, 77 [1] 014110, 11pp (2008).
33
E. I. Speranskaya, V. M. Skorikov, E. Y. Rode, and V. A. Terekhova,
The Phase Diagram of the System Bismuth Oxide Ferric Oxide, Bull.
Acad. Sci. U.S.S.R, 5 [87] 34 (1965).
34
S. M. Skinner, Magnetically Ordered Ferroelectric Materials, IEEE
Trans. PMP, 6 [2] 6890 (1970).

2008

Journal of the American Ceramic SocietyRojac et al.

35
V. S. Filipev, N. P. Smolyaninov, E. G. Fesenko, and I. N. Belyaev,
Synthesis of BiFeO3 and Determination of the Unit Cell, Sov. Phys. Crystallogr., 5, 9134 (1960).
36
S. A. Fedulov, Y. N. Venevtsev, G. S. Zhdanov, and E. G. Smazhevskaya,
High-Temperature X-Ray and Thermal-Analysis Results for Bismuth
Ferrite, Sov. Phys. Crystallogr., 6, 6401 (1961).
37
G. D. Achenbach, W. J. James, and R. Gerson, Preparation of SinglePhase Polycrystalline BiFeO3, J. Am. Ceram. Soc., 50 [8] 4377 (1967).
38
M. I. Morozov, N. A. Lomanova, and V. V. Gusarov, Specic Features
of BiFeO3 Formation in a Mixture of Bismuth(III) and Iron(III) Oxides,
Russ. J. Gen. Chem., 73 [11] 167680 (2003).
39
M. W. Lufaso, T. A. Vanderah, I. M. Pazos, I. Levin, R. S. Roth, J. C.
Nino, V. Provenzano, and P. K. Schenck, Phase Formation, Crystal Chemistry, and Properties in the System Bi2O3Fe2O3Nb2O5, J. Solid State Chem.,
179 [12] 390010 (2006).
40
T. T. Carvalho and P. B. Tavares, Synthesis and Thermodynamic Stability of Multiferroic BiFeO3, Mater. Lett., 62 [24] 39846 (2008).
41
S. M. Selbach, M. A. Einarsrud, and T. Grande, On the Thermodynamic
Stability of BiFeO3, Chem. Mater., 21 [1] 16973 (2009).
42
S. Phapale, R. Mishra, and D. Das, Standard Enthalpy of Formation
and Heat Capacity of Compounds in the Pseudo-Binary Bi2O3Fe2O3 System, J. Nucl. Mater., 373 [13] 13741 (2008).
43
M. Valant, A. K. Axelsson, and N. Alford, Peculiarities of a Solid-State
Synthesis of Multiferroic Polycrystalline BiFeO3, Chem. Mater., 19 [22] 5431
6 (2007).
44
H. A. Harwig, On the Structure of Bismuthsesquioxide: The a, b, c, and
d-Phase, Z. Anorg. Allg. Chem., 444 [1] 15166 (1978).
45
L. G. Sillen, X-Ray Studies on Bismuth Trioxide, Ark. Kemi. Mineral.
Geol., 12A [18] 115 (1937).
46
E. M. Levin and R. S. Roth, Polymorphism of Bismuth Sesquioxide. II
Eect of Oxide Additions on the Polymorphism of Bi2O3, J. Res. Nat. Bur.
Stand. A, 68A [2] 197206 (1964).
47
D. C. Craig and N. C. Stephenson, Structural Studies of Some BodyCentered Cubic Phases of Mixed Oxides Involving Bi2O3: The Structures of
Bi25FeO40 and Bi38ZnO60, J. Solid State Chem., 15 [1] 18 (1975).
48
T. Rojac, M. Kosec, B. Malic, and J. Holc, Mechanochemical Synthesis
of NaNbO3, Mater. Res. Bull., 40 [2] 3415 (2005).
49
B. Yu, M. Li, J. Wang, L. Pei, D. Guo, and X. Zhao, Enhanced Electrical Properties in Multiferroic BiFeO3 Ceramics Co-Doped by La3+and V5+,
J. Phys. D, 41 [18] 185401, 5pp (2008).
50
W. S. Kim, Y. K. Jun, K. H. Kim, and S. H. Hong, Enhanced Magnetization in Co and Ta-Substituted BiFeO3 Ceramics, J. Magn. Magn. Mater.,
