You are on page 1of 23

www.advmat.

de

REVIEW

Physics and Applications of Bismuth Ferrite


By Gustau Catalan* and James F. Scott*

BiFeO3 is perhaps the only material that is both magnetic and a strong ferroelectric at room temperature. As a result, it has had an impact on the eld of multiferroics that is comparable to that of yttrium barium copper oxide (YBCO) on superconductors, with hundreds of publications devoted to it in the past few years. In this Review, we try to summarize both the basic physics and unresolved aspects of BiFeO3 (which are still being discovered with several new phase transitions reported in the past few months) and device applications, which center on spintronics and memory devices that can be addressed both electrically and magnetically.

1. Introduction
1.1. History The basic idea that crystals could be simultaneously ferromagnetic and ferroelectric probably originated with Pierre Curie in the 19th century.[1] After switching was discovered in ferroelectric Rochelle Salt by Valasek in 1920[2] there was a rash of supposed discoveries of magnetoelectric properties by Perrier,[3,4] but unfortunately in materials such as Ni in which they are now understood to be impossible. A history of this period of solid-state physics is given in ODells text.[5] True magnetoelectricity dened as a linear term in the free energy G(P,M,T) aij Pi Mj , where P is the polarization and M is the magnetization was rst understood theoretically by Dzyaloshinskii[6] with special predictions being made for Cr2O3 and discovered experimentally in that material by Astrov.[7] However this material is paraelectric and antiferromagnetic, making microelectronics applications impractical. The more interesting case of ferromagnetic ferroelectrics waited for some years until the work of Schmid on boracites.[8] The boracites are also impractical materials for device applications: they have low symmetry with large unit cells and grow in needle shapes; more importantly, they exhibit magnetoelectricity only at extremely low temperatures. Meanwhile Smolenskiis group in Leningrad pioneered[9] the study of bismuth ferrite, BiFeO3, but they found that they could not grow single crystals and that ceramic specimens were too highly conducting (probably caused by oxygen vacancies and mixed Fe valences) to be used in applications.[10] They tried to address the conductivity problem by
[*] Dr. G. Catalan, Prof. J. F. Scott Department of Earth Sciences University of Cambridge Downing Street, Cambridge CB2 3EQ (United Kingdom) E-mail: gcat05@esc.cam.ac.uk; jsco99@esc.cam.ac.uk

doping other ions into both the A and B sites of the lattice, but no practical devices were obtained. Reviews of the general study of magnetoelectricity appeared by Schmid in 1994[11] and more recently by Fiebig[12] and by Eerenstein et al.[13] The current interest in bismuth ferrite was stimulated primarily by a 2003 paper from Rameshs group,[14] which showed that it had unexpectedly large remnant polarization, Pr, 15 times larger than previously seen in bulk, together with very large ferromagnetism of ca. 1.0 Bohr magneton (mB) per unit cell. Single crystals grown more recently in France in 20067 have conrmed the large value of the polarization rst observed in the lms, showing also that it is intrinsic;[15-19] on the other hand, the intrinsic magnetization of thin lms is now thought to be near zero[20, 21] ca. 0.02 magnetons/cell and possible reasons for discrepancies between the magnetization values encountered in the literature are discussed later in this review. At any rate, the 2003 Science paper has proved enormously stimulating, and has inspired both new fundamental physics and exciting device applications.

1.2. The Hypothesis of Spaldin (Hill) In parallel with the specic investigation of bismuth ferrite and related compounds has been a more general approach to the idea of multiferroics. Nicola Hill (now Spaldin) has asked[22] why there are so few materials that are magnetic and ferroelectric; implicitly limiting her discussion to transition-metal oxides, especially perovskites, she observed that the ferroelectrics (e.g., titanates) have B-site ions with d8 electrons,[23] whereas the magnets require dj electrons with j different from zero. This line of thinking has also proved very stimulating, although it is helpful to remind ourselves that there are many potential multiferroics that are not oxides, as further discussed in Section 1.3. On the other hand, oxide perovskites do not all have the same mechanism of ferroelectricity: the center Ti ion plays the key role in BaTiO3 but the lone-pair Pb ion is dominant in PbTiO3.[24] Indeed, this seems to be the case in BiFeO3, where the polarization is mostly caused by the lone pair (s2 orbital) of Bi3, so that the polarization comes mostly from the A site while the magnetization comes from the B site (Fe3); this same idea has led Spaldin and co-workers to propose a host of other perovskites with possible A-site ferroelectricity and B-site magnetism, such as Bi(Cr,Fe)O3 and BiMnO3. Her work has also been instrumental in triggering the quest for other ways of achieving coexistence of ferroelectricity and magnetism in oxides;

DOI: 10.1002/adma.200802849

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2463

www.advmat.de

Gustau Catalan studied Physics at the University of Barcelona and earned his PhD at the Queens University of Belfast, followed by postdoctoral appointments in CSIC (Spain) and the Zernike Institute for Advanced Materials (Holland). He is currently a Senior Research Associate in the Ferroics Laboratory of the University of Cambridge, which he joined in 2005. His areas of interest include the properties of perovskite oxides and the study of how small size affects the phase transitions and functionality of thin lms and nanocrystals.

James F. Scott was educated at Harvard (B.A., Physics 1963) and Ohio State University (Ph.D., Physics 1966). After six years in the Quantum Electronics Research Department at Bell Labs he was appointed professor of Physics at Univ. Colorado (Boulder). He was Dean of Science and Professor of Physics for eight years in Australia (UNSW, Sydney, and RMIT, Melbourne) and has been Professor of Ferroics in the Earth Sciences Department at the University of Cambridge since 1999. among the new ndings are ferroelectricity induced by spiral spin order,[25,26] magnetic exchange striction,[27] or charge order.[28] Based on these ideas, new oxides have been predicted to be multiferroics, some of which await direct experimental verication.[26,29,30] In addition, there are materials in which the ferroelectricity itself causes spin canting[31,32] and these should be explored further. 1.3. Oxide vs. Fluoride Multiferroics There are hundreds of different crystals with ferroelectric transition temperature, Tc, above room temperature at atmospheric pressure, with approximately a log-normal distribution of transition temperatures.[33] Many of these materials are not oxides, yet almost all work on multiferroics has emphasized oxides. This is convenient as they are easy to grow, particularly in thin-lm form, and researchers in high-Tc superconductors, colossal magnetoresistance manganites, and ceramics all study

oxides. However, there is no a priori reason to expect that they will have the most interesting multiferroic physics or will make the best devices. The exclusive emphasis on oxides seems unwise, and magnetoelectric uorides should probably receive more attention in this context. Abrahams has given a list of magnetic materials that are probably ferroelectric.[3436] His criterion was the existence of a structure that has very small acentric displacements of ions, with the assumption that ferroelectric switching is likely in such lattices. Most of these predicted ferroelectrics are indeed oxides, but some are uorides.[37] BaMnF4, for example, is the earliest known example of a material with spin canting induced by ferroelectricity,[28] the physics being the same as in the ferroelectrically induced local spin canting in BiFeO3, discussed later in this Review. K3Fe5F15[3840] is a ferroelectric ferrimagnet with 2 Fe3 ions and 3 Fe2 ions per unit cell, adding to a weak net ferromagnetic moment. A particularly interesting multiferroic is Sr3(FeF6)2, which merits further study.[41] An additional family of multiferroics that seems promising is Pb5Cr3F19.[42,43] There are many other uoride multiferroics, and (NH4)3FeF6 has received careful study.[44,45] These systems include both pure uorides and oxyuorides,[45] with the latter including Bi2TiO4F2; Ravez has given a review encompassing both uorides and oxyuorides,[46] and Nenert and Palstra[29] have also recently reviewed other possible multiferroic uorides.

REVIEW

2. The Phase Diagram of BiFeO3


2.1. Phase Decomposition and Impurities The phase diagram for the system Bi2O3/Fe2O3 has been mapped out[47,48,49] and is shown in Figure 1. BiFeO3 is usually prepared from equal parts of Bi2O3 and Fe2O3, and under high temperatures it can decompose back into these starting materials, as shown in Equation 1

2BiFeO3 ! Fe2 O3 Bi2 O3

(1)

Figure 1. Compositional phase diagram of BiFeO3. Reproduced with permission from [49]. Copyright 2008, American Physical Society.

2464

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

platinum crucibles which, given the high mutual reactivity of Bi and Pt, may not be the most suitable receptacle. The above discussion underlines what is currently one of the most important difculties in implementing BiFeO3 for practical applications, namely, its compositional instability, with associated ckleness of functional behavior. Addressing this problem is essential if BiFeO3 is to succeed as a technologically relevant material.

REVIEW

2.2. Crystal Structure


Figure 2. Impurities have a strong effect on functional properties. In this gure, the remnant magnetization, M, of BiFeO3 thin lms grown in oxygen-decient conditions is shown to be directly correlated to the amount of g-Fe2O3 parasitic phase as extracted from X-ray diffraction (XRD) peaks. Figure courtesy of Manuel Bibes (Thales-CNRS).

Bismuth ferrite is very prone to show parasitic phases that tend to nucleate at grain boundaries and impurities.[50] It has been argued that BiFeO3 is in fact metastable in air, with optically visible impurity spots appearing well below the melting temperature.[49,51,52] Impurities and oxygen vacancies are also important for thin lms, because they are known to articially enhance the remnant magnetization[19,21] (Fig. 2). Minimizing them requires very careful tuning of growth parameters, particularly oxygen pressure.[21] At room temperature under applied elds of ca. 200 kV cm1 (typical switching voltages across thin lms), BiFeO3 decomposes, yielding magnetite Fe3O4 as a by-product.[53] This was somewhat surprising and is thought to occur via the following reaction

6BiFeO3 ! 2Fe3 O4 3Bi2 O3 O

(2)

The magnetite phase was unambiguously identied by means of micro-Raman studies; the Raman spectra of Fe3O4 is quite distinct and unlike those of Fe2O3. However, the Bi2O3 was not detected, possibly because it is a well-known glass-forming compound, or perhaps because of its evaporation during thermal decomposition. Bi2O3 melts at a temperature slightly above 800 8C.[49] Similar phenomena occur in the electrical stressing of lead zirconate titanate (PZT),[54] which decomposes into rutile TiO2 (but not anatase), and both a-PbO and b-PbO (litharge and massicot). In both bismuth ferrite and PZT these observations raise concerns about the lifetime of ferroelectric memories made from them. In the case of BiFeO3, this decomposition mechanism also provides another possible explanation for the appearance of remnant magnetization in thin lms: it may come from localized spots of magnetite in the sample. We also point out that Bi2O3 and Pt are known to react easily and exothermically with each-other.[55,56] Bismuth has a low melting temperature of 270 8C, at which it readily forms an eutectic alloy with Bi2Pt and, at higher temperatures, other BiPt alloys are also formed.[57] Therefore, it is probably best NOT to use Pt electrodes when probing the high-temperature properties of bismuth-based materials,[58] including bismuth ferrite. We also note that most BiFeO3 ceramics and single crystals are made in

The room-temperature phase of BiFeO3 is classed as rhombohedral (point group R3c).[59] The perovskite-type unit cell has a lattice parameter, arh, of 3.965 A and a rhombohedral angle, arh, of ca. 89.389.48 at room temperature,[60,61] with ferroelectric polarization along [111]pseudocubic.[61] The unit cell can also be described in a hexagonal frame of reference, with the hexagonal c-axis parallel to the diagonals of the perovskite cube, i.e., [001]hexagonal jj [111]pseudocubic. The hexagonal lattice parameters are ahex 5.58 A and chex 13.90 A.[6062] The coefcient of thermal expansion is neither completely linear nor isotropic,[6264] and reported values[62,63] differ notably, ranging from ca. 6.5 106 to ca. 13 106 K1. A very important structural parameter is the rotation angle of the oxygen octahedra. This angle would be 08 for a cubic perovskite with perfectly matched ionic sizes. A measure of how well the ions t into a perovskite unit cell is the ratio rBi r0 =l, where r is the ionic radius of the respective ion and l is the length of the octahedral edge. This is completely analogous to the commonly used Goldschmid tolerance factor,[65] which is dened p as t rBi rO 2rFe rO . For BiFeO3 we obtain t 0.88 using the ionic radii of Shannon,[66] with Bi3 in eightfold coordination (the value for 12-fold coordination is not reported) and Fe3 in sixfold coordination and high spin. When this ratio is smaller than one, the oxygen octahedra must buckle in order to t into a cell that is too small. For BiFeO3, v is ca. 11148 around the polar [111] axis,[59,61,67] with the directly related FeOFe angle, u ca. 1541568.[61,64] The FeOFe angle is important because it controls both the magnetic exchange and orbital overlap between Fe and O, and as such it determines the magnetic ordering temperature and the conductivity, as will be discussed in later sections

2.3. Symmetry of the High-Temperature b and g Phases At approximately 825 8C there is a rst-order transition to a high-temperature b phase that is accompanied by a sudden volume contraction.[49,68] The transition is also accompanied by a peak in the dielectric constant;[68,69] this has been taken as an indication of a ferroelectricparaelectric transition, although dielectric peaks can also occur in ferroelectricferroelectric transitions, such as the orthorhombicrhombohedral transition in the archetypal perovskite ferroelectric BaTiO3 (which is also rst order). Nevertheless, although there is disagreement about the exact symmetry of the b phase above 825 8C, most reports agree that it is centrosymmetric,[7075] so it is probably a safe bet

