Ferroelectric interfacial devices consist of materials systems whose interfacial electronic properties (such as a 2D electron gas or an interfacial magnetic spin configuration) are modulated by a ferroelectric layer set in its immediate vicinity. While the prototypical example of such a system is the ferroelectric field effect transistor first proposed in the 1950s, only with the recent advances in the controlled growth of epitaxial thin films and heterostructures, and the recent physical understanding down to the atomic scale of screening processes at ferroelectric-semiconducting and -metallic interfaces made possible by first principles calculations, have the conditions been met for a full development of the field. In this review, we discuss the recent advances in ferroelectric interfacial systems with emphasis on the ferroelectric control of the electronic properties of interfacial devices with well ordered (epitaxial) interfaces. In particular, we consider the cases of ferroelectric interfacial systems aimed at controlling the correlated state, including superconductivity, Mott metallic-insulator transition, magnetism, charge, and orbital order, and charge and spin transport across ferroelectric tunnel junctions. The focus is on the basic physical mechanisms underlying the emergence of interfacial effects, the nature of the ferroelectric control of the electronic state, and the role of extreme electric field gradients at the interface in giving rise to new physical phenomena. Such understanding is key to the development of ferroelectric interfacial systems with characteristics suitable for next generation electronic devices based on controlling the correlated state of matter.

The current interest in interface phenomena stems from the desire to discover and explore new physical effects emerging at the immediate atomic region forming the boundary between two different materials, in particular, since the change in crystal symmetry allows new physical phenomena that are distinct from those of the respective bulk phases. As a consequence, modifications in the electronic structure, which result from electron transfer or charge confinement effects, can lead to the emergence of novel effects and to new functionalities that are intrinsic to the interface.1,2 An important class of interfacial systems utilizes a ferroelectric material to modulate the electronic properties of the interfacial system through the ferroelectric polarization. We refer to such device structures as ferroelectric interfacial devices in order to emphasize the fact that such phenomena are characteristic of the interface, with the latter acquiring its own physical, chemical, and electronic intrinsic properties. Two different types of device structures can be distinguished: one based on the ferroelectric field effect approach used to induce large and switchable modulations in the charge carrier density at the interface through charge screening effects,3–5 and the other based on ferroelectric tunnel junctions, where the changes in the electric potential as a function of the direction of the ferroelectric polarization give rise to large changes in the transport characteristics across the ferroelectric used as a tunnel barrier.6,7 Important physical requirements for both types of device structures, include sharp interfaces for optimal effects, robust ferroelectric effects at ultrathin thicknesses, and precise control of the channel and ferroelectric layer thickness for ferroelectric field effects and tunneling devices, respectively. Under such conditions, the ferroelectric control of several interfacial physical phenomena have been achieved, including interfacial magnetism, superconductivity, and orbital order in ferroelectric field effect devices, and of the tunneling resistance in various ferroelectric tunnel junction architectures, as illustrated in Fig. 1. The control of those physical phenomena involve a laborious identification of the mechanisms linking the ferroelectric order parameter to that of the interfacing material, often a joint experimental and theoretical endeavor. The progress achieved over the last two decades in our understanding of thin film ferroelectricity8–14 and of interface phenomena,1,2,15–28 from both practical and fundamental perspectives, has been tremendous. The aim of this review is to take stock of the recent developments in the field to gauge the challenges for further progress in view of potential technological applications. We pay particular attention to the aspects associated with ultrathin ferroelectricity, since ferroelectric films are one of the fundamental building blocks of interfacial ferroelectric devices, as well as to the physical mechanisms underlying the emergence and control of the interfacial effects between the ferroelectric and the other material of interest.

FIG. 1.

Schematic diagram showing various types of control of the correlated state in interfacial ferroelectric (FE) devices, including control of superconductivity, Mott metal to insulator (MIT) transition, tunneling conductivity, charge, orbital, and magnetic order.

FIG. 1.

Schematic diagram showing various types of control of the correlated state in interfacial ferroelectric (FE) devices, including control of superconductivity, Mott metal to insulator (MIT) transition, tunneling conductivity, charge, orbital, and magnetic order.