321 [19] 32625 (2009).
51
F. Azough, R. Freer, M. Thrall, R. Cernik, F. Tuna, and D. Collison,
Microstructure and Properties of Co-, Ni-, Zn-, Nb- and W-Modied Multiferroic BiFeO3 Ceramics, J. Eur. Ceram. Soc., 30 [3] 72736 (2010).
52
Z. Chaodan, Y. Jun, Z. Duanming, Y. Bin, W. Yunyi, W. Longhai, W.
Yunbo, and Z. Wenli, Processing and Ferroelectric Properties of Ti-Doped
BiFeO3 Ceramics, Integr. Ferroelectr., 94 [1] 316 (2007).
53
S. J. Kim, S. H. Han, H. G. Kim, A. Y. Kim, J. S. Kim, and C. l. Cheon,
Multiferroic Properties of Ti-Doped BiFeO3 Ceramics, J. Korean Phys. Soc.,
56 [12] 43942 (2010).
54
I. O. Troyanchuk, A. N. Chobot, O. S. Mantytskaya, and N. V. Tereshko,
Magnetic Properties of Bi(Fe1 xMx)O3 (M = Mn, Ti), Inorg. Mater., 46
[4] 4248 (2010).
55
M. S. Bernardo, T. Jardiel, M. Peiteado, A. C. Caballero, and M. Villegas,
Sintering and Microstuctural Characterization of W6+, Nb5+ and Ti4+ IronSubstituted BiFeO3, J. Alloy. Compd., 509 [26] 72906 (2011).
56
M. S. Bernardo, T. Jardiel, M. Peiteado, F. J. Mompean, M. Garcia-Hernandez, M. A. Garcia, M. Villegas, and A. C. Cabalero, Intrinsic Compositional Inhomogeneities in Bulk Ti-Doped BiFeO3: Microstructure
Development and Multiferroic Properties, Chem. Mater., 25 [9] 153341
(2013).
57
S. M. Selbach, T. Tybell, M. A. Einarsrud, and T. Grande, Phase Transitions, Electrical Conductivity and Chemical Stability of BiFeO3 at High Temperatures, J. Solid State Chem., 183 [5] 12058 (2010).
58
R. Haumont, I. Kornev, S. Lisenkov, L. Bellaiche, J. Kreisel, and B.
Dkhil, Phase Stability and Structural Temperature Dependence in Powdered
Multiferroic BiFeO3, Phys. Rev. B, 78 [13] 134108, 8pp (2008).
59
J. D. Bucci, B. K. Robertson, and W. J. James, The Precision Determination of the Lattice Parameters and the Coecients of Thermal Expansion of
BiFeO3, J. Appl. Crystallogr., 5 [3] 18791 (1972).
60
M. S. Bernardo, T. Jardiel, M. Peiteado, A. C. Caballero, and M. Villegas,
Reaction Pathways in the Solid State Synthesis of Multiferroic BiFeO3, J.
Eur. Ceram. Soc., 31 [16] 304753 (2011).
61
M. V. Slinkina and G. I. Doncov, Difuzija Sovstvenih Komponentov V
Keramike Cirkonata-Titanata Svinca., Neorgani
ceskie Materiali., 28 [3] 567
70 (1992).
62
B. Malic, D. Jenko, J. Holc, M. Hrovat, and M. Kosec, Synthesis of
Sodium Potassium Niobate: A Diusion Couples Study, J. Am. Ceram. Soc.,
91 [6] 191622 (2008).
63
P. J. Harrop, Self-Diusion in Simple Oxides (a Bibliography), J.
Mater. Sci., 3 [2] 20622 (1968).
64
J. L. Mukherjee and F. F. Y. Wang, Kinetics of Solid-State Reaction of
Bi2O3 and Fe2O3, J. Am. Ceram. Soc., 54 [1] 314 (1971).
65
A. Beauger, J. C. Mutin, and J. C. Niepce, Synthesis Reaction of Metatitanate BaTiO3, J. Mater. Sci., 18 [12] 354350 (1983).
66
M. R
ossel, H. R. H
oche, H. S. Leipner, D. V
oltzke, H. P. Abicht, O.
Hollricher, J. M
uller, and S. Gablenz, Raman Microscopic Investigations of