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2465

www.advmat.de

that the ab transition at TC 825 8C is indeed the ferroelectricparaelectric transition. Palai et al.[49] propose that the symmetry of the b phase is orthorhombic, although their data does not allow establishing the exact space group with certainty. Some authors have argued that the b phase may be tetragonal or pseudotetragonal,[67,71] but that is impossible, since the domain structure rules out a tetragonal symmetry and the perovskite a,b,c lattice constants are each quite different.[49,72] It was also proposed that this phase may instead be monoclinic;[71,72] the measured monoclinic angle was nevertheless initially quoted as 90o within experimental error,[71] so that the b phase was in effect metrically orthorhombic (i.e., the angles may be 908, but internal ion positions in each unit cell do not satisfy orthorhombic constraints). More recently, however, Haumont et al. have quoted a monoclinic angle of 90.018.[72] On the other hand, our group does not see in our specimens the extra XRD lines used to infer monoclinic structure. Also, the domains studied optically do not reveal the many extra wall orientations that would exist if the symmetry were monoclinic instead of orthorhombic. Furthermore, we nd that the bg phase transition to the cubic metallic phase encountered at 1204 K (atmospheric pressure) seems second-order, and a cubicmonoclinic second-order phase transition would violate the principle of maximal subgroup.[76] This is an additional argument in favor of the b phase being orthorhombic: cubicorthorhombic transitions are allowed to be second order and still satisfy the maximal subgroup criterion. A very recent work has added more fuel to the re regarding this problem. Selbach et al.[73] claim that the paraelectric b phase may neither be orthorhombic nor monoclinic, but rhombohedral (space group R3c). It is, however, hard to reconcile this claim with

the splitting of the pseudocubic lattice parameters observed by other groups,[49,71,72,75] or with the symmetry of the domain walls in this phase. Perhaps, more importantly, if both the a and b phases belong to the same crystal class (rhombohedral), then the transition cannot be ferroelastic, which would appear to contradict the observed change in the ferroelastic domain conguration at this transition.[49] Part of the disagreements between all the above works can perhaps be justied by the fact that X-rays are not particularly sensitive to the position of the oxygen ions as a result of the low electronic density of O compared with the Bi and Fe ions. In this respect, neutron diffraction is a far more helpful technique. High-temperature neutron-diffraction experiments have recently been undertaken by Arnold et al.,[74] who show that the b phase is orthorhombic Pbnm, which is the same non-polar orthorhombic symmetry of the GdFeO3 orthoferrite family. In retrospect, of course, this seems quite obvious: once the distinctive feature of BiFeO3 (its ferroelectric polarization) is removed, one may expect this material to be just like all the other perovskite orthoferrites, i.e., it should be orthorhombic. This is, of course, just an ad hoc argument, but apart from the neutron diffraction and our XRD experiments, there may be additional indirect support for an orthorhombic b phase based on chemical-doping experiments with ions other than Bi, as discussed in the next Section. It is worth mentioning as well that the neutron-diffraction experiments of Arnold et al.[74] suggest phase coexistence in the b phase and, indeed, optical viewgraphs at high temperature do show a coexistence of rhombohedral and orthorhombic domains.[49] This is consistent with the strongly rst-order nature of this transition, and perhaps the mixture of rhombohedral and orthorhombic phases could also explain why the b phase may seem monoclinic in some experiments.[71,72] One nal note relevant to this issue is that differential thermal analysis (DTA) measurements (Fig. 3) show a smaller anomaly ca. 30 8C below the transition to the orthothombic b phase. This suggests the existence of an intermediate phase, which may either be the monoclinic phase seen by Haumont and co-workers, or a region of phase coexistence as seen by Arnold et al. and us. As for the symmetry of the highesttemperature g phase, Redferns XRD data[75] show that the most intense Bragg peak, (110)pseudocubic, has a large splitting at room temperature and atmospheric pressure, but is unsplit (resolution 0.058 in 2u) in the g phase above 931 8C (Fig. 3), with the proposed symmetry for the cubic phase being Pm3m.[49,75] Unfortunately, BiFeO3 is very unstable at the high temperature of the bg Figure 3. Specic-heat measurements by a) Kaczmarek et al. [124] and b) Palai et al. [49] suggest transition and it rapidly decomposes into that the transition to the orthorhombic b phase (815 8C in Kaczmareks measurement, 825 8C in parasitic phases such as Bi2Fe4O9 or Fe2O3 ours) may proceed via an intermediate phase which nucleates at ca. 2535 8C below. This may be (Bi2O3 becomes a liquid at that temperature so the monoclinic phase reported by Haumont and co-workers, or else it may be a region of phase it does not show up in the diffraction scans). coexistence between rhombohedral and orthorhombic. c) At temperatures above 930 8C the Accordingly, measurements of the bg transi(110) diffraction peaks appear unsplit (the small split in the diffractogram is due to a1/a2 radiation), indicating that the g phase is cubic. Reproduced with permission from [124]. tion at 930 8C have to be performed on very high-quality samples (preferably single Copyright 1974, Elsevier (a). Reproduced with permission from [49]. Copyright 2008, American crystals) and using very fast heating/cooling Physics Society (b).

REVIEW
2466

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

Figure 4. Properties of Bi1-xLaxFeO3 as a function of La doping. There are indications of at least three phase transitions as x increases. Reproduced with permission from [68]. Copyright 1974, Wiley.

ramps; the different measurement protocols used by different groups mean that not all of them have been able to measure the cubic phase at high temperature. On the other hand, a tendency towards cubic symmetry has also been reported for BiFeO3 as a function of decreasing grain size,[77] which indirectly supports the conclusion that the highest-symmetry phase is cubic.

2.4. Phase Transitions with Pressure or with La Doping The diffraction peaks are also unsplit in Redferns measurements at room temperature and high pressure above 47 GPa (although the experimental resolution is low, with 0.588 peak broadening); this suggests the same cubic symmetry at high pressure as observed at high temperature, in agreement with the early phase diagram of Scott et al.[84] Gavriliuk et al. report a rhombohedral structure instead,[78] in agreement with ab initio simulations.[79] The experimental resolution precludes a completely unambiguous answer at this stage, as very small splittings may have been

masked by broadening in our measurements or, conversely, lack of hydrostatic equilibrium may cause peak asymmetry that can be wrongly interpreted as peak splitting in Gavriliuks. As a side comment, we note that the question of whether materials tend towards cubic symmetry with high pressure is, surprisingly, unresolved even for simple elements.[80] Recently, Pashkin et al. [81] have reported additional phase transitions at room temperature at ca. 35 and 7.510 GPa. The reported pressure-induced transition near 10 GPa is to an orthorhombic Pnma (Pbnm) state. This could also be the orthorhombic symmetry for the high-temperature b phase proposed by Palai et al.[49] and Arnold et al.[74] While these low-pressure transitions have not been conrmed by highpressure studies in the USA,[82] or Russia,[82,83] the transition to the orthothombic b-phase near 10 GPa has been recently reproduced by Redfern et al.[75] It thus seems that the rhombohedral-orthorhombic-cubic sequence of phase transitions is the same as a function of pressure as it is as a function of temperature. An alternative way of inducing pressure in a crystal is by chemical substitution of an ion for another of the same valence but different size what is sometimes called chemical pressure. The most common isovalent substitute in BiFeO3 is La3 for Bi3. However, interpretation of the effects of La doping in terms of chemical pressure is not straightforward because La3 has almost exactly the same ionic radius as Bi3 (1.16 and 1.17 A, respectively[66]). Furthermore, the lone-pair orbital of Bi3 (6s2) is stereochemically active and responsible for the ferroelectric distortion; distortions induced by La doping are therefore more likely to be caused by the turning off of the lone-pair activity (i.e., the turning off of the ferroelectricity) than to direct differences in ionic size. The rst phase diagram for Bi1xLaxFeO3 was published by Polomska et al.,[68,85] who looked at the dielectric constant and volume expansion as a function of La doping concentration (Fig. 4). In their study, there is a rst-order transition with sharp volume contraction for x ca. 0.2 (Fig. 4) and several other transitions, the last one of which is at x ca. 0.75 to the orthorhombic Pnma (centric) phase of pure LaFeO3, also reported for pure BiFeO3 above ca. 10 GPa[81] at room temperature or above 825 8C at ambient pressure.[74] The nature of the intermediate bridging phases, however, is unclear. Gabbasova et al.[86] and Zalesskii et al.[87] claim a noncentric orthorhombic phase (C222) for 0.2 < x < 0.6 that could also be associated with the hightemperature b phase of pure BiFeO3 the sudden volume contraction at x 0.2 is in this context very reminiscent of the volume contraction observed at the temperature-induced ab transition (825 8C). At any rate, the exact nature and even the number of structural phase transitions as a function of La doping is still an open question.[63,8688]

REVIEW

2.5. Other Anomalies above Room Temperature: Phase Transitions vs. Defects Krainik et al.[51] measured the GHz dielectric constant and thermal expansion of BiFeO3 between room temperature and 900 8C. They found small anomalies at 130, 200, 280, 370, 460, 600, 670, 740, and 845 8C. However, the authors themselves mention that the samples are almost phase pure, which is

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2467

www.advmat.de

another way of saying that they are not pure; accordingly, some of their anomalies could be due to parasitic phases and defects (see Section 2.1), particularly since many have never been reproduced in later measurements. Nonetheless, some of these anomalies are clearly correlated with known phase transitions: the 845 8C peak is almost certainly the ab phase transition, whereas the anomaly at 370 8C, also reported by Polomska,[68] is caused by magnetoelectric coupling to the antiferromagnetic Neel temperature (TNeel). With the exception of the anomaly at the Neel temperature, none of the other possible phase transitions shows up in the refractive index as a function of temperature.[89] The most intriguing of these ghost transitions is perhaps the anomaly in both dielectric constant and thermal expansion reported by Polomska near 185 8C or 458 K.[68,85] It is possible that the phase transition reported at 458 K and ambient pressure could be the same as that observed at room temperature and ca. 4 GPa.[81] Both sides of this phase boundary were rst reported by Pashkin et al. as rhombohedral[81a] and, hence, the transitionif it were a transition would not be ferroelastic.[90] Other ferroelectrics, such as nickel iodine boracite, have had isomorphic transitions proposed for them which do not change symmetry,[91] so this is not out of the question. On the other hand, the transition at 458 K is not universally observed: our own single-crystal dielectric measurements do not show any clear feature around that temperature. Nor is there a clear signature of that transition in the phonon behavior,[49,70] all of which argues in favor of an extrinsic origin of the anomalies. Having said that, the 458 K dielectric anomaly has been reported by other groups,[93] and also appears in electrical resistivity measurements in our laboratory (Fig. 5). This could, of course, still be related to impurities rather than being intrinsic; comparison between two-probe and four-probe measurements, for example, shows the anomaly to be much stronger in the former, suggesting contact- resistance effects, although the derivative of the four-probe resistivity does still show a peak near 185 8C. However, other reports of resistivity do not show any anomaly near ca. 185 8C.[94,95] While the coincidence of our resistive anomaly with Polomskas (whose impedance and dilatometry measurements were in a completely different set of samples) is tantalizing, at this point the evidence for and against an intrinsic origin seems to be split down the middle, so this possible transition certainly merits

REVIEW

Figure 6. Sketch of a possible phase diagram as a function of pressure and temperature. Solid points are experimental data, the lines are only a visual guide. The ground state is rhombohedral, and the b phase is orthorhombic. The reported monoclinic phase transition [71,72,81] has not been conrmed, and may actually be a coexistence of rhombohedral and orthorhombic (not unusual for a strongly rst-order phase transition). Pressure is known to increase TNeel in orthoferrites at a rate of 4.07.5 K GPa1 [204]. Accordingly, we expect that TNeel will rise slowly with hydrostatic pressure up until the triple point is reached, but there is no direct experimental evidence of this; above the triple point, the pressure-induced metalinsulator (MI) transition (TMI) delocalizes the electrons and induces a Pauli paramagnetic state, so that the magnetic-ordering temperature, TN, is forced to track down TMI The metallic state at high pressures and low temperatures has been claimed by Gavriliuk et al. [78] and Gonzalez Vazquez and Iniguez [79] to be rhombohedral, while the data of Redfern et al. is consistent with cubic[49,75]. This schematic phase diagram does not include any of the new magnetic phase transitions observed at low temperatures and discussed in later Sections.

further careful studies in order to unambiguously establish its nature. Based on the pressure/doping effects and the known behavior of the magnetic transition in orthoferrites, we propose a schematic temperaturepressure phase diagram for BiFeO3 (Fig. 6). At present this is but an informed guess, with enormous gaps in real experimental data, so we very much encourage the careful exploration of this map.

3. Conductivity, Bandgap, and MetalInsulator (MI) Transition


3.1. Resistivity of BiFeO3 The dc resistivity of good-quality bulk samples of BiFeO3 exceeds 1010 Ohm cm.[49,95] As temperature increases, the resistivity decreases as would be expected from any widebandgap semiconductor. Around the TN (370 8C) there is no change in the absolute value of resistivity, but Arrhenius plots show a change in slope (Fig. 7), with the

Figure 5. Resistance as a function of temperature of BiFeO3 single crystals: two-probe (left) and four-probe (right) measurements. The sharp anomaly at 185 oC/458 K in the two-probe measurement is very washed-out in the four-probe resistivity, although the derivative (inset) does show a peak at almost exactly the same temperature. The four-probe measurements were performed by Julia Herrero-Albillos, University of Cambridge.

2468

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

Figure 7. Arrhenius plot of the two-probe resistivity of a single crystal, showing a change of slope at the Neel temperature.

activation energy of the charge carriers decreasing from ca. 1.3 to ca. 0.6 eV as the material is heated above TN. Resistive anomalies at TN have also been reported by Selbach et al.[77] This indicates that magnetic ordering affects the conductivity bandgap, increasing it in the antiferromagnetic phase, which is consistent with ab initio calculations.[92] The correlation between bandgap and magnetic ordering suggests that BiFeO3 could be magnetoresistive. Indirect evidence of this exists from the dielectric measurements of Kamba et al.[96] and direct measurements are currently underway in our laboratory. At even higher temperatures there are further resistive anomalies correlated with the ab (rhombohedralorthorhombic) transition, the bg (orthorhombiccubic) transition and, nally, the decomposition temperature.[49] Specically, the resistivity decreases (but remains semiconducting) at the ab transition[77] and the slope of the resistivity as a function of temperature changes sign[48] at the bg transition, which is consistent with a metalinsulator (MI) transition, as discussed below.