Close modal

Ferroelectric materials are characterized by the presence of a spontaneous electric polarization, whose direction can be switched between states that are identical in the crystal structure (enantiomorphous) by means of an applied electric field.29,30 Ferroelectricity involves space inversion symmetry breaking, and the spontaneous ferroelectric polarization can originate from a number of different microscopic mechanisms, including displacement of ions from high symmetry positions, electronic charge asymmetries (as in lone-pair ferroelectricity), charge ordering and/or disproportionation, and geometrical frustration.9,31–40 Hence, ferroelectric phenomena tend to be more challenging to investigate than their magnetic counterparts,35,41 also due to the strong interaction with the electric fields generated by free electric charges and the strong coupling to lattice distortions, including strain, which often make modeling and interpretation of the experimental results difficult. Theoretically also, only recently a consistent theory of the ferroelectric polarization42–47 (expressed in terms of a Berry phase)48 and the development of ab initio methods to treat the electric field termination of ferroelectric slabs9,31,34 have been made available. From a materials' perspective, ferroelectric systems tend to be multicomponent systems, typically containing three or more elements, with a concomitantly complex defect chemistry that make the preparation of good quality, pure samples difficult, but a task made possible in many instances by the advances in the controlled epitaxial growth of thin ferroelectric films and multilayer heterostructures. The ability to grow high quality thin ferroelectric films is important not only for fundamental studies but also for device applications, where size scaling and a long term retention of the ferroelectric properties at small lateral dimensions and reduced thickness are important requirements, and because the onset of interfacial phenomena between dissimilar materials and the control of those properties by means of external stimuli, such as electric fields, require well defined and abrupt interfaces. In this context, state-of-the-art techniques for ferroelectric thin film growth and optimization of the growth process have resulted in the routine fabrication of epitaxial films of very high structural quality, as we illustrate later for a few prototypical ferroelectric systems (Sec. II B). Also significant has been the development of new experimental techniques dedicated to the study of nanoscale ferroelectricity and interfacial phenomena (Sec. II A); in particular, the development of continuous and operando modes of measurement, where the physical properties of the device are probed while subject to a varying external stimulus, are particularly appealing, since they allow one to follow the changes in physical properties in real time while minimizing the influence of other parameters. The combined progress achieved in these theoretical and experimental fronts has led to a strong increase in interest and in the understanding of physical phenomena in ferroelectrics.

Of almost equally importance as the intrinsic properties of ferroelectrics are the conditions under which the ferroelectric system is physically and electrically connected, particularly for thin films, where such boundary conditions have a tremendous impact on the ferroelectric properties. As illustrated in Fig. 2, both three-terminal ferroelectric field effect devices and two-terminal ferroelectric tunnel junctions require a supporting substrate and electrical connections to the ferroelectric layer and to the interface. Since ferroelectricity is generally linked to local structural distortions in the atomic lattice, the constraints imposed by the substrate where the ferroelectric film is grown strongly impact its ferroelectric properties, including ferroelectric polarization, critical temperature, and piezoelectric response. Electrical boundary conditions are also present in terms of metallic contacts to the ferroelectric film, both to switch the ferroelectric state of the device and to establish the functional interface, for example, with a superconductor in a field effect device or with a ferromagnet in a multiferroic ferroelectric tunnel junction. The metallic electrode acts as a charge reservoir that screens the electric polarization at the dielectric interface. In fact, the dielectric/metallic interface has since long received considerable attention since it constitutes the basis of the capacitor structure through which the properties of dielectrics are often investigated. Surprisingly, a strong dependence of the dielectric properties with the electrode material has been observed, in the extreme case leading to a strong reduction of the dielectric properties of the dielectric at reduced thicknesses, built-in electric fields, and fatigue.49–57 To a large extent, such phenomena are linked to the physical characteristics of the metal/ferroelectric interface, and the key aspects associated with this interface are discussed in Sec. II C.

FIG. 2.

Schematic of two and three terminal devices based on ferroelectrics and interfacial switching. (a) The three terminal device, which is similar to a transistor, consists of source and drain contacts to an active channel. The conductivity of the channel is strongly dependent on interface effects between the channel material and the ferroelectric polarization. (b) Tunnel transport in a two-terminal device through the ferroelectric is controlled by the interaction of the ferroelectric polarization and the top and bottom electrode.

FIG. 2.

Schematic of two and three terminal devices based on ferroelectrics and interfacial switching. (a) The three terminal device, which is similar to a transistor, consists of source and drain contacts to an active channel. The conductivity of the channel is strongly dependent on interface effects between the channel material and the ferroelectric polarization. (b) Tunnel transport in a two-terminal device through the ferroelectric is controlled by the interaction of the ferroelectric polarization and the top and bottom electrode.