Vol. 97, No. 7

BaTiO3 Precursors With Core-Shell Structure, Anal. Bioanal. Chem., 380 [1]
15762 (2004).
67
M. Thrall, R. Freer, C. Martin, F. Azough, B. Patterson, and R. J. Cernik, An in Situ Study of the Formation of Multiferroic Bismuth Ferrite
Using High Resolution Synchrotron X-Ray Powder Diraction, J. Eur.
Ceram. Soc., 28 [13] 25672 (2008).
68
T. Rojac, M. Kosec, B. Budic, N. Setter, and D. Damjanovic, Strong
Ferroelectric Domain-Wall Pinning in BiFeO3 Ceramics, J. Appl. Phys., 108
[7] 074107, 8pp (2010).
69
M. Li and J. L. MacManus-Driscoll, Phase Stability, Oxygen Nonstoichiometry and Magnetic Properties of BiFeO3d, Appl. Phys. Lett., 87 [25]
252510, 3pp (2005).
70
L. B. Kong, J. Ma, T. S. Zhang, W. Zhu, and O. K. Tan, Pb(ZrxTi1 x)
O3 Ceramics Via Reactive Sintering of Partially Reacted Mixture Produced by
a High-Energy Ball Milling Process, J. Mater. Res., 16 [6] 163643 (2001).
71