3.2. Bandgap and MI Transition

Reported values for the optical bandgap of BiFeO3 at room temperature range from ca. 2.3 to ca. 2.8 eV.[49,95,97100] According to some authors, this bandgap is direct,[98,99] although other reports suggest also the presence of an indirect bandgap roughly 0.41.0 eV smaller than the direct one.[95,97] Ab initio calculations using screened exchange formalism show that bismuth ferrite is a semiconductor with a room-temperature gap of ca. 2.8 eV.[49,100] The valence-band maximum is at the R-point corner of the Brillouin zone, whereas the conduction-band minimum is at the center, G, so that the gap is indirect. However, the calculated valence band in the rhombohedral state is in fact almost at[100] so that BiFeO3 should in practice behave as a direct-bandgap semiconductor at Figure 8. Optical bandgap of BiFeO3 as a function of pressure and temperature. Figures room temperature. The same calculations, reproduced with permission from [83] and [49], respectively. Copyright 2007, Materials Research however, show an evolution towards indirect Society and 2008, American Physics Society, respectively.

bandgap as BiFeO3 goes from rhombohedral to orthorhombic to cubic, with the indirect bandgap decreasing markedly at each of these transitions. Note that the screened exchange band structure calculation gave good results for the bandgap versus temperature, in comparison with experiment, but it was not a total energy calculation and hence does not assess stability of phases. As temperature increases, Palai et al.[49] have indeed measured that the optical bandgap decreases and goes to zero abruptly at the g-phase, signaling a temperature-driven MI transition. MI transitions are of particular interest in solid-state physics, and are often studied as a function of pressure as well as temperature. An MI phase transition has indeed been observed in bismuth ferrite at room temperature and at a pressure of ca. 50 GPa,[78,82,83] as well as at 1204 K at a pressure of 1 atm.[49] As far as we know, MI transitions have not been reported in perovskite ferrites other than BiFeO3. The evidence for the MI transition at the orthorhombiccubic transition near TMI ( ca.1204 K) is rst, that the optical bandgap goes to zero at that temperature[49] or at room temperature and high pressure[78,82,83] (Fig. 8); second, that the magnetism disappears;[78,101] third, that the temperature derivative of resistivity changes sign;[49,78] and fourth, that the reectivity increases abruptly.[49,78] As the temperature rises, the deviation from cubic structure decreases and the experimentally measured gap[49] decreases to ca. 1.6 eV by 500 8C (723 K, still in the rhombohedral phase). At 1204 K the structure becomes cubic via a second-order transition,[49,75] and the conduction-band minimum now overlaps the valence-band maximum. Thus a semimetal is formed, as in elemental Bi or graphite. Although the behavior is now metallic, the material is not strictly a conventional metal with a half-lled band. The pressure-induced MI transition is correlated with the loss of magnetism.[101] A possible interpretation of this is that the MI transition triggers the magnetic one by delocalizing the magnetic electrons. However, a different interpretation has been proposed by Gavriliuk et al.,[78] who think instead that the order of precedence is different, i.e., the magnetic transition induces the MI change rather than the other way round. These authors suggest that the MI transition may be Mott-type. This means that the bandgap is caused by electron electron Coulombic repulsion (the Hubbard parameter, U), and that there is a critical value of U which can be reached with either temperature or pressure.[104,105]

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2469

REVIEW

www.advmat.de

According to Gavriliuk et al., the change in U would be due to a change in the spin conguration from the ground-state high spin (the ve d-shell electrons occupying one each of the t2g and eg levels, giving a total magnetic moment, S 5/2[102,103]) to low spin (no electrons in the eg level and S 1/2), with the weaker magnetic interactions in the low-spin state being consistent with the observed paramagnetism.[101] While there is no direct experimental evidence for this, ab initio calculations do agree with a low-spin conguration at high pressures and low temperatures.[79] The mechanism proposed by Gavriliuk et al. for the pressure-driven MI transition is unlikely to work for the temperature-driven one. For one thing, BiFeO3 is magnetically disordered both above and below TMI. Furthermore, the transition we observe appears to be second order, which violates one of Motts principal requirements.[104] We think instead that the MI transition is triggered by the structural change, a hypothesis supported by the screened-exchange model,[49] which shows that BiFeO3 is metallic in only the cubic phase. BiFeO3 is viewed as a charge-transfer insulator, with the bandgap controlled by the orbital overlap between the O 2p and the Fe 3d levels.[49,100] The overlap integral in turn depends on the FeOFe exchange angle; accordingly, the observed straightening of the bond angle with increasing temperature[64,75] would result in the observed decrease of bandgap with increasing temperature. This mechanism is in fact completely analogous to that of the MI transition in the perovskite nickelates[106108] (parenthetically we note that perovskite nickelates are also thought to be multiferroic[25,26,29]), with the main difference being that, whereas in the nickelates the bond angle is tuned by the ionic size, in BiFeO3 the FeOFe bond angle is controlled by the ferroelectric distortion.[92,100,109] The correlation between orbital overlap and bandgap is also very relevant for the local properties including conductivity of the domain walls, as further discussed in Section 8. It seems strange to have two different mechanisms depending on whether the MI transition is induced by pressure or temperature. On the other hand, the high-pressure symmetry is reported as rhombohedral[78,79] (although some experiments suggest cubic, as mentioned in Section 2) whereas the high temperature one is cubic.[49,75] So, structurally at least, the two phases may indeed be different and different physical mechanisms for the MI transition could therefore be expected. However, we suggest that there could be a third way, reconciling aspects of the two models. Here we note that the low-spin ionic radius of Fe3 is rLS 0.55 A, whereas that of the high-spin one is rHS 0.645 A;[66] the smaller low-spin radius is, of course, the reason why pressure can induce the transition to low spin in the rst place. The smaller size of the low-spin Fe must necessarily result in a shrinking of the oxygen octahedron around it, meaning that the FeOFe bond angle can straighten: the Goldschmid tolerance factor is 0.88 for the high-spin conguration, and 0.93 for the low-spin one, with the octahedral rotation angle being closer to 0 as it approaches 1. Thus, while the idea of the pressure-induced MI transition associated with a change to low spin may be correct, we also believe that the change in bandgap may not itself be caused by a reduction in the MottHubbard electronelectron repulsion, but by the low-spin-induced straightening of the FeOFe bond angle.

REVIEW

4. Ferroelectricity
4.1. Bulk The ferroelectric polarization of bulk bismuth ferrite is along the diagonals of the perovskite unit cell ([111]pseudocubic/ [001]hexagonal). Early measurements of bulk ferroelectricity in the 1960s and 1970s yielded only small values of the polarization. However, the small value of Pr (ca. 6 mC cm2) reported by Teague et al.[110] for single crystals was viewed by those authors as limited by lack of saturation, and they remarked, presciently, that . . .the actual polarization of BiFeO3 is an order of magnitude higher than we have measured.. It took more than 30 years before they were proved right by measurements on high-quality thin lms,[14] single crystals,[15,16] and ceramics.[111] The unprecedented large polarization of the thin lms was initially thought[14] to be due to strain enhancement, but this is no longer the case: good single crystals were eventually grown[1519] with Pr values very similar (Fig. 9) to those of the lms: ca. 60 mC cm2 normal to (001) and, therefore, approximately 100 mC cm2 along [111]pseudocubic, and high polarization was also found in ceramics.[111] Ab initio calculations also agree with the statement that the polarization of bulk BiFeO3 is intrinsically high[92,109] (ca. 90100 mC cm2) and relatively insensitive to strain.[109] 4.2. Thin Films and Strain Effects In addition to having excellent ferroelectric properties as discussed above, thin lms of BiFeO3 often have different crystallographic structures than single crystals do. Freestanding

Figure 9. Polarization of BiFeO3: bulk single crystal (top) and epitaxial thin lm (bottom). Figures reproduced with permission from [15] and [14], respectively. Copyright 2007, American Institute of Physics (top) and 2003, AAAS (bottom).

2470

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

dendritic lms were prepared as early as the mid 1980s by Hans Schmid and others in Geneva, and these are like single crystals. In particular, their crystal class at ambient temperatures is rhombohedral. However, when bismuth ferrite is epitaxially grown as a thin lm onto, for example, an SrTiO3 [001] substrate, the resulting morphology is monoclinic, where the symmetrylowering distortion arises from in-plane contraction and outof-plane elongation as a result of lattice mismatch between lm and substrate. This has been characterized by several groups.[112] An as yet unresolved issue is whether there is a further change in symmetry, from monoclinic to tetragonal, as lm thickness is reduced. XRD and Raman spectroscopy data suggest that epitaxial BiFeO3 grown on SrTiO3 becomes tetragonal[112,113] below a critical thickness of ca. 100 nm. On the other hand, piezoresponse atomic force microscopy (PFM) studies performed on ultrathin lms[115] still show eight polarization variants, consistent with polarization oriented along the diagonals, as expected from a monoclinic structure, rather than the two variants expected from an in-plane-compressed tetragonal phase. While this discrepancy is still unresolved, we note that both observations are not necessarily incompatible, as the ultrathin lms could perhaps be metrically tetragonal but with the actual point group being monoclinic. That is, while the external shape of the unit cell may be tetragonal, the internal degrees of freedom responsible for the polarization might remain monoclinic. Such a decoupling between crystal class and internal symmetry has been previously reported in other epitaxially strained perovskite thin lms.[116,117] The in-plane compression was initially thought[14,112] to enhance the polarization, a natural assumption given the strong effect of strain on the ferroelectricity of other perovskite lms.[118,119] As discussed in the previous section, however, this is now known not to be the case. Direct experimental proof of the small sensitivity of the polarization to the strain state was recently published by Kim et al.,[114] who show that the polarization of epitaxial BiFeO3 stays constant even as the epitaxial strain is relaxed with increasing lm thickness (Fig. 10). A newer study has also been published by Jang et al.[120] looking closely at the relationship between strain and polarization in BiFeO3; these authors conrm that the spontaneous polarization does not change in absolute magnitude but can be rotated out-of-plane through the monoclinic symmetry plane.[120] The reason for the relatively small sensitivity of BiFeO3 to epitaxial strain is that its piezoelectric constant, which links strain to polarization, is also relatively low (between 1560 pm V1[14,15,111,121] compared with 1001000 pm V1 for other perovskite ferroelectrics). The piezoelectric constant of proper ferroelectrics/improper ferroelastics with a centrosymmetric paraphase can itself be linked to a more fundamental parameter: the electrostrictive coefcient, Q. This relates the strain, s, to the square of the polarization, s QP2. The effective piezoelectric coefcient is dened as the derivative of the strain with respect to the electric eld
eff d33 

REVIEW
Figure 10. The absolute value of the ferroelectric polarization in thin lms (c) is essentially independent from in-plane compression of the lms (a,b). Reproduced with permission from [114]. Copyright 2008, American Institute of Physics.

the effective electrostrictive coefcients are Qeff33 ca. 14 102 m4 C2, with the lower limit being in good agreement with experimental estimates[121]. The full electrostrictive tensor has in fact been calculated by Zhang et al.,[122] who give values of Q1111 0.032 m4 C2, Q1122 0.016 m4 C2, and Q1212 0.01 m4 C2, all of which are in the range estimated here from reported piezoelectric constants. It is interesting to note that these electrostrictive values are similar to those of BaTiO3 or SrTiO3,[123] and yet the strain effect on the ferroelectricity of BaTiO3 and SrTiO3 is much bigger. In other words, although the piezoelectric coefcient of BiFeO3 is small, its electrostrictive coefcient is not. The reason for this seemingly surprising result is likely to be the small dielectric constant (see Section 5), which affects piezoelectricity as given in Equation 3. Possible reasons for the smallness of the dielectric constant are discussed in the next section.

5. Dielectric Properties
5.1. Dielectric Constant from Radio Frequency to Optical Frequency

@s @s @P 2QP" @E @P @E

(3)
The GHz dielectric constant of BiFeO3 at room temperature is er ca. 30.[68,51,96,124] It peaks at the rhombohedralorthorhombic transition (825840 8C), possibly though not necessarily due to a ferroelectricparaelectric transition. This dielectric constant

where e is the dielectric constant. Using the experimentally measured values of d33[14,111,121] we obtain from Equation 3 that

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2471

www.advmat.de

is small compared with those of typical perovskite ferroelectrics such as BaTiO3, (Ba, Sr)TiO3 and Pb(Zr,Ti)O3 (PZT), which, as argued in the previous section, is the reason why the associated piezoelectric coefcient is also smaller. The mean refractive index, n, of BiFeO3 is[89] ca. 2.62, so the optical frequency dielectric constant can be estimated as er n2 ca. 6.86. This is only an average value, however; BiFeO3 is in fact strongly birefringent with Dn ca. 0.34[89] meaning that the dielectric constant at optical frequencies is very anisotropic. Although ca. 30 can be regarded as the intrinsic dielectric constant of this compound at radiofrequencies, the impedance measurements in parallel-plate capacitors often yield higher values: between 50 and 300 depending on sample morphology, orientation, and frequency range. This is because at the frequencies typically accessible by impedance analyzers (100 Hz to 1 MHz), domain-wall motion and space-charge contributions can be important and add to the measured permittivity. While the intrinsic value er ca. 30 may seem small for a ferroelectric, this value is not unreasonable. For one thing, the ferroelectric Curie temperature of BiFeO3 is very high, meaning that at room temperature the ferroelectric polarization is already saturated and, thus, small electric elds will barely affect it (the dielectric constant is essentially a measure of polarizability). Furthermore, this is a strongly rst-order transition to start with, so again there is little phonon softening and thus the dielectric constant predicted by the frequency-shift according to the LyddaneSachsTeller relationship can also be expected to be very low. Finally, and this is just a hypothesis, it may be that perovskite ferroelectrics in which the polarization comes from the A site (e.g., PbTiO3 and BiFeO3) have intrinsically lower dielectric constants than those where polarization comes from the B site (e.g., BaTiO3). Experimentally this certainly seems to be the case, but at present we know of no satisfactory explanation for this fact, if indeed it is more than just a coincidence. At low frequencies or at high temperatures, colossal dielectric constants have also been reported[96,125] and these are clearly due to nite conductivity leading to MaxwellWagner (MW) behavior.[96,125-127] The temperature at which the MW effects set in depends on the sample conductivity; for some samples this effect happens at temperatures as low as 200 K;[96] in our own single crystal and ceramics the nite resistivity effects typically appear above room temperature, enabling a more condent analysis of intrinsic dielectric effects. 5.2. Dielectric Anomalies at Magnetic Transitions Because bismuth ferrite is piezoelectric at all temperatures below 1100 K, any magnetoelastic phenomena at its magnetic-phase transitions are apt to create responses in the dielectric response. These are shown in Figure 11. The subtle low-temperature anomalies at 200 and at 50 K coincide with the temperatures where magnetic, magneto-optic and elastic anomalies have been seen, as discussed in the next section. None of the dielectric anomalies is strong and, curiously, none seems to affect the dielectric loss. Their

weakness shows that they do not correspond to ferroelectric phase transitions, but arise instead from weak coupling to another order parameter, most likely magnetic. Additional dielectric and conductivity anomalies are reported[68] at TNeel 643 K (370 8C), clearly related to magnetoelectric coupling, and magnetodielectric coupling is also responsible for the reported anomaly in the birefringence of BiFeO3 at TNeel.[89] Another is reported at the heretofore mysterious transition at 458 K (185 8C), although this dielectric anomaly may itself be an artifact caused by the change in resistivity.[125128]

REVIEW

6. Magnetism
6.1. Magnetic Symmetry and Spin Cycloid The local short-range magnetic ordering of BiFeO3 is G-type antiferromagnet, that is, each Fe3 spin is surrounded by six antiparallel spins on the nearest Fe neighbors. The spins are in fact not perfectly antiparallel, as there is a weak canting moment caused by the local magnetoelectric coupling to the polarization (see next section). Superimposed on this canting, however, is also a long-range superstructure consisting of an incommensurate spin cycloid of the antiferromagnetically ordered sublattices. The cycloid has a very long repeat distance of ca. 6264 nm, and a propagation vector along the [110] direction.[129,130] The magnetic easy plane (the plane within which the spins rotate) is dened by the propagation vector and the polarization vector (Fig. 12). The magnetic Neel temperature is ca. 643 K (370 8C) and the exponent characterizing the sublattice magnetization as a function of temperature, b, is known to be approximately 0.43 from birefringence[89] and 0.37 from Mossbauer hyperne splittings.[102] Other critical exponents are discussed in the literature.[131,132] The cycloidal model of spin ordering in bismuth ferrite was rst proposed by Sosnowska et al. (1982),[130] whose group has made a number of detailed studies via XRD, neutron scattering, Mossbauer measurements, etc.[130135] However, in recent years Zalesskii and co-workers[136138] have proposed that the simple cycloid is distorted at low temperatures. However, no published data from either group indicate the phase-transition temperature where the spin reorientation transition should occur.