Close modal

The prototypical example of a ferroelectric interfacial device is the ferroelectric field effect transistor, first proposed in the 1950s,58,59 before the inventions of the integrated circuit and SiO2 as the transistor gate dielectric. The goal of this invention is to combine the logic function of a CMOS transistor with the switching of the ferroelectric polarization to achieve nonvolatile on/off states. Two challenges need to be overcome in such structures: the control of the interface structure in order to avoid spurious screening of the ferroelectric surface polarization that may arise from impurities or charge traps from bandgap states and ensuring that the (semi)conducting channel can provide the requisite charge screening of the ferroelectric polarization to stabilize ferroelectricity and the orientation of the ferroelectric polarization. Such challenges prevented the immediate success of the ferroelectric field effect transistor (FeFET), but recent efforts have resulted in approaches to using ferroelectrics in fast and energy efficient memory devices.60 The FeFET concept remains very appealing and has been adapted to controlling the properties of other material systems susceptible to charge modulation [Fig. 1]. For example, instead of a semiconductor channel, large changes in the channel conductivity can be achieved by controlling the insulator to metal transition of a Mott insulator or the superconducting state of a high-critical temperature superconductor. Here, one controls not simply the conductivity through the charge carrier density but directly the correlated state via its strong dependence on the carrier density. This approach to controlling the correlated state of matter through changes in the charge carrier density using the ferroelectric field effect has proved to be a powerful method to observe new physical interfacial phenomena. That the processes are largely confined to the interface is a result of the relatively large charge density of strongly correlated electron systems and the correspondingly short Thomas-Fermi screening lengths of the order of the unit cell.3 By taking advantage of the very large surface bound charges that are characteristic of ferroelectrics, of up to one electron per unit cell, one can reach modulations in the interfacial carrier density well beyond what is possible with silicon oxide gates.3 The ferroelectric field effect approach has been used to electrostatically control the superconducting state, magnetism, the Mott metal to insulator (MIT) transition, and the orbital state, as we discuss in detail in Sec. III. The challenges for these more exotic field effect devices are increasing the size of the observed effects to above room temperature, as desirable for electronic devices.

While tunneling transport phenomena across dielectrics are now well established, the particular case when the dielectric is also a ferroelectric has been explored only recently, first theoretically,61–64 and later experimentally,65–68 to reveal surprisingly large changes in conductivity as a function of the ferroelectric polarization direction, in what is termed tunneling electric resistance (TER) in analogy with the magnetic counterpart, tunneling magnetoresistance (TMR).69,70 The experimental study of such effects is made possible by the availability of high quality, ultrathin ferroelectric films, as discussed in Sec. II. Several mechanisms are responsible for the TER effect, some related to the bulk of the ferroelectric film, including changes in thickness with the applied electric field, while other contributions are linked to modifications of the electronic potential and changes in bonding at the interface.6 We overview in Sec. IV the current status of the field.

An important motivation for the research work in this field is the prospect of applying the physical effects observed at ferroelectric interfaces to functional devices.71 While such systems offer good scaling properties,72 novel functionalities, and robust electric field control over electronic properties, the main challenges to realizing ferroelectric devices are the requirement for sustaining the ferroelectric state at ultrathin thicknesses for ferroelectrics of technological interest (which is presently more a practical rather than a fundamental issue) and of achieving sufficiently large interfacial electronic effects at ambient temperature. Also, rather than competing with binary logic operation, the class of multifunctional devices should aim for more complex operations, including devices with selective or self-enforcing memory suitable for simulating neural networks for artificial intelligence, or to control quantum entanglement for quantum computation.73 For example, while current artificial neural networks are based on classical computing, it can be anticipated that better scaling, speed, and lower power consumption of the millions of interconnected artificial neurons and synapses can be better achieved with memristor devices (systems showing hysteresis in the resistivity response) based on oxide-based material systems.74 In this context, resistive switching associated with reversible, redox processes that give rise to a nonvolatile change in resistivity with the applied electric field that develops in many thin films,75 including oxides, has been proposed as a potential building block for neural networks. Since they rely on filamentary disruption of the insulating barrier between two conducting electrodes, they can be easily scaled up; however, the control of their onset within a device and the spread in their characteristics remain a challenge. Another device structure capable of the same memristor function is the ferroelectric tunnel junction.76–78 Compatibility and integration of such types of devices with current CMOS technology would also be desirable.79 A brief overview of the proposals being put forward in this area is provided in Sec. V.

Given the size of the field, we restrict ourselves to discussing epitaxial systems, where better control over the interface enables a more targeted control of the electronic effects and minimizes extrinsic factors. In particular, our focus is on oxide materials, where the interfacial properties between ferroelectric oxides and complex oxides have been extensively studied and are better understood. Heterostructures consisting of two-dimensional (2D) materials interfaced with ferroelectrics are not discussed; the physical and chemical properties of 2D materials, such as graphene and the transition metal di- and trichalcogenides, have generated much interest due to the wide range of properties exhibited by these systems, including pseudo-spin in the case of graphene,80,81 magnetism, ferroelectricity, and optical properties in the chalcogenides.82–84 An increasing body of work has focused on interfacing such materials with ferroelectrics, which has been the focus of several recent review articles.85–92 The main approach in such device structures has been to modulate the charge carrier density via electrostatic doping, and a significant effort has been placed at forming a clean interface between these systems and the ferroelectric to achieve the expected response to the ferroelectric charge modulation.93–95 In these systems, epitaxy does not seem critical to device function, which remains at the level of the 2D system itself.

Our goals in this paper are to provide both a timely appraisal of the current developments on ferroelectric interfacial devices and to stimulate further work by pointing, in our view, to promising approaches, as we discuss succinctly in the conclusions and outlook (Sec. VI). Our overview highlights that the large practical and theoretical efforts in fabricating and understanding thin film ferroelectricity have led to a good overall control of the properties of ferroelectric thin films and to the discovery of new ferroelectric systems with enhanced properties that are imminently suited to device applications; in turn, the growth of high quality ferroelectric/complex oxide interfaces have enabled the discovery and study of new interfacial phenomena that have the potential for new functional devices.