S. Kamba, D. Nuzhnyy, M. Savinov, J. Sebek,
J. Petzelt, J. Proleska, R.
Haumont, and J. Kreisel, Infrared and Terahertz Studies of Polar Phonons
and Magnetodielectric Eect in Multiferroic BiFeO3 Ceramics, Phys. Rev. B,
75 [2] 024403, 7pp (2007).
72
J. Lu, A. G
unther, F. Schrettle, F. Mayr, S. Krohns, P. Lunkenheimer, A.
Pimenov, V. D. Travkin, A. A. Mukhin, and A. Loidl, On the Room Temperature Multiferroic BiFeO3: Magnetic, Dielectric and Thermal Properties,
Eur. Phys. J. B, 75 [4] 45160 (2010).
73
A. K. Jonscher, Dielectric Relaxation in Solids. Chelsea Dielectrics Press,
U.K., 1983.
74
M. I. Morozov and D. Damjanovic, Charge Migration in Pb(Zr,Ti)O3
Ceramics and Its Relation to Ageing, Hardening, and Softening, J. Appl.
Phys., 107 [3] 034106, 10pp (2010).
75
R. Mazumder, S. Ghosh, P. Mondal, D. Bhattacharya, S. Dasgupta, N.
Das, A. Sen, A.K. Tyagi, M. Sivalumar, T Takami, and H. Ikuta, Particle
Size Dependence of Magnetization and Phase Transition Near Tn in Multiferroic BiFeO3, J. Appl. Phys., 100 [3] 033908, 9pp (2006).
76
S. Hunpratub, P. Thongbai, T. Yamwong, R. Yimnirun, and S. Maensiri,
Dielectric Relaxations and Dielectric Response in Multiferroic BiFeO3
Ceramics, Appl. Phys. Lett., 94 [6] 062904, 3pp (2009).
77
P. Lunkenheimer, R. Fichtl, S. G. Ebbinghaus, and A. Loidl, Nonintrinsic Origin of the Colossal Dielectric Constants in CaCu3Ti4O12, Phys. Rev. B,
70 [17] 172102, 4pp (2004).
78
Z. Dai and Y. Akishige, Electrical Properties of Multiferroic BiFeO3
Ceramics Synthesized by Spark Plasma Sintering, J. Phys. D, 43 [44] 445403,
5pp (2010).
79
J. Chen, X. Xing, A. Watson, W. Wang, R. Yu, J. Deng, L. Yan, C. Sun,
and X. Chen Rapid Synthesis of Multiferroic BiFeO3 Single-Crystalline
Nanostructures, Chem. Mater., 19 [15] 3598600 (2007).
80
M. Kumar, K. L. Yadav, and G. D. Varma, Large Magnetization and
Weak Polarization in Sol-Gel Derived BiFeO3 Ceramics, Mater. Lett., 62 [8
9] 115961 (2008).
81
R. Mazumder, D. Chakravarty, D. Bhattacharya, and A. Sen, Spark
Plasma Sintering of BiFeO3, Mater. Res. Bull., 44 [3] 5559 (2009).
82
R. Mazumder and A. Sen, Eect of Pb-Doping on Dielectric Properties
of BiFeO3 Ceramics, J. Alloy Compd., 475 [12] 57780 (2009).
83
H. O. Rodrigues, G. F. M. Pires Junior, A. J. M. Sales, P. M. O. Silva, B.
F. O. Costa, P. Alcantara Jr, S. G. C. Moreira, and A. S. B. Sombra, BiFeO3
Ceramic Matrix With Bi2O3 or PbO Added: M
ossbauer, Raman and Dielectric
Spectroscopy Studies, Phys. B, 406 [13] 25329 (2011).
84
Z. H. Dai and Y. Akishige, BiFeO3 Ceramics Synthesized by Spark
Plasma Sintering, Ceram. Int., 38, S4036 (2012).
85
G. L. Yuan, S. W. Or, Y. P. Wang, Z. G. Liu, and J. M. Liu, Preparation and Multi-Properties of Insulated Single-Phase BiFeO3 Ceramics, Solid
State Commun., 138 [2] 7681 (2006).
86
T. Rojac, M. Kosec, and D. Damjanovic, Large Electric-Field Induced
Strain in BiFeO3 Ceramics, J. Am. Ceram. Soc., 94 [12] 410811 (2011).
87
E. Markiewicz, B. Hilczer, M. Baszyk, A. Pietraszko, and E. Talik,
Dielectric Properties of BiFeO3 Ceramics Obtained From Mechanochemically
Synthesized Nanopowders, J. Electroceram., 27 [34] 15461 (2011).
88
Y. P. Wang, L. Zhou, M. F. Zhang, X. Y. Chen, J. M. Liu, and Z. G.
Liu, Room-Temperature Saturated Ferroelectric Polarization in BiFeO3
Ceramics Synthesized by Rapid Liquid Phase Sintering, Appl. Phys. Lett., 84
[10] 17313 (2004).
89
A. K. Pradhan, K. Zhang, D. Hunter, J. B. Dadson, G. B. Loutts, P.
Bhattacharya, R. Katiyar, J. Zhang, D. J. Sellmyer, U. N. Roy, Y. Cui, and
A. Burger, Magnetic and Electrical Properties of Single-Phase Multiferroic
BiFeO3, J. Appl. Phys., 97 [9] 093903, 4pp (2005).
90
F. Chen, Q. F. Zhang, J. H. Li, Y. J. Qi, C. J. Lu, X. B. Chen, X. M.
Ren, and Y. Zhao, Sol-Gel Derived Multiferroic BiFeO3 Ceramics With
Large Polarization and Weak Ferromagnetism, Appl. Phys. Lett., 89 [9]
092910, 3pp (2006).
91
G. L. Yuan, K. Z. Baba-Kishi, J. M. Liu, S. W. Or, Y. P. Wang, and Z.
G. Liu, Multiferroic Properties of Single Phase Bi0.85La0.15FeO3 Lead-Free
Ceramics, J. Am. Ceram. Soc., 89 [10] 31369 (2006).
92
H. Ke, W. Wang, Y. Wang, H. Zhang, D. Jia, Y. Zhou, X. Lu, and P.
Withers, Dependence of Dielectric Behavior in BiFeO3 Ceramics on Intrinsic
Defects, J. Alloy. Compd., 541, 948 (2012).
93
B. Yu, M. Li, J. Liu, D. Guo, L. Pei, and X. Zhao, Eects of Ion Doping at Dierent Sites on Electrical Properties of Multiferroic BiFeO3 Ceramics, J. Phys. D, 41 [6] 065003, 4pp (2008).
94
Z. Zhang, P. Wu, L. Chen, and J. Wang, Density Functional Theory Plus
U Study of Vacancy Formations in Bismuth Ferrite, Appl. Phys. Lett., 96
[23] 232906, 3pp (2010).