Figure 11. Anomalies in the relative dielectric constant (), possibly due to coupling to magnetic (or magnetoelastic) transitions at low temperature. The anomalies do not seem to affect the dielectric loss (tan d). Adapted from [142], with permission. Copyright 2008, Insitute of Physics.

2472

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

Figure 12. Schematic representation of the spin cycloid. The canted antiferromagnetic spins (blue and green arrows) give rise to a net magnetic moment (purple arrows) that is spacially averaged out to zero due to the cycloidal rotation. The spins are contained within the plane dened by the polarization vector (red) and the cycloidal propagation vector (black). Figure reproduced with permission from [129]. Copyright 2008, American Physical Society.

It is also worth noting that in some single-crystal monodomain samples the cycloid propagates along only one of the three symmetry-equivalent <110> directions. This unique propagation direction is suggestive of a magnetic symmetry-lowering effect (from rhombohedral to monoclinic), as emphasized by Lebeugle et al.[129] and Schmid.[139]. In 20072008 two groups reported evidence for further magnetic phase transitions at 140 and 200 K. Cazayous et al. rst reported a transition at 140 K[141] and, independently, Singh et al. found that one and an apparently stronger one at 200 K.[103] A possible origin of these transitions is discussed below together with evidence for spin-glass behavior.[142,143] 6.2. Spin Reorientation in Orthoferrites

however, that the phenomena at 140 K are different from those at 200 K. The orthoferrites are, as the name implies, orthorhombic, whereas BiFeO3 is crystallographically rhombohedral; however, its local magnetic structure is monoclinic,[146] with a monoclinic angle very near 908, which justies approximating the spin structure as orthorhombic, as in the orthoferrites. We note also that the FeOFe exchange angle (ca.1568), octahedral rotation angle (ca. 128) and Neel temperature (ca.640 K) are all in the same range as those of the rare-earth orthoferrites.[147] On this basis, one may hypothesize that the magnon anomalies observed at 140 and 200 K may be indicative of spin reorientation in BiFeO3 analogous to that observed in orthoferrites such as ErFeO3. On the other hand, the spin reorientation in orthoferrites such as ErFeO3 is thought to be brought about by the magnetic inuence of the rare-earth ions.[148] Clearly this cannot be the case in BiFeO3 as bismuth is not magnetic. So, again, whether or not the phase transitions at 140 and 200 K are indeed due to spin reorientation remains uncertain.

REVIEW

6.3. Spin-Glasslike Behavior The evidence for spin-glass (or, at least, nonergodic) behavior in BiFeO3 is[143,149] rst that there is a large difference between its eld-cooled (FC) and zero-eld-cooled (ZFC) magnetization below ca. 240 K (Fig. 14) (weaker FC effects were also reported by Pradhan et al.[150] and Nakamura et al.[151]); second, that there is a cusp at ca. 50 K in the magnetic susceptibility;[143] and third, that the temperature of the cusp in magnetic ac susceptibility appears to be dependent upon the frequency of the magnetic eld.[143] When magnetic spins are subjected to competing forces and geometric constrains, frustration can result in a chaotic glassy

In the magnetic orthoferrites (e.g., ErFeO3) there are phase transitions within the antiferromagnetic phase at which the sublattice spin orientations rotate. These occur at temperatures (90 and 103 K in ErFeO3) far below the Neel temperature (TN 633 K in ErFeO3) and, hence, have nothing to do with loss of magnetic order. Generally these transitions occur in pairs: at the upper temperature the spins begin to rotate out of plane; and at the lower temperature, the rotation is complete so that the spins are now 908 from their original directions, perpendicular to the plane. These phenomena are well understood in orthoferrites[144] and Raman spectroscopy of magnons near the reorientation temperatures shows frequency dips, cross-section enhancements, and linewidth narrowing.[145] The dip in frequency would be 100% (to zero) if there were no magnetoelastic behavior, but actually reaches 50% in ErFeO3[145] and only 5% in BiFeO3.[103] The cross-section divergences are shown for bismuth ferrite in Figure 13. The linewidth narrowing for the magnons goes from 3.5 to <1.9 cm1 (resolution-limited) at the reorientation transition temperatures. Also Figure 13. (Left) Intensity of magnon peaks in the Raman spectra as a function of temperature. shown in Figure 13 is the EPR susceptibility, These show clear phase transitions at ~140 K and ~200 K, the origin of which is as yet unclear but which is tentatively attributed to spin reorientations. (Above right) Magnon linewidth narrowing which shows discontinuities at both 140 shows "critical slowing down" of spin uctuations near 140 K, proof that the cross section and 200 K; these anomalies conrm the divergence cannot come from impurities. (Below right) Preliminary electron paramagnetic interpretation of these two temperatures as resonance measurements show clear anomalies also at 140 and 200 K. Figures courtesy of M. those of magnetic-phase transitions. Note, Singh (Puerto Rico) and Pavle Cevc (Ljubliana).

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2473

www.advmat.de

section. We remind readers that strain is always unscreened and therefore very long-range and generally mean-eld: if the magnetic-order parameter is coupled to strain, the mean-eld nature of the strain might induce mean-eld behavior in the magnetic state. It is also useful to consider the absolute magnitude of the shift in Tf with f. This is usually dened by a dimensionless sensitivity parameter, K

REVIEW

K DTf =Tf D log f

(5)

Figure 14. Field-cooled (FC) vs. zero-eld-cooled (ZFC) magnetization in single crystals of BiFeO3 for different magnetic elds (H). The strong difference between the two is consistent with a spin-glass state. Figure courtesy of M. Singh, University of Puerto Rico.

state. The original spin-glass model of Kirkpatrick and Sherrington[152] gave detailed predictions for such systems within a mean-eld theory. The spin glasses studied experimentally are centric; that is, their spatially averaged structure has an inversion center. More recent work has generally applied Ising-model statistics to such spin glasses, but bismuth ferrite would be a rare (perhaps unique) case of a spin glass that is ferroelectric and hence non-centrosymmetric (acentric). As Fisher and Hertz have emphasized in their text,[153] no published theories apply to acentric spin glasses, and Ising models denitely cannot apply to them. Spin glasses are characterized by the frequency dependence of the peak in their magnetic susceptibilities. As the frequency is increased, the peak in the ac magnetic susceptibility moves to higher temperatures. If we call Tf the temperature at which the ac susceptibility has a maximum for a measurement frequency f, and TSG the extrapolated value of Tf at f = 0, then the spin-relaxation time, t, varies as:

Typically in a superparamagnetic crystal, 0.01 < K < 0.1; whereas in a conventional spin glass, 0.001 < K < 0.01. In BiFeO3, Singh et al. [143] found K 0.014, which is at the margin between the two cases, and hence they infer that bismuth ferrite is not a conventional spin glass. Further, they note that the magnitude of the ac magnetic susceptibility increases with frequency; which is not reasonable for any glass, since glassy states are always less responsive as frequency increases. They suggest that this might be due to electrical or mechanical resonances in the kHz regime, but more work clearly is warranted to clarify these issues, including what is the origin of the glassy state and whether or not it is an intrinsic feature as opposed to a defect-related phenomenon. An unpublished report was given on glassy behavior in single crystals of BiFeO3 very recently which may be helpful in this regard. Shvartsman et al. conrm re-entrant non-ergodic behavior of the low-eld magnetization at low temperature, but they exclude a generic spin-glass phase, since only cumulative relaxation is found after isothermal aging below Tg instead of classic hole burning and rejuvenation.[155] Other technical details need attention, such as the possible presence in bismuth ferrite of the AlmeidaThouless (AT) transition line, which describes the stability of a spin glass at nite temperatures and magnetic elds.[156] Such an AT line is shown as a function of magnetic eld in Figure 15 for BiFeO3.[149] Although the physical interpretation is not yet clear, it is worth noticing that the extrapolation temperature of the AT line is ca.140 K, which is one of the two critical temperatures of the electromagnon spectra. It does not, however, coincide with the freezing temperature extracted from the ac-susceptibility analysis.

tTf aTSG =Tf TSG zn

(4)

where a is a constant independent of T and the exponent zn is a characteristic spin-glass critical exponent describing the slowing down of spin uctuations near TSG. Although zn 79 for Ising models, there are other systems known to have 1 < zn < 2, as in the present case; La0.5Mn0.5FeO3 is a good example of such a nonstandard spin glass,[154] with zn 1.0. When the susceptibility data are analyzed quantitatively, a spin-glass freezing temperature of 29.4 K is estimated, and the critical exponent zn 1.4 0.2 that characterizes the relaxation dynamics. This critical exponent zn is 79 for Ising models, but 2 in mean-eld theory.[152] This suggests that the spin glass in BiFeO3 may be mean eld; this would be reasonable in view of the strong elastic effects manifest near the magnetic transition temperatures, as discussed in the next

Figure 15. AlmeidaThouless t of the irreversibility temperature determined from the FC vs. ZFC data in Figure 14.

2474

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

It is not easy to prove the existence of a spin-glass state experimentally: superparamagnets can exhibit AT lines, pinned domains can exhibit aging and rejuvenation, and relaxors exhibit frequency dependent susceptibilities; so unambiguous evidence will require many different kinds of measurement. 6.4. Low-Temperature Ferromagnetism? As explained earlier, BiFeO3 is antiferromagnetic at room temperature, with the weak local canting moment being completely cancelled by the averaging out effect of the cycloid. However, there are several reports, including hysteresis measurements in single crystals (Fig. 16) suggesting that at very low temperatures there could be a weakly ferromagnetic state.[157,158] It is important to conrm whether or not this is intrinsic because, although the net magnetic moment is minuscule (ca.106 mB per Fe), it would have important consequences regarding magnetic symmetry and, thus, also whether or not the linear magnetic coupling is allowed. The existence of ferromagnetism at very low temperatures would also reect an underlying competition between antiferromagnetic and ferromagnetic interactions, which, of course, would be consistent with the spin-glass state in the intermediate temperature range. On the other hand, the observation of ferromagnetic hysteresis at low temperatures is not universal and may be explained by even a very small concentration of impurities; Lebeugle et al., for example, note that just 1 mol% of paramagnetic Fe3 (probably due to the presence of Bi25FeO39) can account for all the low-temperature magnetic enhancement in their single crystals, and that removing such impurities with HNO3 removes virtually all traces of ferromagnetism in their samples.[15] 6.5. Elastic Anomalies at Magnetic Transitions Redfern et al. have used dynamic mechanical analysis (an oscillating three-point bending measurement) to estimate the elastic constant of BiFeO3 ceramics below room temperature. [142] Their reported results show an anomaly between 200250 K and

perhaps another one near 140 K. The exact temperature of the anomaly at ca. 225 K depends strongly on the frequency of the applied mechanical stress. This frequency dependence can be due to a thermally activated defect state, with an Arrhenius-type behavior, or else it is an indication of some glassy underlying process. While the mechanical data alone does not allow elucidation of the answer, it is worth pointing out that between 200 and 250 K there are also indications of a magnetic transition, the nature of which is still unclear but possibly related to a spin-glass state (see Section 6.3). It is therefore possible that the elastic anomaly is due to magnetoelastic (magnetostrictive) coupling to a glassy magnetic transition. We have also measured the mechanical response of BiFeO3 above room temperature and several broad peaks are apparent. These are rather puzzling, since no structural transition has been reported between room temperature and the Neel temperature. It is again possible that these mechanical anomalies are caused by defect states, but they may also be real: as discussed in Section 2, a large number of ghost transitions, which are awaiting clarication, have been reported for BiFeO3. Resonant ultrasound spectroscopy has also been deployed to characterize the elastic behavior of BiFeO3. Preliminary lowtemperature data (Fig. 17) shows a very clear transition in the region 3060 K, where elastic attenuation leads to the disappearance and re-entrance of the elastic resonances. This massive attenuation is suggestive of a highly dissipative state. Given that the magnetic measurements suggest a spin-glass state in this temperature region, the elastic measurements are consistent with coupling between elasticity and the spin glass, lending support to a magnetoelastic mean-eld character for the transition.

REVIEW

7. Magnetoelectric Coupling
7.1. Magnetoelectric Coupling and Spin Cycloid The existence of a spin cycloid averages out any linear magnetoelectric (ME) coupling between polarization (P) and

Figure 16. Magnetization of BiFeO3 single crystals at low temperatures. Figure courtesy of M. Singh, University of Puerto Rico.

Figure 17. Resonant ultrasound spectroscopy of a BiFeO3 ceramic. All the resonant peaks disappear in the low-temperature region between ca. 50 and 30 K, a temperature range in which the ac magnetic susceptibility and dielectric constant have also been reported to show broad anomalies. Other elastic anomalies can also be seen at higher temperatures. Figure courtesy of Julia Herrero-Albillos and Michael Carpenter, University of Cambridge.