The experimental investigation of ferroelectric phenomena at the nanoscale requires techniques with sufficient sensitivity and spatial resolution to measure the physical properties of systems at reduced volumes. Measurements of the dielectric and ferroelectric response to electric stimuli, such as permittivity, dielectric loss, breakdown voltage, ferroelectric polarization, piezoelectric response, and ferroelectric switching, are often carried out in a parallel plate capacitor geometry with metallic electrodes in direct contact to the ferroelectric system. In this geometry, a time-varying, spatially uniform electric field is applied to the ferroelectric system across the metallic plates and the displacement current is measured as a function of time to yield the dielectric constant for small electric field excitations or the ferroelectric polarization by integration of the displacement current associated with the ferroelectric polarization switching. Commercial ferroelectric testers can be used to obtain polarization vs electric field hysteresis curves that provide not only measurements of the coercive field, saturation, and remanent polarization but also convey information about the ferroelectric switching process, leakage current, built-in electric fields at the contact interfaces, and fatigue (i.e., loss in switchable polarization upon repeated switching) due to defects. Alternatively, the ferroelectric polarization can be obtained through the pyroelectric effect by measuring the charge required to compensate the onset of the ferroelectric polarization when cooling the sample down from the paraelectric state.96–98 In addition, to probe nanoscale volumes and ferroelectric dynamics, the size of the contacts can be defined to be arbitrary small, for example, down to the nanoscale by means of advanced e-beam lithography techniques and on ultrathin ferroelectric films. However, while conceptually simple, such types of electrical measurements are challenging for a number of reasons: in addition to the small magnitude of the signal, tunneling currents in ultrathin films may overwhelm the ferroelectric displacement currents, and the impact of the metallic electrical contacts needs to be taken into consideration in order not to mask the intrinsic dielectric properties of the ferroelectric system.99,100 For example, for PZT, it is found that Cu, Au, Ag, and Pd are suitable materials that form Schottky contacts, while Ta and Cr can form Ohmic contacts unsuitable for electric characterization.101–104 Since the impact of the electric contact/ferroelectric interface is fundamental for ferroelectric field effect devices, we will discuss this aspect in more detail in Sec. II C 2.

A second type of measurement probes local ferroelectric phenomena at the nano- or atomic scale and employs techniques that have high spatial resolution. One such technique is piezoresponse force microscopy (PFM),105–108 a scanning probe technique that relies on the measurement of the sample piezoelectric response to an alternating electric field applied locally through a conducting tip of an atomic force microscope (in contact mode) and an extended bottom electric contact (typically, a conducting substrate or a conduction layer grown between the substrate and the ferroelectric film). At an elementary level, the amplitude of the response provides a direct measure of the local piezoresponse amplitude, while the phase signal provides direct information about the orientation of the ferroelectric domain (in-phase when the polarization points opposite to the electric field and 180° for the opposite direction). While a powerful technique for obtaining spatially resolved maps of the ferroelectric domain structure and of the local piezoelectric response, instrumental artifacts, such as local charge injection from the metal tip to the ferroelectric film surface, can make interpretation of PFM results difficult.109 Also, quantitative values for the piezoresponse of thin ferroelectric films can be challenging to extract due to the small vertical displacements induced by the electric field and due to the highly non-homogeneous electric field generated by the sharp metallic tip, and multiple strategies have been developed for overcoming such difficulties.110–116 With the same proviso, local electric fields can be applied to the sample to directly visualize the ferroelectric switching process.117,118 Another technique that has been used to visualize ferroelectric domains is photoemission electron microscopy (PEEM), which measures the intensity of photoemitted electrons excited by a light source, such as ultra-violet light, monochromatic electrons, or x-rays.119–122 Since the photoemitted electron intensity depends strongly on the surface potential, differences in the latter associated with the ferroelectric polarization of the different domains at the sample surface can be measured to provide direct images of the ferroelectric domains.121,123–126 For similar reasons, low energy electron diffraction (LEED) can be used also as a tool to derive the surface potential of ferroelectric surfaces, although without the spatial resolution.56 

Techniques based on x-rays, and synchrotron x-ray light in particular, are also powerful probes for ferroelectric properties, providing precise measurements of the local atomic structure, including changes in atomic displacements, periodicity of ferroelectric domains, and ferroelectric switching, as a function of electric field, temperature, and pressure via x-ray diffraction and scattering;127–129 of the electronic state and band alignment via x-ray photoemission spectroscopy (XPS)130–134 and of the electronic band structure using x-ray absorption spectroscopy (XAS).135–141 With the advent of free electron laser sources142 capable of producing extremely intense, highly transverse coherent, and ultra-short (few fs long) pulsed x-ray beams, it is now possible to investigate ultrafast dynamical processes in ferroelectric systems, including phonon dynamics and ultrafast switching induced by strong THz light beams.143 