July 2014

BiFeO3 Ceramics: Processing and Properties

95
T. R. Paudel, S. S. Jaswal, and E. Y. Tsymbal, Intrinsic Defects in
Multiferroic BiFeO3 and Their Eect on Magnetism, Phys. Rev. B, 85 [10]
104409, 8pp (2012).
96
S. J. Clark and J. Robertson, Energy Levels of Oxygen Vacancies in
BiFeO3 by Screened Exchange, Appl. Phys. Lett., 94 [2] 022902, 3pp (2009).
97
M. V. Raymond and D. M. Smyth, Defects and Charge Transport in
Perovskite Ferroelectrics, J. Phys. Chem. Solids, 57 [10] 150711 (1996).
98
I. Wrnhus, T. Grande, and K. Wiik, Surface Exchange of Oxygen in
La1xSrxFeO3d (x = 0, 0.1), Top. Catal., 54 [1315] 100915 (2011).
99
A. S. Poghossian, H. V. Abovian, P. B. Avakian, S. H. Mkrtchian, and V.
M. Haroutunian, Bismuth Ferrites: New Materials for Semiconductor Gas
Sensors, Sens. Actuators B, 4 [34] 5459 (1991).
100
C. Wang, M. Takahashi, H. Fujino, X. Zhao, E. Kume, T. Horiuchi, and
S. Sakai, Leakage Current of Multiferroic (Bi0.6 Tb0.3La0.1)FeO3 Thin Films
Grown at Various Oxygen Pressures by Pulsed Laser Deposition and Annealing Eect, J. Appl. Phys., 99 [5] 054104, 5pp (2006).
101
B. Vengalis, J. Devenson, A.K. Oginskis, R. Butkute, A. Maneikis, A.
Steikuniene, L. Dapkus, J. Banys, and M. Kinka, Growth and Investigation
of Heterostructures Based on Multiferroic BiFeO3, Acta Phys. Pol. A, 113 [3]
10958 (2008).
102
T. H. Kim, B. C. Jeon, T. Min, S. M. Yang, D. Lee, Y. S. Kim, S.-H. Baek,
W. Saenrang, C.-B. Eom, T. K. Song, J.-G. Yoon, and T. W. Noh, Continuous
Control of Charge Transport in Bi-Decient BiFeO3 Films Through Local Ferroelectric Switching, Adv. Funct. Mater., 22 [23] 49628 (2012).
103
N. Mas
o and A. R. West, Electrical Properties of Ca-Doped BiFeO3
Ceramics: From p-Type Semiconduction to Oxide-Ion Conduction, Chem.
Mater., 24 [11] 212732 (2012).
104
M. C. Li, J. Driscoll, L. H. Liu, and L. C. Zhao, The Phase Transition
and Phase Stability of Magnetoelectric BiFeO3, Mater. Sci. Eng. A, 438440,
3469 (2006).
105
J. F. Scott, Ferroelectrics Go Bananas, J. Phys. Condens. Mat., 20 [2]
021001, 2pp (2008).
106
D. Damjanovic, Hysteresis in Piezoelectric and Ferroelectric Materials;
pp. 337465 in The Science of Hysteresis, Vol. 3. Edited by G. Bertotti and I.
Mayergoyz. Academic Press, Oxford, U.K., 2006.
107
S. T. Zhang, M. H. Lu, D. Wu, Y. F. Chen, and N. B. Ming, Larger Polarization and Weak Ferromagnetism in Quenched BiFeO3 Ceramics With a Distorted
Rhombohedral Crystal Structure, Appl. Phys. Lett., 87 [26] 262907, 3pp (2005).
108
S. T. Zhang, L. H. Pang, Y. Zhang, M. H. Lu, and Y. F. Chen, Preparation, Structures, and Multiferroic Properties of Single Phase Bi1xLaxFeO3
(x = 00.40) Ceramics, J. Appl. Phys., 100 [11] 114108, 6pp (2006).
109
W. N. Su, D. H. Wang, Q. Q. Cao, Z. D. Han, J. Yin, J. R. Zhang, and
Y. W. Du, Large Polarization and Enhanced Magnetic Properties in BiFeO3
Ceramic Prepared by High-Pressure Synthesis, Appl. Phys. Lett., 91 [9]
092905, 3pp (2007).
110
G. L. Yuan, Y. Yang, and S. W. Or, Aging-Induced Double Ferroelectric Hysteresis Loops in BiFeO3 Multiferroic Ceramic, Appl. Phys. Lett., 91
[12] 122907, 3pp (2007).
111
S. Zhang, L. Wang, Y. Chen, D. Wang, Y. Yao, and Y. Ma, Observation of Room Temperature Saturated Ferroelectric Polarization in Dy Substituted BiFeO3 Ceramics, J. Appl. Phys., 111 [7] 074105, 5pp (2012).
112
A. Y. Kim, Y. J. Lee, J. S. Kim, S. H. Han, H. W. Kang, H. G. Lee,
and C. I. Cheon, Ferroelectric Properties of BiFeO3 Ceramics Sintered Under
Low Oxygen Partial Pressure, J. Korean Phys. Soc., 60 [1] 837 (2012).
113
K. Carl and K. H. Hardtl, Electrical After-Eects in Pb(Ti,Zr)O3
Ceramics, Ferroelectrics, 17 [1] 47386 (1977).
114
M. I. Morozov and D. Damjanovic, Hardening-Softening Transition in
Fe-Doped Pb(Zr,Ti)O3 Ceramics and Evolution of the Third Harmonic of the
Polarization Response, J. Appl. Phys., 104 [3] 034107, 8pp (2008).
115
P. V. Lambeck and G. H. Jonker, The Nature of Domain Stabilization
in Ferroelectric Perovskites, J. Phys. Chem. Solids, 47 [5] 45361 (1986).
116
L. Zhang and X. Ren, Aging Behavior in Single-Domain Mn-Doped BaTiO3 Crystals: Implication for a Unied Microscopic Explanation of Ferroelectric Aging, Phys. Rev. B, 73 [9] 094121, 6pp (2006).
117
X. Ren, Large Electric-Field-Induced Strain in Ferroelectric Crystals by
Point-Defect-Mediated Reversible Domain Switching, Nat. Mater., 3 [2] 914
(2004).
118
H. Birol, D. Damjanovic, and N. Setter, Preparation and Characterization of (K0.5Na0.5)NbO3 Ceramics, J. Eur. Ceram. Soc., 26 [6] 8616 (2006).
119
R. A. Eichel, Structural and Dynamic Properties of Oxygen Vacancies in
Perovskite Oxides Analysis of Defect Chemistry by Modern Multi-Frequency
and Pulsed EPR Techniques, Phys. Chem. Chem. Phys., 13 [2] 36884 (2011).
120
U. Robels and G. Arlt, Domain Wall Clamping in Ferroelectrics by Orientation of Defects, J. Appl. Phys., 73 [7] 345460 (1993).
121
W. L. Warren, D. Dimos, G. E. Pike, K. Vanheusden, and R. Ramesh,
Alignment of Defect Dipoles in Polycrystalline Ferroelectrics, Appl. Phys.
Lett., 67 [12] 168991 (1995).
122
W. L. Warren, G. E. Pike, K. Vanheusden, D. Dimos, B. A. Tuttle, and
J. Robertson, Defect-Dipole Alignment and Tetragonal Strain in Ferroelectrics, J. Appl. Phys., 79 [12] 92507 (1996).
123
W. L. Warren, K. Vanheusden, D. Dimos, G. E. Pike, and B. A. Tuttle,
Oxygen Vacancy Motion in Perovskite Oxides, J. Am. Ceram. Soc., 79 [2]
5368 (1996).
124
P. Erhart, P. Traskelin, and K. Albe, Formation and Switching of
Defect Dipoles in Acceptor-Doped Lead Titanate: A Kinetic Model Based on
First-Principles Calculations, Phys. Rev. B, 88 [2] 024107, 10pp (2013).
125
A. Chandrasekaran, D. Damjanovic, N. Setter, and N. Marzari, Defect
Ordering and DefectDomain-Wall Interactions in PbTiO3: A First-Principles
Study, Phys. Rev. B, 88 [21] 214116, 7pp (2013).