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2475

www.advmat.de

magnetization (M). Any macroscopic magnetoelectric coupling must therefore be higher order (quadratic). Indeed, up to magnetic elds of several Tesla the magnetically induced polarization is found to be proportional to the square of the magnetic eld (Fig. 18a). The full magnetoelectric tensor was rst characterized by Tabares-Munoz et al.,[159] and is given by (in hexagonal coordinate axis with P3 parallel to the spontaneous polarization)
2 2 P1 b111 H1 H2 b113 H1 H3

(6) (7) (8)

P2 b113 H2 H3 2b111 H1 H2
2 2 2 P3 b311 H1 H2 b333 H3

with experimentally measured coefcients b111 5.0 1019 s A1, b113 8.1 1019 s A1, b311 0.3 1019 s A1, and b333 2.1 1019 s A1.[159] Above a certain critical eld, however, the magnetoelectric polarization markedly changes, signaling a change in the spin conguration. Ismailzade et al.[158] rst reported spin op at a critical eld, HC, of only 5 kOe, (1 Oe 104 T) but this was almost certainly an artifact, since no subsequent measurements[159,161,162] were able to reproduce it. The real critical eld appears to be much higher, at ca. 20 T[161,162] (Fig. 18). Above this critical value, the magnetoelectric polarization changes sign and becomes linearly dependent on magnetic eld (Fig. 18, left). Since the linear magnetoelectric effect is forbidden by the cycloid, its onset signals that the cycloid has been destroyed by the high magnetic eld. A second observation is that above the critical eld for the cycloid destruction (or spin op), the eld-induced magnetization jumps to a higher value. Linear extrapolation of this eld-induced magnetization to zero-eld yields a remnant magnetization ca.0.3 emu g1 (Fig. 18, right). The theory behind these effects is subtle. The local (shortrange) magnetic symmetry of BiFeO3 is such that, if it were centrosymmetric (paraelectric), it would be a perfect G-type

antiferromagnet with no net magnetic moment. However, the ferroelectric polarization breaks the center of symmetry and induces a small canting of the spins via the Dzyaloshinskii Moriya interaction. This canting results in the very small magnetization of 0.3 emu g1.[32,162] Ferroelectrically induced canting magnetism is neither new nor unique to BiFeO3, as it was already reported for BaMnF4 in the 1970s.[28] What is special about BiFeO3 is that, in addition to this canting, there is also a ferroelectrically induced spin cycloid that averages out the local canted magnetism. This cycloid appears because polarization can also couple to gradients of magnetization, thereby inducing an inhomogeneous spin conguration (the spin cycloid).[161,162] This is the converse effect of the ferroelectric polarization induced by magnetic spirals.[25,26] Although the cycloid averages out the macroscopic canting moment, this is locally still present at the unit-cell level. High magnetic elds can destroy the cycloid, thereby recovering the canted state and its associated remnant magnetization (Fig. 18b) and, in this state, the linear magnetoelectric is allowed (Fig. 18a), so both effects in Figure 18 are consistent with each other. The spin cycloid can also be destroyed by doping[164] or by epitaxial strain,[165] so fully strained epitaxial thin lms can in principle display a weak remnant magnetization, although not as big as initially reported.[20,21]

REVIEW

7.2. Ferroelectric Control of Magnetism

Recent experimental works have explored in more detail the relationship between the ferroelectric polarization and magnetic symmetry. With the exception of the work by Kubel and Schmid,[61] early bulk research was performed on samples which were mostly either polycrystalline or polydomain single crystals and, hence, some subtle directional effects were averaged out. Groups at Saclay[129] and Rutgers,[140] however, have managed to make ferroelectric monodomain crystals of BiFeO3 by growing them at temperatures below the ferroelectric transition, and have deployed very high-resolution neutron diffraction to elucidate the relationship between ferroelectricity and antiferromagnetism in this compound. Specically, they have shown that the magnetic moments rotate within the plane dened by the polarization (P// [111]pseudocubic) and the cycloid propagation vector (k// [1,0,1]pseudocubic) (Fig. 19). This has profound consequences, for if the direction of the polarization is changed, so too will the magnetic easy plane: indeed, by applying a voltage and switching the polarization by 718, both Lebeugle et al.[129] and Lee et al.[140] were able to show that the magnetic easy planes were rotated. Importantly too, the magnetic easy plane can be switched only if the Figure 18. Magnetoelectric effect in BiFeO3 (left): at low elds, P is proportional to H2, (quadratic polarization changes direction, but not if it ME coupling). Above BC 20 T, P is linearly dependent on H instead. Since the linear ME is merely changes polarity; 1808 switching of the forbidden in the presence of a cycloid, the cycloid is destroyed above 20 T. Note that, in any case, polarization should not affect the magnetic 5 the magnetically induced polarization is very small (ca.10 times smaller than the ferroelectric orientation. polarization). Once the cycloid is destroyed, the small canted magnetic moment is recovered The recent observations in single crystals (right). Extrapolation of the magnetization to zero eld yields a small net magnetization of ca. are closely related to a previous investigation 0.3 emu g1. Figures adapted from [161], with permission. Copyright 2006, Elsevier.

2476

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

Figure 19. Relationship between the magnetic easy plane containing the spins, the vector of ferroelectric polarization, and the vector of cycloid propagation. Rotating the polarization by 718 (i.e., switching only one of the components of the polarization) results in a change of the magnetic easy plane, meaning that sublattice magnetization can be switched by an applied voltage. Reproduced with permission from [129]. Copyright 2008, American Physical Society.

antiparallel polarities for each direction: hence there are eight different polar domains in BiFeO3. Separating adjacent domains, there are three possible types of ferroelectric domain wall, which are usually labeled according to the angle formed between the polarization vectors on either side of the wall. When only one component of the diagonal polarization is reversed (say, one domain has [111] orientation and the adjacent one is [111]), then the polar vectors form an angle of approximately 718 and the domain wall that separates the two polarizations is called a 718 wall. When two polar components are reversed, it is a 1098 wall, and when all three components of the polarization are reversed it is a 1808 wall. This is schematically depicted in Figure 20. Minimization of electrostatic and elastic elds imposes constraints on the orientation of the walls; typically, 718 walls are parallel to {110} planes, 1098 walls are in {100} planes, and 1808 walls should be on planes containing the [111] polar vector.[167] Of course, nonequilibrium walls can be in different planes and may even be very irregular in shape.

REVIEW

8.2. Domain Size 8.2.1. Stripe Domains Since the work of Landau and Lifshitz in 1935,[168] and later Kittel in 1946,[169] it is understood that domain size scales as the square root of lm thickness. While their arguments were initially proposed for ferromagnetic lms, they were later extended to ferroelectric materials[170] and ferroelastics,[171] and more recently also to multiferroics[172] and nanostructures other than thin lms.[173] The basic argument is that domain size results from the competition between a surface energy (demagnetization, depolarization, strain) which is directly proportional to domain width, w, and a domain-wall energy that is proportional to the number density of walls and, hence, inversely proportional to w. The energy of the walls is also proportional to their size, which scales as the lm thickness, d. Minimization of these components leads p to the familiar expression w A d, where A is a constant. This scaling has been experimentally veried for the ferroelectric stripe domains of BiFeO3 lms,[174] but not for ultrathin lms, which have a very different domain morphology (Fig. 21).

on thin lms reported by Zhao et al.[166] The thin lms had no spin cycloid (due to either strain or reduced thickness), and instead they had the homogeneous G-type antiferromagnetism (with a slight canting), with the magnetic easy plane perpendicular to the ferroelectric polarization. Using a combination of piezoelectric-force microscopy and X-ray photoelectron microscopy, these authors have managed to visualize simultaneously the ferroelectric and the antiferromagnetic domains, establishing that both types of domains are completely correlated with each other. Again, switching the polarization by an angle other than 1808 (in rhombohedral symmetry, the polar vectors can be switched by 718 and 1098 as well as 1808, as discussed in Section 8) changes the magnetic easy plane. This electrically induced switching of the magnetic easy plane is of seminal importance, because it opens the possibility of magnetoelectric devices based on the voltage control of magnetization (see Section 9).

8. Domains and Domain Walls


Research on domains and domain walls has intensied recently because i) the behavior of domains is directly responsible for switching characteristics (switching of polarization takes place through nucleation and growth of domains) and ii) domain size scales with sample size, so thin lms can have very small domains and, therefore, a high volume density of domain walls. BiFeO3 displays new domain-wall related phenomena of its own which make this subject particularly fascinating. 8.1. Domain Walls in Rhombohedral Symmetry In rhombohedral BiFeO3 the ferroelectric polarization can point along any of the four diagonals of the perovskite unit cell, with two

Figure 20. Schematic of the three types of ferroelectric domain walls separating domains with one, two, or all three components of the polarization switched. The domain walls are labeled according to the angle between the polarization vectors on either side. Note that in this simplied picture the 718 and 1808 walls are not in their most stable congurations, since the head-to-head polarization perpendicular to the wall would lead to large electric elds at the interface.

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2477

www.advmat.de

Figure 21. Ferroelectric stripe domains (left) and fractal domains (right) in epitaxial thin lms of BiFeO3. The latter tend to appear in very thin lms only. (PFM images of stripe domains and fractal domains courtesy of Y. H. Chu and M. Bibes, respectively).

8.2.2. Fractal Domains For ultrathin lms the domains in BiFeO3 are no longer striped and instead form irregularly shaped mosaic structures (Fig. 20).[115,175] The critical thickness depends on substrate and electrodes, among other things. The domain walls are rough, and are well described by a fractal Hausdorff dimension, h ca. 2.5 (for a perfectly smooth domain wall it should be two). The Hausdorff dimension describes the scaling between the area of the domains and their perimeters; since the domain size results from a competition between domain energy (proportional to area) and the wall energy (proportional to perimeter), it is reasonable to expect that Kittels law will need to be modied in order to incorporate this fractality. The modied equation turns out to be[175]
h?

the conductivity of the walls is directly related to the type of domains they separate. Thus, 1808 walls are the most conductive, followed by 1098 walls and, nally, the 718 walls, which, in fact, do not have any measurable transport enhancement. The authors argue that there are at least two reasons for the enhanced conductivity of the walls. First, the polarization normal to the domain wall is observed not to be constant across it; this generates an electrostatic depolarization eld that may attract charge carriers. Second, the electronic bandgap is considerably reduced for the 1808 and 1098 domain walls. A plausible explanation for the bandgap decrease has to do with the local distortion of the FeOFe bond angle, which, as discussed in Section 3, controls the orbital overlap. In the middle of the domains, the octahedra are quite buckled (and hence the gap is big). If the unit cell expands, the buckling angle can become straighter, thereby increasing the orbital overlap and reducing the bandgap. The local suppression of polarization at the domain walls leads to precisely such a volume expansion via the cancelling of the spontaneous strain (Fig. 22) and, hence, the local straightening of the bond angles at the walls reduces the gap in much the same way as temperature or pressure would. This model is consistent with the correlation between the type of wall and its conductivity. The absolute value of the polarization is smallest in the middle of the 1808 walls (Pwall 0), intermediate p in the 1098 walls (Pwall P0/ 3 ca. 0.57P0), and maximum in p the 718 walls (Pwall P0p2 ca.0.82P0); therefore, the volume 3 change (and the associated change in conductivity) will be biggest for the 1808 walls, intermediate in the 1098 walls, and smallest in the 718 walls, in agreement with the observed sequence of wall conductivities. 8.4. Domain-Wall Magnetization: PrivratskaJanovec and Daraktchiev Theories Privatska and Janovec[177] observed that magnetoelectric coupling could lead to the appearance of net magnetization in the middle of antiferromagnetic domain walls. Specically, they showed that this effect is allowed for R3c space groups, which is the symmetry of BiFeO3, but their group-symmetry arguments do not allow any quantitative estimate. Later, Fiebig and co-workers analyzed magnetoelectric coupling in the walls of multiferroics such as YMnO3 and HoMnO3.[178,179] Daraktchiev et al. have proposed a thermodynamic (Landau-type) model[172] with the aim of quantitatively estimating whether the walls of BiFeO3 can be magnetic and, if so, to what extent they might contribute to the observed enhancement of

REVIEW
w A0 d 3hk
0

(9)

where A is a constant and h?, hjj are the Hausdorff dimensions of the domain wall in the directions perpendicular and parallel to the plane of the lm, respectively. This becomes the standard Kittels law when the walls are perfectly smooth (h? hjj 1). The reason for the thickness-induced transition to fractal morphology remains unknown, but we note that irregular walls are elastically costly, so they cannot be the equilibrium conguration in a perfect crystal. Two necessary ingredients for their appearance must be a random distribution of pinning defects and a low crystal anisotropy (so that the wall can deform without much elastic cost). In this respect, the possible thickness transition to a tetragonal symmetry for ultrathin lms of BiFeO3 [113] would indeed facilitate the latter, since tetragonal c-axis domains face no elastic constraints (other than surface minimization) on the in-plane orientation of the 1808 walls. 8.3. Domain-Wall Conductivity Domain walls have their own local symmetry and, hence, also their own properties. In the case of BiFeO3, this includes enhanced local conductivity. Ramesh and co-workers have recently reported that certain domain walls of BiFeO3 are much more conductive than the domains themselves.[176] Furthermore,

Figure 22. The suppression of the ferroelectrically induced in-plane contraction (Q13 is negative) leads to a lattice expansion in the direction perpendicular to the wall, which is accommodated by straightening of the bond angle. This results in increased orbital overlap and higher conductivity.

2478

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

magnetization in ultrathin lms. Their starting point is a very basic two-parameter expansion with biquadratic coupling (which is always symmetry-allowed) between net polarization and net magnetization

REVIEW

DG

a 2 b 4 a b P P krP2 M2 M4 2 4 2 4 g lrM2 P 2 M2 2

(10)

Analysis of the phase space of this thermodynamic potential shows that it is possible for net magnetization to appear in the middle of ferroelectric walls even when the domains themselves are not ferromagnetic (Fig. 23). This, however, is presently just only a toy model which does not take into account the exact symmetry of BiFeO3, so it cannot yet quantitatively estimate how much domain walls can contribute to the magnetization of BiFeO3. The exact theory of magnetoelectric coupling at the domain walls of BiFeO3 remains to be formulated.

Figure 24. Ferromagnetic hysteresis due to uncompensated surface spins in BiFeO3 nanocrystals. Reproduced with permission from [180]. Copyright 2007, American Chemical Society.

9. Bismuth Ferrite Nanotubes, Nanowires, Nanocrystals


There is a fast-growing body of research devoted to the manufacture and characterization of complex nanoscopic shapes other than thin lms. These 3D nanostructures generally have their own distinctive size effects, and multiferroic BiFeO3 is no exception. For example, nanocrystals of BiFeO3 show enhanced magnetization and superparamagnetism correlated with decreasing diameter[180] (Fig. 24). Similar size-induced magnetism has also been reported for BiFeO3 nanowires[181] and nanopowders.[125] This is thought to be due to the large fraction of uncompensated spins from the surfaces of the nanocrystals, an effect that is well known from classic antiferromagnets such as NiO.[182] Prof. Wong at SUNY Stony Brook reported crystalline BiFeO3 nanotubes in 2004.[183] These tubes were 240300 nm in diameter and as much as 50 mm long. Prepared via a sol-gel technique and

a porous alumina (AAO) template technique, they were polycrystalline with some amorphous content. Those authors removed the alumina template completely by immersion in NaOH, meaning that the tubes were left out in a pile which was hard to characterize electrically. Zhang et al.[184] used porous alumina templates instead, managing to make ordered arrays of standing BiFeO3 nanotubes and measure their piezoelectric hysteresis loops. This proved that they were indeed ferroelectric. Such nanotubes may be of considerable interest in terms of new theories relating to them.[185] Using a Landau-type thermodynamic analysis, the phase diagrams of magnetoelectric nanotubes as a function of radius have been calculated. The small diameter of the tubes affects the ferroelectric critical temperature of the ferroelectric state as a result of the stress produced by the curvature; when the ferroelectric critical temperature approaches the magnetic one, the magnetoelectric coupling can be enhanced by several orders of magnitude (Fig. 25). The theoretical predictions regarding magnetoelectric enhancement still await experimental conrmation.

10. Device Applications


10.1. Ferroelectricity and Piezoelectricity Being a room-temperature multiferroic, BiFeO3 is an obvious candidate for applications. Interestingly, however, the rst application that may reach the market might not use the multiferroic properties of BiFeO3 at all. The remnant polarization of BiFeO3 is very large, 100 mC cm2 along the polar [111] direction. To put this into context, this is the biggest switchable polarization of any perovskite ferroelectric, and is roughly twice as big as

Figure 23. Domain-wall prole in a hypothetical multiferroic with biquadratic magnetoelectric coupling between polarization and magnetization. Note that the domain walls have net magnetic moment, even though the domains themselves do not.