Finally, we mention here high resolution scanning transmission electron microscopy (HR-STEM), which is a particularly powerful technique for probing the atomic structure in real space of both ferroelectric films and interfaces and also capable of locally probing some aspects of the electronic structure via electron energy loss spectroscopy (EELS).28,144 The direction of the ferroelectric polarization can be determined from an accurate measurement of the relative atomic displacement between cation and anion species, such that ferroelectric domains and domain walls can be identified;145–147 also from the amplitude of the atomic displacement, estimates of the ferroelectric polarization can be obtained from the relation between effective Born charges and ferroelectric polarization.147–149 In many systems, local electrical contacts to the sample enable one to carry out in situ ferroelectric switching studies,146 including ferroelectric domain wall dynamics.145,150–152 Atomically resolved EELS spectra can provide additional information on the interface atomic structure (including the presence of atomic interdiffusion),153–155 valence profile across the interface,147,156–161 and atomic layer stacking across the interface.162 

A challenging aspect in the study of thin film ferroelectricity relates to the preparation of such systems, typically achieved by means of film growth techniques such as molecular beam epitaxy (MBE), pulsed layer deposition (PLD), and rf magnetron sputtering.163–167 In such deposition techniques, the film grows on a supporting substrate by condensation of the various elements, or molecules, that compose the material. To promote crystalline order, the growth is carried out at elevated temperature, such that atomic and molecular surface diffusion is high but bulk mass mobility is low.168 Often various stable crystalline phases exist for the same material composition or for a subset of the constituent elements, such that spurious phases may develop during growth which can, in some instances, strongly impact the materials properties or lead to an inconclusive interpretation of the experimental findings. A related aspect is that, in the majority of cases, the window in parameter space for the growth of the material of interest is relatively narrow, making the growth of high quality epitaxial films a demanding task that requires a good control over stoichiometry and oxidation state; in addition, a low density of defects, low surface and interface roughness, and the avoidance of spurious phases should be attained as well.

For ferroelectric interface devices, it is important not only that the ferroelectric interface be sharp but also that the ferroelectric properties are robust, including having a high spontaneous polarization, a high ferroelectric critical temperature, a large breakdown voltage, and that the system can be grown on suitable substrates as high quality thin films. A number of ferroelectric systems fulfill these conditions and have been grown as high quality single crystalline films, a representative subset of which are discussed next and whose bulk properties are listed in Table I.

TABLE I.

Bulk properties of selected ferroelectric materials at room temperature; for HfO2, only thin films have been stabilized. Values from Ref. 169 except where mentioned otherwise. Structure: T: tetragonal, R: rombohedral, O: orthorhombic. Tc: ferroelectric critical temperature, Ec: coercive field, Eg: bandgap, χ: electron affinity, ϵr: relative dielectric constant ϵ33/ϵ0, and ϵ: relative electronic dielectric constant (taken as the square of the ordinary refractive index no at 633 nm).170 

MaterialStructureLatt. param. (Å)Tc (K)Ps (μC/cm2)Ec (kV/cm)Eg (eV)χ (eV)ϵrϵ
BaTiO3 {a=3.9920c=4.0361 393 25 1.2 3.2a 3.8l 130 5.79h 
PbTiO3 {a=3.9072c=4.1187 763 75 6.75 3.2a 2.8m 80b 7.16h 
Pb(Zr0.2Ti0.8)O3 {a=3.933c=4.134 730d 70d 100d 3.4a 3.5g 90n 6.30j 
BiFeO3 {a=5.616α=59.35° 1123 100f 10f 2.8a 4.0o 65 8.82i 
HfO2 {a=5.07b=4.88c=4.89eee 723c 15-45c 1000c 5.5c 1.7p 23c 3.98k 
MaterialStructureLatt. param. (Å)Tc (K)Ps (μC/cm2)Ec (kV/cm)Eg (eV)χ (eV)ϵrϵ
BaTiO3 {a=3.9920c=4.0361 393 25 1.2 3.2a 3.8l 130 5.79h 
PbTiO3 {a=3.9072c=4.1187 763 75 6.75 3.2a 2.8m 80b 7.16h 
Pb(Zr0.2Ti0.8)O3 {a=3.933c=4.134 730d 70d 100d 3.4a 3.5g 90n 6.30j 
BiFeO3 {a=5.616α=59.35° 1123 100f 10f 2.8a 4.0o 65 8.82i 
HfO2 {a=5.07b=4.88c=4.89eee 723c 15-45c 1000c 5.5c 1.7p 23c 3.98k 
a

Reference 171.

b

Reference 172.

c

Reference 173.

d

Reference 174.

e

First principles calculations, Ref. 175.

f

Reference 176.

g

Reference 177.

h

Reference 178.

i

Reference 179.

j

Reference 180.

k

References 181 and 182.

l

Reference 183.

m

Reference 184.

n

References 185 and 186.

o

Reference 187.