2009

126
Z. Wen, G. Hu, C. Yang, and W. Wu, Eect of Annealing Process on
Asymmetric Coercitivities of Mn-Doped BiFeO3 Thin Films, Appl. Phys. A,
97, 93741 (2009).
127
Z. Feng and X. Ren, Striking Similarity of Ferroelectric Aging Eect in
Tetragonal, Orthorhombic and Rhombohedral Crystal Structures, Phys. Rev.
B, 77 [13] 134115, 6pp (2008).
128
L. Zhang, E. Erdem, X. Ren, and R. A. Eichel, Reorientation of
(MnTiVo)9 Defect Dipoles in Acceptor-Modied BaTiO3 Single Crystals:
An Electron Paramagnetic Resonance Study, Appl. Phys. Lett., 93 [20]
202901, 3pp (2008).
129
A. Kontsos and C. M. Landis, Computational Modeling of Domain
Wall Interactions With Dislocations in Ferroelectric Crystals, Int. J. Solids
Struct., 46 [6] 14918 (2009).
130
A. Lubk, M. D. Rossell, J. Seidel, Y. H. Chu, R. Ramesh, M. J. Hytch,
and E. Snoeck, Electromechanical Coupling Among Edge Dislocations,
Domain Walls, and Nanodomains in BiFeO3 Revealed by Unit-Cell-Wise
Strain and Polarization Maps, Nano Lett., 13 [4] 14105 (2013).
131
J. Seidel, L. W. Martin, Q. He, Q. Zhan, Y.-H. Chu, A. Rother, M. E.
Hawkridge, P. Maksymovych, P. Yu, M. Gajek, N. Balke, S. V. Kalinin, S.
Gemming, F. Wang, G. Catalan, J. F. Scott, N. A. Spaldin, J. Orenstein, and
R. Ramesh, Conduction at Domain Walls in Oxide Multiferroics, Nat.
Mater., 8 [3] 22934 (2009).
132
J. Seidel, P. Maksymovych, Y. Batra, A. Katan, S.-Y. Yang, Q. He, A.
P. Baddorf, S.V. Kalinin, C.-H. Yang, J.-C. Yang, Y.-H. Chu, E. K. H. Salje,
H. Wormeester, M. Salmeron, and R. Ramesh, Domain Wall Conductivity
in La-Doped BiFeO3, Phys. Rev. Lett., 105 [19] 197603 (2010).
133
S. Farokhipoor and B. Noheda, Conduction Through 71 Domain Walls
in BiFeO3 Thin Films, Phys. Rev. Lett., 107 [12] 127601 (2011).
134
P. Maksymovych, J. Seidel, Y. H. Chu, P. Wu, A. P. Baddorf, L. Q.
Chen, S. V. Kalinin, and R. Ramesh, Dynamic Conductivity of Ferroelectric
Domain Walls in BiFeO3, Nano Lett., 11 [5] 190612 (2011).
135
V. S. Postnikov, V. S. Pavlov, and S. K. Turkov, Internal Friction in
Ferroelectrics Due to Interaction of Domain Boundaries and Point Defects,
J. Phys. Chem. Solids, 31 [8] 178591 (1970).
136
V. S. Postnikov, V. S. Pavlov, S. A. Gridnev, and S. K. Turkov, Interaction Between 90o Domain Walls and Point Defects of the Crystal Lattice in
Ferroelectric Ceramics, Sov. Phys. Solid State, 10 [6] 126770 (1968).
137
Z. Li, H. Wu, and W. Cao, Piezoelectric Response of Charged Non180 Domain Walls in Ferroelectric Ceramics, J. Appl. Phys., 111 [2] 024106,
5pp (2012).
138
T. Rojac, A. Bencan, G. Drazic, M. Kosec, and D. Damjanovic, Piezoelectric Nonlinearity and Frequency Dispersion of the Direct Piezoelectric
Response of BiFeO3 Ceramics, J. Appl. Phys., 112 [6] 064114, 12pp (2012).
139
Y. F. Gong, P. Wu, W. F. Liu, S. Y. Wang, G. Y. Liu, and G. H. Rao,
Switchable Ferroelectric Diode Eect and Piezoelectric Properties of
Bi0.9La0.1FeO3 Ceramics, Chin. Phys. Lett., 29 [4] 047701, 4pp (2012).
140
D. Damjanovic, Ferroelectric, Dielectric and Piezoelectric Properties of
Ferroelectric Thin Films and Ceramics, Rep. Prog. Phys., 61 [9] 1267324
(1998).
141
K. Uchino, Ferroelectric Devices. Marcel Deker, New York, USA, 2000.
142
T. Tsurumi, Y. Kumano, N. Ohashi, T. Takenaka, and O. Fukunaga,
o
90 Domain Reorientation and Electric-Field-Induced Strain of Tetragonal
Lead Zirconate Titanate Ceramics, Jpn. J. Appl. Phys., 36 [1] 59705
(1997).
143
R. J. Zeches, M. D. Rossell, J. X. Zhang, A. J. Hatt, Q. He, C.-H. Yang,
A. Kumar, C. H. Wang, A. Melville, C. Adamo, G. Sheng, Y.-H. Chu, J. F.
Ihlefeld, R. Erni, C. Ederer, V. Gopalan, L. Q. Chen, D. G. Schlom, N. A.
Spaldin, L. W. Martin, and R. Ramesh, A Strain-Driven Morphotropic
Phase Boundary in BiFeO3, Science, 326 [5955] 97780 (2009).
144
J. X. Zhang, B. Xiang, Q. He, J. Seidel, R. J. Zeches, P. Yu, S. Y. Yang,
C.H. Wang, Y. H. Chu, L. W. Martin, A. M. Minor, and R. Ramesh, Large
Field-Induced Strains in a Lead-Free Piezoelectric Material, Nat. Nanotechnol., 6 [2] 98102 (2011).
145
R. K. Vasudevan, M. B. Okatan, Y. Y. Liu, S. Jesse, J.-C. Yang, W.-I.
Liang, Y.-H. Chu, J. Y. Li, S. V. Kalinin, and V. Nagarajan, Unraveling the
Origins of Electromechanical Response in Mixed-Phase Bismuth Ferrite,
Phys. Rev. B, 88 [2] 020402, 7pp (2013).
146
J. E. Daniels, A. Pramanick, and J. L. Jones, Time-Resolved Characterization of Ferroelectrics Using High-Energy X-Ray Diraction, IEEE Trans.
UFFC, 56 [8] 153945 (2009).
147
A. Pramanick, J. E. Daniels, and J. L. Jones, Subcoercive Cyclic Electrical Loading of Lead Zirconate Titanate Ceramics II: Time-Resolved X-Ray
Diraction, J. Am. Ceram. Soc., 92 [10] 230010 (2009).
148
J. L. Jones, E. B. Slamovich, and K. J. Bowman, Domain Texture
Distributions in Tetragonal Lead Zirconate Titanate by X-Ray and Neutron
Diraction, J. Appl. Phys., 97 [3] 034113, 6pp (2005).
149
A. Pramanick, D. Damjanovic, J. E. Daniels, J. C. Nino, and J. L. Jones,
Origins Of Electro-Mechanical Coupling in Polycrystalline Ferroelectrics
During Subcoercive Electrical Loading, J. Am. Ceram. Soc., 94 [2] 293309
(2011).
150
J. L. Jones, M. Homan, and K. J. Bowman, Saturated Domain Switching Textures and Strains in Ferroelastic Ceramics, J. Appl. Phys., 98 [2]
024115, 6pp (2005).
151
Q. H. Jiang, C. W. Nan, and Z. J. Shen, Synthesis and Properties of
Multiferroic La-Modied BiFeO3 Ceramics, J. Am. Ceram. Soc., 89 [7] 2123
7 (2006).
152
V. V. Shvartsman, W. Kleemann, R. Haumont, and J. Kreisel, Large
Bulk Polarization and Regular Domain Structure in Ceramic BiFeO3, Appl.
Phys. Lett., 90 [17] 172115, 3pp (2007).