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2479

www.advmat.de

REVIEW

Figure 25. Magnetoelectric coefcient as a function of nanotube (R12 > 0) and nanowire (R12 < 0) radius. The magnetoelectric coefcient increases as the external radius decreases. Reproduced with permission from [185]. Copyright 2008, American Physics Society.

Figure 26. Enhanced piezoelectric coefcient, d33, of thin lms at the morphotropic phase boundary between pure BiFeO3 and SmFeO3. Reproduced with permission from [190]. Copyright 2008 American Institute of Physics.

the polarization of the most widely used material in ferroelectric memories, PZT. Moreover, unlike PZT, BiFeO3 is a lead-free material, a bonus regarding health and safety. It is therefore not surprising that manufacturers of ferroelectric memories such as Fujitsu are considering BiFeO3 as the potential active material in their next generation of ferroelectric memory devices.[186] For such an application to ever come into fruition, however, important obstacles must be removed, such as: i) the higher conductivity (and thus also dielectric losses) of BiFeO3 relative to PZT, ii) its tendency to fatigue,[187] and iii) the fact that it appears to thermally decompose at voltages quite close to the coercive voltage.[53] A second potential application unrelated to magnetoelectric properties is piezoelectricity. The piezoelectric coefcient of pure BiFeO3 is actually quite small, as argued in Section 4.2. However, its rhombohedral ground state means that mixing it with a tetragonal ferroelectric such as PbTiO3 leads to a morphotropic phase boundary (MPB) at a composition of 30% mol PbTiO3.[188] This is important because MPBs are commonly thought to be the key behind the large piezoelectric coefcients of PZT and relaxors,[189] so the MPB of BiFeO3PT might lead to equally large piezoelectric constants. A newer MPB with enhanced piezoelectric coefcients has been reported[190] for a solid solution of rhombohedral BiFeO3 with orthorhombic SmFeO3 (Fig. 26); it is likely that other solid solutions between BiFeO3 and any of the other orthoferrites should also display similar MPBs. In this context, it is surprising that the widely studied solid solution between BiFeO3 and LaFeO3 does not appear to have yet been piezoelectrically characterized; perhaps this is due to its high conductivity.

a femtosecond laser pulse, BiFeO3 lms emit THz radiation, which is currently of great interest for many applications ranging from telecommunications to security.[192] Furthermore, the authors note that the THz radiation is completely correlated with the poling state of the lms (Fig. 27); accordingly, THz emission could provide an ultrafast (picosecond response time) and non-destructive method for ferroelectric

10.2. Terahertz Radiation Another possible application which does not make use of the magnetoelectric properties of BiFeO3 is its reported emission of THz radiation. Takahashi et al.[191] reported that, when hit with

Figure 27. a) Experimental setup used by Takahashi et al. [191] for measuring the THz emission of BiFeO3. (LT: low temperature; LSAT: LaAlO3Sr2AlTaO6). b) Experimentally measured radiation and c) the Fourier components of the amplitude of the measured THz radiation switch when the ferroelectric polarization is switched, indicating the possibility of using THz-based optical methods for reading ferroelectric memories. Copyright 2006, American Physics Society.

2480

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

memory readout. As an additional advantage, at such high frequencies the response is insensitive to leakage, which automatically gets rid of one of the major obstacles for implementing BiFeO3 as a ferroelectric memory material. 10.3. Spintronics But the real drivers behind most of the applied research on BiFeO3 are magnetoelectric and spintronic applications.[193] Chief among these would be memories that can be written using a voltage and read using a magnetic eld. Using a voltage for writing has three advantages: i) this can be implemented in a solid-state circuit without mobile parts, ii) it has a low-energy requirement, and iii) the voltage requirements automatically scale down with thickness. Reading the memory magnetically, on the other hand, has the advantage that it is a non-destructive readout process, unlike direct ferroelectric reading, which requires switching the polarization in order to read it. For such memories to actually work, the magnetic state therefore must be a) electrically switchable and b) magnetically readable. As discussed in the previous section, the rst condition is met by BiFeO3, because the easy plane of its antiferromagnetic domains is correlated with the polar direction, and rotating the ferroelectric polarization results in a rotation of the sublattice magnetization,[129,140,166] i.e., the magnetic state of the sample can be changed by a voltage. On the other hand, the second condition is not directly met, because antiferromagnetic (or, at best, weakly canted antiferromagnetic) domains cannot be easily read. An elegant solution to the problem of reading antiferromagnetic states consists in using the mechanism known as exchange bias. Crudely, exchange bias is the magnetic interaction between the spins at the uppermost layer of an antiferromagnet and a thin ferromagnetic layer attached to it. The exchange bias modies the hysteresis loops of the ferromagnetic layer, either offsetting or widening them.[194] What is relevant here is that voltage-induced changes to the underlying antiferromagnetic domains will result in changes to the ferromagnetic hysteresis of the upper layer, which can then be read by conventional mechanisms. The implementation of this concept for Cr2O3 (which is magnetoelectric but not ferroelectric) was rst done by Borisov et al.,[195] and the rst investigation with an actual multiferroic (YMnO3) was done by Laukhin et al.[196] The race to implement this idea using BiFeO3 (which has the advantage over YMnO3 that it works at room temperature) has been on for a while, and has been punctuated by several important milestones, such as the observation of exchange bias in thin ferromagnetic layers grown on BiFeO3[197,198] (Fig. 28), the
Figure 29. The magnetization of a small island of ferromagnetic CoFe is switched when a voltage is applied to the underlying BiFeO3 (white and black correspond to different magnetic polarities). Figure reproduced with permission from [200]. Copyright 2008, Nature Publishing Group.

REVIEW

Figure 30. Tunneling magnetoresistance (TMR) of BiFeO3 sandwiched between (La,Sr)MnO3 and Co. Reproduced with permission from [193]. Copyright 2008, Institute of Physics.

correlation between exchange bias and ferroelectric domains,[199] the observation that the antiferromagnetic domains can be switched by a voltage,[166] and, most recently, the nal proofof-concept that the exchange-biased ferromagnetic layer can indeed be switched by a voltage[200] (Fig. 29). A second line of work uses BiFeO3 as a barrier layer in spintronics. Sandwiching BiFeO3 between two ferromagnetic metals results in tunneling magnetoresistance[193,198] (Fig. 30). For this, the only requirement is that the BiFeO3 layer be reasonably insulating down to tunneling thicknesses. However, an extra ingredient provided by BiFeO3 is the fact that it also remains a robust and switchable ferroelectric down to a thickness of 2 nm,[201] and thus it could in principle be used as an electrically switchable tunnel junction, whereby the ferroelectric state controls the magnetic state of the thin ferromagnetic electrodes, thus modifying the tunneling magnetoresistance. A similar concept using a ferromagnetic multiferroic (LaBMO) was indeed demonstrated by Gajek et al.,[202] who showed that the tunneling resistivity could be controlled both by electric and magnetic elds, giving rise to a four-state memory device. The voltage-dependent barrier characteristics of BiFeO3 have not yet been established. The above developments show that, at least in principle, it is now possible to develop an MERAM (Magnetoelectric Random Access Memory) based on BiFeO3. A schematic of such a device has been proposed by Bibes and Barthelemy,[203] and is reproduced in Figure 31.

11. Closing Remarks


Figure 28. Exchange-biased magnetic hysteresis loops of thin Co grown on BiFeO3. The four gures correspond to measurements along the four in-plane orientations [100], [010], [010], [100]. Reproduced with permission from [198]. Copyright 2006, American Institute of Physics.

It is difcult to write a review on a topic that is so popular and that is continuing to develop at

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2481

www.advmat.de

REVIEW

certain models. We emphasize that the study of BiFeO3 is NOT all wrapped up. So this is not a Bible about bismuth ferrite properties; rather, it should be read as an inspiration for further studies. We hope the reader nds it useful as such.

Acknowledgements
During the elaboration of this review we have had many fruitful exchanges that have enhanced our understanding of this material. We would particularly like to thank Michel Viret, John Robertson, Jorge Iniguez, and Hans Schmid for their insights. There are also a number of unpublished results that have been kindly made available to us for this review; we are indebted to Simon Redfern, Julia Herrero-Albillos, Hong Jiawang, Michael Carpenter, Manooj Singh, Ram Katiyar, Manuel Bibes, Brahim Dhkil, Jens Kreisel, Finlay Morrison, Eddie Chu, Ramamoorthy Ramesh, Maren Daraktchiev, Maria Polomska, and Pavle Cevc for sharing their work with us. Received: September 25, 2008 Published online: May 4, 2009 Note added in proof: After submission of this paper, Choi et al. reported [205] a polarization-controlled diode effect and large photovoltaic currents in BiFeO3, opening the way for novel device applications that combine both the ferroelectric and semiconducting properties of this material.

Figure 31. MERAM based on exchange-bias coupling between a multiferroic that is ferroelectric and antiferromagnetic (FE-AFM, green layer), and a thin ferromagnetic electrode (blue). A tunneling barrier layer between the two top ferromagnetic layers provides the two resistive states. Interestingly, BiFeO3 could act not only as the magnetoelectric active layer, but also as the tunneling barrier. Reproduced with permission from [203]. Copyright 2008, Nature Publishing Group.

such a rapid rate. So, rather than just summarizing what is known so far about BiFeO3, we have chosen to focus our attention here on four basic issues which we believe are still open: 1) What are the high-temperature phases and in particular what is the nature of the MI transition (at high temperature? at high pressure? Are they the same? Are they Mott-like? is the Polomska transition at 458 K extrinsic?). 2) What are the low-temperature magnetic phases? (Is there a spin glass? If so, is it Ising-like? long-range? mean-eld? How does it couple to strain? Is Sosnowskas magnetic structure correct or Zalesskiis or neither? Is there an AlmeidaThouless AT line? Are there two magnetic transitions at 140 and 201 K or four? Is the spin-glass onset at 230 or 50 K? Is the VolgelFulcher glass-freezing temperature 29 K? Is it a ferromagnet below ca. 10 K? Are these magnetic transitions intrinsic?). 3) What are the magnetoelectric properties near room temperature? (Is there a linear magnetoelectric term? quadratic? Is it useful? for spin lters and spintronics? for memories?). 4) What are the intrinsic properties of the domain walls? (Are they ferromagnetic? If so, how much? Do they affect the functional properties of thin lms? What is the mechanism for enhanced conductivity? Is it intrinsic or extrinsic? How will that affect the performance of thin-lm devices with a high density of domains?) There is an urgent need for many more studies focusing on the phase diagram and the dynamics. Very little yet is known about switching processes. As a nal, concrete example: If we dont know the magnetic space group or even point-group symmetry of this material below room temperature in what could be several magnetic phases plus a glassy phase, how can we even decide whether a linear magnetoelectric coupling is allowed or forbidden? We cannot provide denitive answers to most of these questions. For some readers that may be a disappointment: Why write a 15 000-word review if you dont know the answers? Instead, our aim has been to provide a fair view of the pertinent works from different sources, stating as clearly as possible what we think the questions are, together with evidence for and against

[1] [2] [3] [4] [5] [6] [7] [8] [9]

[10]

[11] [12] [13] [14]

[15]

[16] [17] [18] [19]

P. Curie, J. Physique 1894, 3, 393. J. Valasek, Phys. Rev. 1920, 15, 537. A. Perrier, A. J. Staring, Arch. Sci. Phys. Nat. (Geneva) 1922, 4, 373. A. Perrier, A. J. Staring, Arch. Sci. Phys. Nat. (Geneva) 1923, 5, 333. T. H. ODell, The Electrodynamics of Magneto-electric Media, NorthHolland, Amsterdam 1970. a) I. E. Dzyaloshinskii, Zh. Eksp. Teor. Fiz. 1959, 37, 88. b) Sov. Phys. JETP 1959, 10, 628. a) D. N. Astrov, Zh. Eksp. Teor. Fiz. 1960, 38, 984. b) Sov. Phys. JETP 1960, 11, 708. E. Ascher, H. Rieder, H. Schmid, H. Stossel, J. Appl. Phys. 1966, 37, 1404. a) G. A. Smolensky, V. A. Isupov, A. I. Agronovskaya, Sov. Phys. Solid State 1959, 1, 150. b) G. A. Smolenskii, A. I. Agranovskaya, S. N. Popov, V. A. Isupov, Zh. Tekh. Fiz. 1958, 28, 2152. c) A. Smolenskii, A. I. Agranovskaya, S. N. Popov, V. A. Isupov, Sov. Phys. Tech. Phys. 1958, 3, 1981. a) G. A. Smolenskii, I. E. Chupis, Usp. Fiz. Nauk 1982, 137, 415. b) G. A. Smolenskii, I. E. Chupis, Sov. Phys. Usp. 1982, 25, 475. c) G. A. Smolenskii, I. E. Chupis, Sov. Phys. Solid State 1962, 4, 807. H. Schmid, Ferroelectrics 1994, 162, 665. M. Fiebig, J. Phys. D 2005, 38, R123. W. Eerenstein, N. D. Mathur, J. F. Scott, Nature 2006, 442, 759. 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, R. Ramesh, Science 2003, 299, 1719. a) D. Lebeugle, D. Colson, A. Forget, M. Viret, P. Bonville, J. F. Marucco, S. Fusil, Phys. Rev. B 2007, 76, 024 116. b) D. Lebeugle, D. Colson, A. Forget, M. Viret, P. Bonville, J. F. Marucco, S. Fusil, Appl. Phys. Lett. 2007, 91, 022 907. D. Lebeugle, Groupement de Recherche sur les Nouveaux Etats Electroniques de la Matiere: GdR NEEM; Gif sur Yvette, November 2006. R. P. S. M. Lobo, R. L. Moreira, D. Lebeugle, D. Colson, Phys. Rev. B 2007, 76, 172 105. M. Cazayous, D. Malka, D. Lebeugle, D. Colson, Appl. Phys. Lett. 2007, 91, 071910. D. Lebeugle, Ph.D. Thesis Saclay Institute of Matter and Radiation, 2007.