p

Reference 188. For HfO2/Si, using the electron affinity of Si as 4.05 eV.189 

1. BaTiO3

Among the ferroelectrics, barium titanate (BaTiO3) has played an important historical role in the study of ferroelectricity, since its crystal structure with a small unit cell permitted a more straightforward understanding of the origin of the spontaneous electric polarization in the solid state as compared to the earlier known ferroelectrics, such as the Rochelle salt with a large unit cell;29 it is, in addition, a technologically important ferroelectric due to its robust ferroelectric and optoelastic properties.190–193 The high-temperature paraelectric phase of BaTiO3 (between 1733 and 393 K) crystallizes in the centrosymmetric cubic perovskite structure; at lower temperatures between 393 and 278 K, the active soft-phonon mode associated with the relative displacement of the Ti cation with respect to the oxygen octahedron freezes, leading to the onset of a ferroelectric dipole moment in a tetragonal crystal structure, with the polarization pointing along the c axis. At lower temperatures (between 278 and 183 K), further distortion in the crystal structure to orthorhombic takes place along with a change in the direction of the ferroelectric polarization to the pseudocubic [110]pc direction; below 183 K, the structure changes further to rhombohedral with the polarization pointing along the pseudocubic [111] direction.169,194

Although high quality BaTiO3 single crystals are commercially available, its high cost, its sensitivity to variations in temperature due to the tetrahedral to orthorhombic phase transition at 278 K, and the long switching times and high voltages required to excite bulk crystals has driven the need for the growth of high quality epitaxial films on standard substrate materials for use in device applications. Epitaxial thin films of BaTiO3 have been grown by MBE, PLD, and sputtering on various substrates, including metal oxides such as SrTiO3(001),49,195–214 MgO(001),196,215–219 GdScO3(110),218,220 DyScO3(110),218 LaAlO3(100),217 and MgAl2O4(001)221 as well as on semiconductors, including Si(001)222–229 and Ge(001).230–232 In most instances, an intermediate buffer layer is used as an electrical contact or for the release of misfit strain. BaTiO3 films are typically grown at elevated temperatures, well above the critical temperature, such that they are brought down from the paraelectric state to the ferroelectric state at room temperature upon cooling. The most relevant film orientation for BaTiO3 in ferroelectric interfacial devices is the [001], i.e., with the polar axis pointing along the out-of-plane direction. It is worth pointing out that for the ferroelectric perovskites, the value of the c/a ratio tends to relate to the amplitude of the electric dipole and, therefore, to the amplitude of the ferroelectric polarization (Sec. II C).198,233–240 Hence, a substrate of high interest is SrTiO3, where a lattice mismatch of –2.18% is expected to lead to a strong compressive strain and to a large c/a ratio.236 On SrTiO3(001), BaTiO3 is found to grow layer by layer in steps of one unit cell (u.c.) up to about 12 u.c., above which the film growth proceeds in a three-dimensional island (Stranski–Krastanov) growth mode;195,200,201,205,208,241 at substrate temperatures of the order of 680 °C, typical for oxide film growth, the adatom surface mobility is high, as deduced from the recovery of the RHEED intensity after a break in the deposition process of about 20 s.195,202,205 However, such adatom surface mobility seems to be detrimental to the growth of smooth and fully strained films: the latter can be achieved at lower deposition temperatures and higher deposition rates, while under opposite conditions early strain relaxation and 3D growth are observed.204,242,243 For the BaTiO3/SrTiO3(001) system, the critical thickness for coherent growth (coherent films have the same in-plane lattice parameter and a symmetry compatible with the substrate) has been found to range from 2 to 10 nm, depending on the growth conditions, above which the in-plane lattice constant starts relaxing toward the bulk value through the onset of misfit dislocations.201,205,207,214,243,244 Although Nb-doped conducting SrTiO3 can be used as a substrate for the BaTiO3 film growth also to form a bottom contact, an alternative solution consists of using a conducting buffer layer, such as LaNiO3,206 SrRuO3,209,211,245 La1−xSrxMnO3,211 YBa2Cu3O7−x,196,246 or Sr2RuO4,212 in the latter case leading to atomically abrupt interfaces, as shown in Fig. 3(a).

FIG. 3.

(a) Cross-sectional high angle annular dark field STEM image of a BaTiO3 film grown on Sr2RuO4(001) electrode. (b) Cross-sectional HR-TEM of 1 μm thick BaTiO3/GdScO3(110) near the interface (indicated by the arrow). (a) Reprinted with permission from Takahashi and Lippmaa, ACS Appl. Mater. Interfaces 9, 21314 (2017). Copyright 2017 American Chemical Society. (b) Reprinted with permission from Schubert et al., Appl. Phys. Lett. 82, 3460 (2003). Copyright 2003 AIP Publishing LLC.

FIG. 3.