2010

Journal of the American Ceramic SocietyRojac et al.

Vol. 97, No. 7

153
Y. Yao, B. Ploss, C. L. Mak, and K. H. Wong, Pyroelectric Properties
of BiFeO3 Ceramics Prepared By a Modied Solid-State-Reaction Method,
Appl. Phys. A, 99 [1] 2116 (2010).
154
D. Damjanovic and M. Demartin, Contribution of the Irreversible Displacement of Domain Walls to the Piezoelectric Eect in Barium Titanate and
Lead Zirconate Titanate Ceramics, J. Phys. Condens. Mat., 9 [23] 494353
(1997).
155
S. Li, W. Cao, and L. E. Cross, The Extrinsic Nature of Nonlinear
Behavior Observed in Lead Zirconate Titanate Ferroelectric Ceramic, J.
Appl. Phys., 69 [10] 721924 (1991).
156
D. Damjanovic, Stress and Frequency Dependence of the Direct Piezoelectric Eect in Ferroelectric Ceramics, J. Appl. Phys., 82 [4] 178897 (1997).

157
R. E. Eitel, T. R. Shrout, and C. A. Randall, Nonlinear Contributions
to the Dielectric Permittivity and Converse Piezoelectric Coecient in Piezoelectric Ceramics, J. Appl. Phys., 99 [12] 124110, 7pp (2006).
158
J. E. Garca, R. Perez, D. A. Ochoa, A. Albareda, M. H. Lente, and J.
A. Eiras, Evaluation of Domain Wall Motion in Lead Zirconate Titanate
Ceramics by Nonlinear Response Measurements, J. Appl. Phys., 103 [5]
054108, 8pp (2008).
159
D. Damjanovic, M. Demartin Maeder, P. Duran Martin, C. Voisard, and
N. Setter, MaxwellWagner Piezoelectric Relaxation in Ferroelectric Heterostructures, J. Appl. Phys., 90 [11] 570812 (2001).
h

Tadej Rojac received his B.Sc. in


2003 in Chemical Engineering and
a Ph.D. degree in 2007 in the
same eld from the University of
Ljubljana, Slovenia. In 2009 he
performed a one-year postdoctoral
study at the Swiss Federal Institute of Technology in Lausanne,
Switzerland. Since 2000 he has
been working as a researcher at
the Electronic Ceramics Department of the Jozef Stefan Institute
in Ljubljana, Slovenia. His main
research interests cover mechanochemical reaction mechanisms, application of the mechanochemical processing in the
synthesis of complex ceramic oxides, and processing-structure-properties relationship in lead-based and lead-free piezoelectric ceramics and thick lms. From 2013 he is an Asst.
Prof. at the Jozef Stefan International Postgraduate School.
He is author and co-author of 25 scientic papers, 3 review
papers, 3 chapters in monographs and 1 patent.

thin lms, tunable ferroelectric thin lms and solutionderived materials for transparent electronics. She is also
involved in studies related to chemical solution deposition of
thin lms, inkjet printing of solutions and particle dispersions. She is author or co-author of more than 130 papers,
10 book-chapters, more than 110 technical reports and 4 Slovenian patents.