2482

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

REVIEW

[20] W. Eerenstein, F. D. Morrison, J. Dho, M. G. Blamire, J. F. Scott, N. D. Mathur, Science 2005, 307, 1203a. [21] H. Bea, M. Bibes, A. Barthelemy, K. Bouzehouane, E. Jacquet, A. Khodan, J.-P. Contour, Appl. Phys. Lett. 2005, 87, 072508. [22] N. A. Hill, J. Phys. Chem. 2000, B104, 6694. [23] A. Filippetti, N. A. Hill, Phys. Rev. B 2002, 65, 195120. [24] a) R. E. Cohen, Nature 1992, 358, 136. b) Z. Wu, R. E. Cohen, Phys. Rev. Lett. 2005, 95, 9 037 601. [25] M. Mostovoy, Phys. Rev. Lett. 2006, 96, 067601. [26] S.-W. Cheong, M. Mostovoy, Nature Materials 2007, 6, 13. [27] I. A. Sergienko, C. Sen, E. Dagotto, Phys. Rev. Lett. 2006, 97, 227204. [28] D. V. Efremov, J. van den Brink, D. I. Khomskii, Nature Materials 2004, 3, 853. [29] G. Nenert, T. T. M. Palstra, J. Phys.: Condens. Matter 2007, 19, 406213. [30] G. Catalan, Phase Transitions 2008, 81, 729. [31] D. L. Fox, J. F. Scott, J. Phys. C: Sol. St. Phys. 1977, 10, L329. [32] C. J. Fennie, Phys. Rev. Lett. 2008, 100, 167203. C. Ederer, C. J. Fennie, J. Phys. Cond. Mat. 2008, 20, 434219. [33] J. F. Scott, Ferroelectrics 1995, 166, 95. [34] S. C. Abrahams, Acta Cryst. B 2007, 63, 257; 2003, 59, 811; 2003, 59, 541; 2002, 58, 34; 2002, 58, 316; 2000, 56, 793; 1996, 52, 790. [35] S. C. Abrahams, K. Mirsky, R. M. Nielson, Acta Cryst. B 1996, 52, 806. [36] S. C. Abrahams, Ferroelectrics 1990, 104, 37; Acta Cryst. B 1999, 55, 494; 1989, 45, 228; 1988, 44, 585. [37] S. C. Abrahams, J. Ravez, Ferroelectrics 1992, 135, 21. [38] J. Ravez, S. C. Abrahams, R. de Pape, J. Appl. Phys. 1989, 65, 3987. [39] Y. Calage, S. C. Abrahams, J. Ravez, J. Appl. Phys. 1990, 67, 430. [40] R. Blinc, G. Tavcar, B. Zemva, D. Hanzel, P. Cevc, C. Filipic, A. Levstik, Z. Jaglicic, Z. Trontelj, N. Dalal, V. Ramachandran, S. Nellutla, J. F. Scott, J. Appl. Phys. 2008, 103, 074114. [41] E. Kroumova, M. I. Aroyo, J. M. Perez-Mato, R. Hundt, Acta Cryst. B 2001, 57, 599. [42] S. Sarraute, J. Ravez, R. VonderMuhll, G. Bravic, R. S. Feigelson, S. C. Abrahams, Acta Cryst. B 1996, 52, 72. [43] S. C. Abrahams, J. Albertsson, C. Svensson, Acta Cryst. B 1990, 46, 497. [44] a) M. Lorient, R. VonderMuhll, J. Ravez, A. Tressaud, Solid State. Commun. 1980, 36, 383. b) M. Lorient, R. VonderMuhll, J. Ravez, A. Tressaud, Solid State. Commun. 1981, 40, 847. [45] a) A. Tressauld, J. Ravez, M. Lorient, J. Fluorine Chem. 1982, 21, 30. b) A. Tressauld, J. Ravez, M. Lorient, Rev. Chim. Minerale 1982, 19, 128. [46] J. Ravez, J. Physique III 1997, 7, 1129. [47] E. J. Speranskaya, V. M. Skorikov, E. Ya. Kode, V. A. Terektova, Bull. Acad. Sci. U. S. S. R. 1965, 5, 873. [48] M. I. Morozov, N. A. Lomanova, V. V. Gusarov, Russ. J. Gen. Chem. 2003, 73, 1680. [49] R. Palai, R. S. Katiyar, H. Schmid, P. Tissot, S. J. Clark, J. Robertson, S. A. T. Redfern, G. Catalan, J. F. Scott, Phys. Rev. B 2008, 77, 014110. [50] M. Valant, A.-K. Axelsson, N. Alford, Chem. Mater. 2007, 19, 5431. [51] N. N. Krainik, N. P. Khuchua, V. V. Zhdanova, V. A. Evseev, Sov. Phys. Solid State 1966, 8, 654. [52] H. Schmid: private communication. [53] X. J. Lou, C. X. Yang, T. A. Tang, Y. Y. Lin, M. Zhang, J. F. Scott, Appl. Phys. Lett. 2007, 90, 262908. [54] X. J. Lou, X. J. M. Zhang, M. S. A. T. Redfern, J. F. Scott, Phys. Rev. Lett. 2006, 97, 177601. [55] G. Garton, S. H. Smith, B. M. Wanklyn, J. Cryst. Growth 1971, 13, 588. [56] D. Elwell, A. W. Morris, B. W. Neate, J. Cryst. Growth 1972, 16, 67. [57] H. Okamoto, J. Phase Equilib. 1991, 12, 207. [58] T. Atsuki, N. Soyama, T. Yonezawa, K. Ogi, Jpn. J. Appl. Phys. 1995, 34, 5906. [59] J. M. Moreau, C. Michel, R. Gerson, W. J. James, J. Phys. Chem. Solids 1971, 32, 1315. [60] a) V.S Filipev, I.P Smolyaninov, E.G Fesenko, I.I Belyaev, Kristallograya 1960, 5, 958. b) Sov. Phys. Crystallogr. 1960, 5, 913.

[61] F. Kubel, H. Schmid, Acta Cryst. B 1990, 46, 698. [62] J. D. Bucci, B. K. Robertson, W. J. James, J. Appl. Cryst. 1972, 5, 187. [63] J. R. Chen, W. L. Wang, J.-B. Li, G. H. Rao, J. Alloys Compd. 2008, 459, 66. [64] A. Palewicz, R. Przenioso, I. Sosnowska, A. W. Hewat, Acta Cryst. 2007, B63, 537. [65] V. M. Goldschmidt, Naturwissenschaften 1926, 14, 477. [66] R. D. Shannon, Acta Cryst. A 1976, 32, 751. [67] H. D. Megaw, C. N. W. Darlington, Acta Cryst. A 1975, 31, 161. [68] M. Polomska, W. Kaczmarek, Z. Pajak, Phys. Stat. Sol. 1974, 23, 567. [69] N. N. Krainik, N. P. Khuchua, A. A. Bierezhnoi, A. G. Tutov, A. J. Cherkashtschenko, Izv. Akad. Nauk ser. Fiz. 1965, 29, 1026. [70] a) R. Haumont, J. Kreisel, P. Bouvier, F. Hippert, Phase Trans. 2006, 79, 1043. b) R. Haumont, J. Kreisel, P. Bouvier, F. Hippert, Phys. Rev. B 2006, 73, 132 101. [71] I. A. Kornev, S. Lisenkov, R. Haumont, B. Dkhil, L. Bellaiche, Phys. Rev. Lett. 2007, 99, 227602. [72] R. Haumont, I. A. Kornev, S. Lisenkov, L. Bellaiche, J. Kreisel, B. Dkhil, Phys. Rev. B 2008, 78, 134108. [73] S. M. Selbach, T. Tybell, M.-A. Einarsrud, T. Grande, Adv. Mater. 2008, 20, 3692. [74] D. C. Arnold, K. S. Knight, F. D. Morrison, P. Lightfoot, Phys. Rev. Lett. 2009, 102, 027602. [75] S. A. T. Redfern, J. N. Walsh, S. M. Clark, G. Catalan, J. F. Scott, arXiv:0901.3748, 2009. [76] a) E. Ascher, J. Phys. C: Solid State Phys. 1977, 10, 1365. b) M. V. Jaric, D. Mukamel, Nucl. Phys. 1990, B336, 475. [77] S. M. Selbach, T. Tybell, M.-A. Einarsrud, T. Grande, Chem. Mater. 2007, 19, 6478. [78] A. G. Gavriliuk, V. V. Struzhkin, I. S. Lyubutin, S. G. Ovchinnikov, M. Y. Hu, P. Chow, Phys. Rev. B 2008, 77, 155 112. [79] O. E. Gonzalez-Vazquez, J. Iniguez, Phys Rev. B 2009, 79, 064102. [80] P. Karen, D. Legut, M. Friak, M. Sob, Phys. Today 2008, 61, 10. [81] a) A. Pashkin, K. Rabia, S. Frank, C. A. Kuntscher, R. Haumont, R. Saint-Martin, J. Kreisel, arXiv:0712.0736v1, 2007.b) R. Haumont, P. Bouvier, A. Pashkin, K. Rabia, S. Frank, B. Dkhil, W. A. Crichton, C. A. Kuntscher, J. Kreisel, arXiv:0811.0047, 2008. [82] a) A. G. Gavriliuk, I. S. Lyubiutin, V. V. Strukin, JETP Lett. 2007, 86, 532. b) A. G. Gavriliuk, I. S. Lyubiutin, V. V. Strukin, JETP Lett. 2007, 86, 197. c) A. G. Gavriliuk, I. S. Lyubiutin, V. V. Strukin, Mater. High Press. 2007, 987, 147. [83] A. G. Gavriliuk, V. Struzhkin, I. S. Lyubutin, I. A. Trojan, Mi, Y. Hu, P. Chow, Mater. Res. Soc. Symp. Proc. 2007, 987, 05. [84] J. F. Scott, R. Palai, A. Kumar, M. K. Singh, N. M. Murari, N. K. Karan, R. S. Katiyar, J. Am. Ceram. Soc. 2008, 91, 1762. [85] M. Polomska, W. Kaczmarek, Acta Phys. Polonica A 1974, 45, 199. [86] Z. V. Gabbasova, M. D. Kuzmin, A. K. Zvezdin, I. S. Dubenko, V. A. Murashov, D. N. Rakov, I. B. Krynetsky, Phys. Lett. A 1991, l58, 491. [87] A. V. Zalesskii, A. A. Frolov, T. A. Khimich, A. A. Bush, Phys. Sol. St. 2003, 45, 141. [88] G. L. Yuan, S. W. Or, H. L. W. Chan, J. Phys. D: Appl. Phys. 2007, 40, 1196. [89] J.-P. Rivera, H. Schmid, Ferroelectrics 1997, 204, 23. [90] J. C. Toledano, Annal. Telecomm. 1974, 29, 249. [91] J. Holakovsky, F. Smutny, J. Phys. C: Solid State 1978, 11, L611. [92] P. Ravindran, R. Vidya, A. Kjekshus, H. Fjellvag, O. Eriksson, Phys. Rev. B 2006, 74, 224412. [93] T. T. Carvalho, European Summer School of Multiferroics, Girona, 2008. [94] J. Rymarczyk, D. Machura, J. Ilczuk, Eur. Phys. J. Special Topics 2008, 154, 191. [95] T. P. Gujar, V. R. Shinde, C. D. Lokhande, Mater. Chem. Phys. 2007, 103, 142. [96] S. Kamba, D. Nuzhnyy, M. Savinov, J. Sebek, J. Petzelt, J. Prokleska, R. Haumont, J. Kreisel, Phys. Rev. B 2007, 75, 024403.

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2483

www.advmat.de

[97] V. Fruth, E. Tenea, M. Gartner, M. Anastasescu, D. Berger, R. Ramer, M. Zaharescu, J. Eur. Ceram. Soc. 2007, 27, 937. [98] J. F. Ihlefeld, N. J. Podraza, Z. K Liu, R. C. Rai, X. Xu, T. Heeg, Y. B. Chen, J. Li, R. W. Collins, J. L. Musfeldt, X. Q. Pan, J. Schubert, R. Ramesh, D. G. Schlom, Appl. Phys. Lett. 2008, 92, 142908. [99] Y. Xu, M. Shen, Mat. Lett. 2008, 62, 3600. [100] S. J. Clark, J. Robertson, Appl. Phys. Lett. 2007, 90, 132903. [101] A. G. Gavriliuk, V. V. Struzhkin, I. S. Lyubutin, M. Y. Hu, H. K. Mao, JETP Lett. 2005, 82, 224. [102] C. Blaauw, F. van der Woude, J. Phys. C: Solid State Phys. 1973, 6, 1422. [103] M. K. Singh, R. S. Katiyar, J. F. Scott, J. Phys. Condens. Mat. 2008, 20, 252203. [104] N. F. Mott, Rev. Mod. Phys. 1968, 40, 677. [105] P. Schlottmann, Phys. Rev. B 1992, 45, 5784. [106] J. B. Torrance, P. Lacorre, A. I. Nazzal, E. J. Ansaldo, C. Niedermayer, Phys. Rev. B 1992, 45, 8209. [107] P. C. Caneld, J. D. Thompson, S. W. Cheong, L. W. Rupp, Phys. Rev. B 1993, 47, 12357. [108] M. Medarde, J. Mesot, P. Lacorre, S. Rosenkranz, P. Fischer, K. Gobrecht, Phys. Rev. B 1995, 52, 9248. [109] C. Ederer, N. A. Spaldin, Phys. Rev. Lett. 2005, 95, 257601. [110] J. R. Teague, R. Gerson, W. J. James, Solid State Commun. 1970, 8, 1073. [111] V. V. Shvartsman, W. Kleemann, R. Haumont, J. Kreisel, Appl. Phys. Lett. 2007, 90, 172115. [112] a) J. Li, J. Wang, M. Wuttig, R. Ramesh, N. Wang, B. Ruette, A. P. Pyatakov, A. K. Zvezdin, D. Viehland, Appl. Phys. Lett. 2004, 84, 5261. b) G. Xu, H. Hiraka, G. Shirane, J. Li, J. Wang, D. Viehland, Appl. Phys. Lett. 2005, 86, 182905. c) S Keisuke, K. Saito, A. Ulyanenkov, V. Grossmann, H. Ress, L. Bruegemann, H. Ohta, T. Kurosawa, S. Ueki, H. Funakubo, Jpn. J. Appl. Phys. 2006, 45, 7311. d) G. Y. Xu, J. F. Li, D. Viehland, Appl. Phys. Lett. 2006, 89, 222901. [113] H. Bea, M. Bibes, S. Petit, J. Kreisel, A. Barthelemy, Philos. Mag. Lett. 2007, 87, 165. [114] D. H. Kim, H. N. Lee, M. D. Biegalski, H. M. Christen, Appl. Phys. Lett. 2008, 92, 012911. [115] Y. H. Chu, T. Zhao, M. P. Cruz, Q. Zhan, P. L. Yang, L. W. Martin, M. Huijben, C. H. Yang, F. Zavaliche, H. Zheng, R. Ramesh, Appl. Phys. Lett. 2007, 90, 252906. [116] F. He, B. O. Wells, S. M. Shapiro, Phys. Rev. Lett. 2005, 94, 176101. [117] G. Catalan, A. Janssens, G. Rispens, S. Csiszar, O. Seeck, G. Rijnders, D. H. Blank, B. Noheda, Phys. Rev. Lett. 2006, 96, 127602. [118] K. J. Choi, M. Biegalski, Y. L. Li, A. Sharan, J. Schubert, R. Uecker, P. Reiche, Y. B. Chen, X. Q. Pan, V. Gopalan, L.-Q. Chen, D. G. Schlom, C. B. Eom, Science 2004, 306, 1005. [119] J. H. Haeni, P. Irvin, W. Chang, R. Uecker, P. Reiche, Y. L. Li, S. Choudhury, W. Tian, M. E. Hawley, B. Craigo, A. K. Tagantsev, X. Q. Pan, S. K. Streiffer, L. Q. Chen, S. W. Kirchoefer, J. Levy, D. G. Schlom, Nature (London) 2004, 430, 758. [120] H. W. Jang, S. H. Baek, D. Ortiz, C. M. Folkman, R. R. Das, Y. H. Chu, P. Shafer, J. X. Zhang, S. Choudhury, V. Vaithyanathan, Y. B. Chen, D. A. Felker, M. D. Biegalski, M. S. Rzchowski, X. Q. Pan, D. G. Schlom, L. Q. Chen, R. Ramesh, C. B. Eom, Phys. Rev. Lett. 2008, 101, 107602. [121] C. W. Bark, S. Ryu, Y. M. Koo, H. M. Jang, H. S. Youn, Appl. Phys. Lett. 2007, 90, 022902. [122] J. X. Zhang, Y. L. Li, Y. Wang, Z. K. Liu, L. Q. Chen, Y. H. Chu, F. Zavaliche, R. Ramesh, J. Appl. Phys. 2007, 101, 114105. [123] J. Zhao, A. E. Glazounov, Q. M. Zhang, B. Toby, Appl. Phys. Lett. 1998, 72, 1048. [124] W. Kaczmarek, Z. Pajak, M. Polomska, Solid State Commun. 1975, 17, 807. [125] R. Mazumder, S. Ghosh, P. Mondal, D. Bhattacharya, S. Dasgupta, N. Das, A. Sen, A. K. Tyagi, M. Sivakumar, T. Takami, H. Ikuta, J. Appl. Phys. 2006, 100, 033908.