(a) Cross-sectional high angle annular dark field STEM image of a BaTiO3 film grown on Sr2RuO4(001) electrode. (b) Cross-sectional HR-TEM of 1 μm thick BaTiO3/GdScO3(110) near the interface (indicated by the arrow). (a) Reprinted with permission from Takahashi and Lippmaa, ACS Appl. Mater. Interfaces 9, 21314 (2017). Copyright 2017 American Chemical Society. (b) Reprinted with permission from Schubert et al., Appl. Phys. Lett. 82, 3460 (2003). Copyright 2003 AIP Publishing LLC.

Close modal

Other oxides used for the epitaxial growth of BaTiO3 include MgO (rock salt with a lattice constant a = 4.2117 Å)247 and LaAlO3 (rhombohedral below 800 K with a = 5.365 Å, c = 13.11 Å, and a pseudo-cubic lattice constant apc=3.79 Å),248 both of which have a large lattice mismatch to BaTiO3, respectively, of +5.0% and –5.5%. Likely as a consequence, the BaTiO3 film quality tends to be poorer in comparison to films grown on SrTiO3, even when buffer layers are incorporated that act to reduce the lattice misfit.196,216,217,219 Metal oxide substrates with smaller lattice misfits include MgAl2O4 (cubic spinel with a = 8.0831 Å),248 DyScO3 (orthorhombic, a = 5.43 Å, b = 5.71 Å, c = 7.89 Å; apc=3.94 Å),249 and GdScO3 (orthorhombic, a = 5.487 Å, b = 5.756 Å, c = 7.925 Å; apc=3.98 Å).250 For BaTiO3 films grown on MgAl2O4(001), where the lattice misfit is about 0.79%, it is found that relatively high growth temperatures (900–1050 °C) are required for full epitaxy, but with a defective interface as determined by TEM, possibly as a consequence of surface reconstruction of MgAl2O4, surface terraces exposing different surface terminations, and the large thermal misfit.221 GdScO3(110) and DyScO3(110) have small lattice misfits to BaTiO3, respectively, of –1.0% and –1.7% and provide excellent templates for the epitaxial growth of fully strained BaTiO3(001) films with sharp interfaces, as illustrated by the example shown in Fig. 3(b).218,220

The growth of complex oxides on semiconducting substrates, particularly Si, has been intensively investigated with the aim of integrating complex oxides in semiconductor platforms to realize multifunctional electronic and electro-optic devices167,251–260 and for using high-κ dielectric oxides as alternatives to SiOx for gate dielectrics in MOSFET transistors.170,261–263 The challenge is to achieve epitaxial growth and a sharp interface without the formation of an interfacial oxide amorphous layer that may be formed by exposure of the clean semiconducting surface to the oxygen species required for the metal oxide growth or due to reaction with the oxide layer that could degrade device performance. In the case of Si, sharp and abrupt interfaces, without the formation of an amorphous interfacial silicon oxide layer, were obtained for BaO, SrO, and SrTiO3.252,264–268 In these instances, the process relies on the passivation of the Si surface with a submonolayer-thick alkaline metal layer that protects it from oxidation while enabling the start of the oxide film growth. A recent review of the growth of BaTiO3 on Si, Ge, and GaAs substrates by MBE has been provided by Mazet et al.254 In most instances reported in the literature, the growth of epitaxial BaTiO3 on Si(001) proceeds through the use of a mediating SrTiO3 buffer layer, which has the added benefit of reducing the large misfit between Si (a = 5.431 Å) and BaTiO3 (of about –3.8%), which otherwise would favor the growth of a-oriented films. It is found that the lattice orientation (a or c) depends strongly on the SrTiO3 buffer layer thickness, surface quality, and oxygen partial pressure during growth, making the ferroelectric state difficult to control.254 Ge (a = 5.658 Å) has a smaller lattice misfit to BaTiO3, of about +0.27% at room temperature, on which high quality BaTiO3 films can be grown directly.265 The in-plane tensile strain leads to a preferred a-orientation; however, this can be circumvented by using buffer layers that induce a compressive in-plane strain, such as Ba1−xSrxTiO3230 or SrTiO3.269 One example of the BaTiO3 film quality that is possible to achieve on Ge(001) is shown in Fig. 4 for a 2.5 u.c. film grown by reactive molecular beam epitaxy by first depositing a 0.5 monolayers of BaO to passivate the Ge surface against oxidation, followed by the deposition of two monolayers each of BaO and TiO2 to form an amorphous BaTiO3 layer that is subsequently crystallized upon annealing.270 The TEM image in Fig. 4(a) shows full crystallization of the BaTiO3 layer in perfect registry to the Ge lattice; the analysis of crystal truncation rod (CTR) x-ray diffraction measurements271 together with the results of ab initio calculations could show that, for this thin BaTiO3 film, rumpling of the BaO interfacial layer occurs in response to Ge dimerization [Figs. 4(b) and 4(c)] and leads to distortions in the BaTiO3 unit cell away from the tetragonal distortion akin to the orthorhombic phase of BaTiO3; a 5.5 u.c. BaTiO3 film grown by additional deposition of three monolyers of BaO and TiO2 at elevated temperature show the tetragonal structure expected for bulk BaTiO3 at room temperature, however.270 Another study has demonstrated epitaxial growth of [001]-oriented BaTiO3 on Ge(001) through the deposition of a 2 nm amorphous BaTiO3 film that was subsequently crystallized upon annealing and that could be used as a seed for further BaTiO3 growth.232 Although GaAs (a = 5.653 25 Å) also has a small lattice misfit to BaTiO3 of about +0.19%, the lack of structural and chemical stability of the GaAs surface to the high temperatures required for BaTiO3 growth and interdiffusion precludes direct growth of BaTiO3 on GaAs, making the use of a buffer layer, such as MgO or SrTiO3, necessary for high quality single crystalline films.254,272 In this sense, the problem is reduced to one of growing the buffer oxide layer on GaAs.254,273