Andreja Bencan obtained her B.Sc.


in 1998 in Chemical technology
and in 2002 a Ph.D. in the eld of
Materials, from the Faculty of
Chemistry and Chemical Technology, University of Ljubljana,
Slovenia. She
is
a senior
researcher at the Electronic
Ceramics Department of the Jozef
Stefan Institute, Ljubljana. Her
main research interest lies in the
development of lead-based and
lead-free piezoceramics, especially
in the microstructure investigations by dierent scanning and
transmission electron microscopy methods. From 2009 she is
also habilitated as an Asst. Prof. at the Jozef Stefan International Postgraduate School. She is co-author of about 60 scientic papers and 4 chapters in monographs.
Barbara Malic received her Ph.D.
degree in 1995 in the eld of
chemistry at the University of
Ljubljana, Slovenia. She is currently the head of the Electronic
Ceramics Department at the Jozef
Stefan Institute, Ljubljana, Slovenia, and an associate professor of
chemistry of materials at the Jozef
Stefan International Postgraduate
School. Her research topics cover
lead-based and lead-free ferroelectric and piezoelectric ceramics and

Goknur Tutuncu is currently a


Postdoctoral Research Associate
at Brookhaven National Laboratory (BNL) and before joining
BNL she spent 2 years as a Postdoctoral researcher at the University of Florida. She received her
M.S. degree in Chemical Engineering in 2002 from the Middle East
Technical University and her
Ph.D. degree in Materials Science
and Engineering in 2010 from
Iowa State University. Her primary research focuses and expertises are structure/property
correlations in piezoelectric materials, processing and characterizations of lead-free ferroelectric materials utilizing sophisticated X-ray and neutron diraction techniques.
Jacob Jones is an Associate Professor of Materials Science and
Engineering at North Carolina
State University and Director of
the Analytical Instrumentation
Facility (AIF). Jones received his
PhD from Purdue University in
Materials Engineering in 2004.
From 2006-2013, Jones was at the
University of Florida where he
was an Assistant and then Associate Professor of Materials Science
and Engineering. Jones is an
experimental materials scientist with research interests in Xray and neutron scattering, crystallography, ceramic materials, and mechanical behavior of materials. He has published
over 100 publications on these topics since 2004. Jones has
received numerous research awards including the National
Science Foundation CAREER award, a Presidential Early
Career Award for Scientists and Engineers, the IEEE Ferroelectrics Young Investigator Award, a National Nuclear
Security Administration Defense Program Award of
Excellence, the HHMI Distinguished Mentor (DM) Award
for undergraduate research mentoring, an International
Educator award, and twice received the Edward C. Henry
Award by the Electronics Division of the American Ceramic
Society.

July 2014

BiFeO3 Ceramics: Processing and Properties

John Daniels is currently in the


School of Materials Science and
Engineering at UNSW as a Senior
Lecturer and Australian Institute
of Nuclear Science and Engineering research fellow. He was awarded his PhD in 2007 from the
School of Physics at Monash University, Melbourne, for work in
the eld of time-resolved neutron
scattering in ferroelectric materials. After his PhD, he spent three
years as a postdoctoral researcher
within the Structure of Materials group at the European
Synchrotron Radiation Facility, Grenoble, France. His current research interests are in the application of advanced xray and neutron scattering techniques to the study of functional and mechanical properties of electro-ceramic materials.
Dragan Damjanovic received BSc
diploma in Physics from the University of Sarajevo in 1980, and
PhD in Ceramics Science from the
Pennsylvania State University
(PSU) in 1987. From 1988 to 1991
he was a research associate in the
Materials Research Laboratory at
the PSU. He joined the Ceramics
Laboratory, Institute of Materials,
at the Swiss Federal Institute of
Technology in Lausanne (EPFL)
in 1991. He is currently a professeur titulaire and teaches undergraduate and graduate courses on electrical properties of materials. He investigates
experimentally physical processes taking place at dierent
driving elds and time scales and how they aect macro-

2011

scopic behavior of ceramics, single crystals and thin layers.


His interests include interaction of defects with domain walls,
symmetry breaking and its eect on electro-mechanical and
electro-thermal coupling, interface dynamics, dispersion,
creep, nonlinearity and hysteresis in dielectric, mechanical
and piezoelectric responses, phase transition-related instabilities, structure/microstructure property relations, and applications of dielectric, piezoelectric and ferroelectric crystals,
lms, and ceramics. He is an IEEE Fellow, was awarded
2007 Outstanding Achievement Award by the International
Symposium on Integrated Ferroelectrics and 2009 Ferroelectrics Recognition Award by the IEEE Ultrasonics, Ferroelectrics and Frequency Control (UFFC) Society, was
Distinguished Lecturer for the IEEE UFFC Society for
2010/11, and won with his colleagues the 2012 Edward C.
Henry Best Paper Award by the Electronics Division of the
American Ceramic Society. He authored and co-authored
more than 190 papers.

Das könnte Ihnen auch gefallen