REVIEW

[126] [127] [128] [129] [130] [131] [132] [133] [134] [135] [136] [137] [138] [139] [140] [141] [142] [143] [144]

[145] [146] [147] [148] [149] [150]

[151] [152] [153] [154] [155] [156] [157] [158] [159] [160] [161]

[162]

G. Catalan, J. F. Scott, Nature 2007, 448, E4E5. G. Catalan, Appl. Phys. Lett. 2006, 88, 102902. J. F. Scott, JETP Lett. 1989, 49, 233. D. Lebeugle, D. Colson, A. Forget, M. Viret, A. M. Bataille, A. Gukasov, Phys. Rev. Lett. 2008, 100, 227602. I. Sosnovska, T. Peterlin-Neumaier, E. Steichele, J. Phys. C 1982, 15, 4835. J. F. Scott, M. K. Singh, R. S. Katiyar, J. Phys.: Condens. Matter 2008, 20, 322203. J. F. Scott, M. K. Singh, R. S. Katiyar, J. Phys.: Condens. Matter 2008, 20, 425223. A. Palewicz, T. Szumiata, R. Przenioso, I. Sosnowska, I. Margiolaki, Solid State Commun. 2006, 140, 359. R. Przenioso, M. Regulski, I. Sosnowska, J. Phys. Soc. Jpn. 2006, 75, 084 718. R. Przenioso, A. Palewicz, M. Regulski, I. Sosnowska, R. M. Ibberson, K. S. Knight, J. Phys.: Condens. Matter 2006, 18, 2069. A. A. Bush, A. A. Gippius, A. V. Zalesskii, E. N. Morozova, JETP Lett. 2003, 78, 389. A. V. Zalesskii, A. A. Frolov, A. K. Zvezdin, A. A. Gippius, E. N. Morozova, D. F. Khozeev, A. A. Bush, V. S. Pokatilov, JETP 2002, 95, 101. A. V. Zalesskii, A. K. Zvezdin, A. A. Frolov, A. A. Bush, JETP Lett. 2000, 71, 465. H. Schmid, J. Phys.: Condens. Matter 2008, 20, 434 201. S. Lee, W. Ratcliff, S.-W. Cheong, V. Kiryukhin, Appl. Phys. Lett. 2008, 92, 192 906. M. Cazayous, Y. Gallais, A. Sacuto, R. de Sousa, D. Lebeugle, D. Colson, Phys. Rev. Lett. 2008, 101, 037 601. S. A. T. Redfern, C. Wang, J. W. Hong, G. Catalan, J. F. Scott, J. Phys.: Cond. Mat. 2008, 20, 452205. M. K. Singh, W. Prelier, M. P. Singh, R. S. Katiyar, J. F. Scott, Phys. Rev. B 2008, 77, 144403. R. M. White, R. J. Nemanich, Conyers. Herring, Phys. Rev. B 1982, 25, 1822. S. Venugopalan, M. Dutta, A. K. Ramdas, J. P. Remeika, Phys. Rev. B 1985, 31, 1490. N. Koshizuka, S. Ushioda, Phys. Rev. B 1980, 22, 5394. C. Ederer, N. A. Spaldin, Phys. Rev. B 2005, 71, 060401R. A. Bombik, B. Lesniewska, J. Mayer, A. W. Pacyna, J. of Mag. Mag. Mat. 2003, 257, 206. T. Yamaguchi, J. Phys. Chem. Solids. 1974, 35, 479. M. K. Singh, R. S. Katiyar, W. Prelier, J. F. Scott, J. Phys.: Condens. Matter 2009, 21, 042202. A. K. Pradhan, Kai Zhang, D. Hunter, J. B. Dadson, G. B. Loiutts, P. Bhattacharya, R. Katiyar, J. Zhang, D. J. Sellmyer J. Appl. Phys. 2005, 97, 093903. S. Nakamura, S. Soeya, N. Ikeda, M. Tanaka, J. Appl. Phys. 1993, 74, 5652. S. Kirkpatrick, D. Sherrington, Phys. Rev. B 1978, 17, 4384. K. H. Fischer, J. Hertz, Spin Glasses, Cambridge Univ. Press, Cambridge 1991. K. De, M. Thakur, A. Manna, S. Giri, J. Appl. Phys. 2006, 99, 013 908. V. V. Shvartsman, R. Haumont, W. Kleemann, unpublished. J. R. L. de Almeida, D. J. Thouless, J. Phys. A 1978, 11, 983. S. T. Zhang, M. H. Lu, D. Wu, Y. F. Chen, N. B. Ming, Appl. Phys. Lett. 2005, 87, 262907. H. Naganuma, S. Okamura, J. Appl. Phys. 2007, 101, 09103. C. Tabares-Munoz, J.-P. Rivera, A. Bezinges, A. Monnier, H. Schmid, Jpn. J. Appl. Phys. 1985, 242, 1051. I. H. Ismailzade, R. M. Ismailov, A. I. Alekperov, F. M. Salaev, Phys. Stat. Sol. A 1980, 57, 99. a) Y. F. Popov, A. K. Zvezdin, G. P. Vorobev, A. M. Kadomtseva, V. A. Murashev, D. N. Rakov, JETP Lett. 1993, 57, 69. b) A. K. Zvezdin, A. M. Kadomtseva, S. S. Krotov, A. P. Pyatakov, Y. F. Popov, G. P. Vorobev, J. Magn. Magn. Mater. 2006, 300, 224. A. M. Kadomtseva, A. K. Zvezdin, Y. F. Popov, A. P. Pyatakov, G. P. Vorobev, JETP Lett. 2004, 79, 571.

2484

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2009, 21, 24632485

www.advmat.de

REVIEW

[163] I. Sosnowska, A. K. Zvezdin, J. Magn. Magn. Mater. 1995, 140, 167. [164] N. Wang, J. Cheng, A. Pyatakov, A. K. Zvezdin, J. F. Li, L. E. Cross, D. Viehland, Phys. Rev. B 2005, 72, 104434. [165] F. Bai, J. Wang, M. Wuttig, J. F. Li, N. Wang, A. P. Pyatakov, A. K. Zvezdin, L. E. Cross, D. Viehland, Appl. Phys. Lett. 2005, 86, 32511. [166] T. Zhao, A. Scholl, F. Zavaliche, K. Lee, M. Barry, A. Doran, M. P. Cruz, Y. H. Chu, C. Ederer, N. A. Spaldin, R. R. Das, D. M. Kim, S. H. Baek, C. B. Eom, R. Ramesh, Nat. Mater. 2006, 5, 823. [167] S. K. Streiffer, C. B. Parker, A. E. Romanov, M. J. Lefevre, L. Zhao, J. S. Speck, W. Pompe, C. M. Foster, G. R. Bai, J. Appl. Phys. 1998, 83, 2742. [168] L. D. Landau, E. M. Lifshitz, Phys. J. Sowjetunion 1935, 8, 153. [169] C. Kittel, Phys. Rev. 1946, 70, 965. [170] T. Mitsui, J. Furuichi, Phys. Rev. 1953, 90, 193. [171] A. L. Roytburd, Phys. Stat. Sol. A 1976, 37, 329. [172] M. Daraktchiev, G. Catalan, J. F. Scott, Ferroelectrics 2008, 375, 122. [173] G. Catalan, A. Schilling, J. F. Scott, J. M. Gregg, J. Phys.: Cond. Mat. 2007, 19, 132201. [174] Y. B. Chen, M. B. Katz, X. Q. Pan, R. R. Das, D. M. Kim, S. H. Baek, C. B. Eom, Appl. Phys. Lett. 2007, 90, 072907. [175] G. Catalan, H. Bea, S. Fusil, M. Bibes, P. Paruch, A. Barthelemy, J. F. Scott, Phys. Rev. Lett. 2008, 100, 027602. [176] 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, H. Lichte, F. Wang, G. Catalan, J. F. Scott, N. A. Spaldin, J. Orenstein, R. Ramesh, Nat. Mater. 2009, 8, 229. [177] J. Privratska, V. Janovec, Ferroelectrics 1997, 204, 321. [178] A. Goltsev, R. Pisarev, T. Lottermoser, M. Fiebig, Phys. Rev. Lett. 2003, 90, 177204. [179] T. Lottermoser, M. Fiebig, Phys. Rev. B 2004, 70, 220407R. [180] T.-J. Park, G. C. Papaefthymiou, A. J. Viescas, A. R. Moodenbaugh, S. S. Wong, Nano Lett. 2006, 7, 766. [181] F. Gao, Y. Yuan, K. F. Wang, X. Y. Chen, F. Chen, J.-M. Liu, Z. F. Ren, Appl. Phys. Lett. 2006, 89, 102506. [182] a) J. T. Richardson, W. O. Milligan, Phys. Rev. 1956, 102, 1289. b) J. T. Richardson, D. I. Yiagas, B. Turk, K. Forster, M. V. Twigg, J. Appl. Phys. 1991, 70, 6977. [183] T. J. Park, Y. B. Mao, S. S. Wong, Chem. Commun. 2004, 2708. [184] X. Y. Zhang, C. W. Lai, X. Zhao, D. Y. Wang, J. Y. Dai, Appl. Phys. Lett. 2005, 87, 143102. [185] M. D. Glinchuk, E. A. Eliseev, A. N. Morozovska, R. Blinc, Phys. Rev. B 2008, 77, 024106.

[186] Fujitsu website: http://www.fujitsu.com/my/news/pr/fmal20060808. html (accessed January 2009). [187] H. W. Jang, S. H. Baek, D. Ortiz, C. M. Folkman, C. B. Eom, Y. H. Chu, P. Shafer, R. Ramesh, V. Vaithyanathan, D. G. Schlom, Appl. Phys. Lett. 2008, 92, 062910. [188] D. I. Woodward, I. M. Reaney, R. E. Eitel, C. A. Randall, J. Appl. Phys. 2003, 94, 3313. [189] a) B. Noheda, D. E. Cox, G. Shirane, J. A. Gonzalo, L. E. Cross, S. E. Park, Appl. Phys. Lett. 1999, 74, 2059. b) R. Guo, L. E. Cross, S.-E. Park, B. Noheda, D. E. Cox, G. Shirane, Phys. Rev. Lett. 2000, 84, 5423. [190] S. Fujino, M. Murakami, V. Anbusathaiah, S.-H. Lim, V. Nagarajan, C. J. Fennie, M. Wuttig, L. Salamanca-Riba, I. Takeuchi, Appl. Phys. Lett. 2008, 92, 202904. [191] K. Takahashi, N. Kida, M. Tonouchi, Phys. Rev. Lett. 2006, 96, 117402. [192] V. Ryzhii, J. Phys.: Condens. Matter 2008, 20, 380301. [193] H. Bea, M. Gajek, M. Bibes, A. Barthelemy, J. Phys.: Cond. Mat. 2008, 20, 434231. [194] J. Nogues, I. K. Schuller, J. Magn. Magn. Mat. 1999, 192, 203. [195] P. Borisov, A. Hochstrat, X. Chen, W. Kleemann, C. H. Binek, Phys. Rev. Lett. 2005, 94, 117203. , [196] V. Laukhin, V. Skumryev, X. Mart D. Hrabovsky, F. Sanchez, M. V. a-Cuenca, C. Ferrater, M. Varela, U. Luders, J. F. Bobo, J. Fontcuberta, Garc Phys. Rev. Lett. 2006, 97, 227201. [197] J. Dho, X. Qi, H. Kim, J. L. MacManus-Driscoll, M. G. Blamire, Adv. Mater. 2006, 18, 1445. [198] H. Bea, M. Bibes, S. Cheri, F. Nolting, B. Warot-Fonrose, S. Fusil, G. Herranz, C. Deranlot, E. Jacquet, K. Bouzehouane, A. Barthelemy, Appl. Phys. Lett. 2006, 89, 242114. [199] H. Bea, M. Bibes, F. Ott, B. Dupe, X.-H. Zhu, S. Petit, S. Fusil, C. Deranlot, K. Bouzehouane, A. Barthelemy, Phys. Rev. Lett. 2008, 100, 017204. [200] Y.-H. Chu, L. W. Martin, M. B. Holcomb, M. Gajek, S.-J. Han, Q. He, N. guez, Balke, C.-H. Yang, D. Lee, W. Hu, Q. Zhan, P.-L. Yang, A. Fraile-Rodr A. Scholl, S. X. Wang, R. Ramesh, Nat. Mater. 2008, 7, 478. [201] H. Bea, S. Fusil, K. Bouzehouane, M. Bibes, M. Sirena, G. Herranz, E. Jacquet, J.-P. Contour, A. Barthelemy, Jpn. J. Appl. Phys. 2006, L187, 45. [202] M. Gajek, M. Bibes, S. Fusil, K. Bouzehouane, J. Fontcuberta, A. Barthelmy, A. Fert, Nat. Mater. 2007, 6, 296. e [203] M. Bibes, A. Barthelemy, Nat. Mater. 2008, 7, 425. [204] N. A. Halasa, G. DePasquali, H. G. Drickamer, Phys. Rev. B 1974, 10, 154. [205] T. Choi, S. Lee, Y. J. Choi, V. Kiryukhin, S.-W. Cheong, Sciencexpress, DOI: 10.1126/science.1168636 (2009).

Adv. Mater. 2009, 21, 24632485

2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2485

You might also like