FIG. 4.

(a) High angle annular dark field TEM image of 2.5 u.c. BaTiO3/Ge(001). (b) Superimposed TEM image taken along the Ge[110] zone axis with a model structure of the system determined from surface x-ray diffraction measurements. The red arrows indicate rumpling of interfacial Ba along the [001] axis. (c) DFT-predicted ground state structure of the 2.5 u.c. BTO/Ge system. Reprinted with permission from Kumah et al., Phys. Rev. Lett. 116, 106101 (2016). Copyright 2016 American Physical Society.

FIG. 4.

(a) High angle annular dark field TEM image of 2.5 u.c. BaTiO3/Ge(001). (b) Superimposed TEM image taken along the Ge[110] zone axis with a model structure of the system determined from surface x-ray diffraction measurements. The red arrows indicate rumpling of interfacial Ba along the [001] axis. (c) DFT-predicted ground state structure of the 2.5 u.c. BTO/Ge system. Reprinted with permission from Kumah et al., Phys. Rev. Lett. 116, 106101 (2016). Copyright 2016 American Physical Society.

Close modal

2. PbTiO3, PbTiO3-based materials

Other important perovskite ferroelectrics are PbTiO3 and PbTiO3-based materials. Since the first reports of the crystal structure and ferroelectricity in the 1940s,274–277 PbTiO3 has attracted significant interest due to its excellent ferroelectric properties,60,278,279 including a high Curie temperature and a large ferroelectric polarization. At ambient conditions, PbTiO3 presents a tetragonal structure with a large c/a1.063.280 The Ti cations are off-centered from the oxygen octahedra, inducing a ferroelectric polarization in the system. The ferroelectric phase is stable up to 763 K, then undergoes a transition to the paraelectric phase accompanied by a structural transition from tetragonal to cubic symmetry.281 The spontaneous polarization of PbTiO3 crystals is around 75 μC/cm2, which is much larger than that of BaTiO3 (see Table I). In addition to the contribution from hybridization of Ti 3d and O 2p orbitals, the 6s lone electron pair of the Pb2+ cations and hybridization between the Pb 6s and O 2p orbitals provide an additional driving force for the structural distortion characteristic of PbTiO3.34,282–284 The ferroelectric distortion for PbTiO3 differs from that of BaTiO3 in that the oxygen octahedra move in the same direction, but with larger magnitude, than the Ti cation.

Given its outstanding ferroelectric characteristics, the epitaxial growth of PbTiO3 thin films has been extensively studied for decades. Various deposition methods, such as MBE, PLD, chemical vapor deposition (CVD), and sputtering, have been exploited to achieve high-quality epitaxial thin films.285–289 The majority of these studies employ [001]-oriented SrTiO3 substrates. On SrTiO3(001), PbTiO3 adatoms exhibit ideal 2D layer-by-layer growth290 or even step-flow growth291 to form atomically flat surfaces and interfaces. Coherent growth without undesired a-domain formation has been observed up to a film thickness of 340 nm,292 which is far thicker than the coherent critical thickness expected from the Matthews–Blakeslee model. The minimum thickness for the onset of ferroelectricity is reported to be extremely small (three-unit cells) on SrTiO3.127 These features make PbTiO3/SrTiO3(001) an ideal platform for ferroelectric interface devices. The successful synthesis of epitaxial PbTiO3 films has been also reported on other substrates, such as LaAlO3,293 MgO,294 Pt,295 and Si.286,287

Lead zirconate titanate, PbZrxTi1−xO3 (PZT), is the most widely used ferroelectric and piezoelectric material due to its large ferroelectric spontaneous polarization and piezoelectric coefficient. PZT can be seen as a solid solution of PbTiO3 and PbZrO3, an antiferromagnetic rhombohedrally distorted perovskite material. The phase diagram of PZT is characterized by an equilibrium tetragonal phase at a large PbTiO3 content (x0.5) and several rhombohedral structures for the PbZrO3 rich phases (x0.5).296,297 At x0.52, PZT is characterized by the coexistence of tetragonal, rhombohedral, and a more recently identified intermediate monoclinic phase,298–303 defining the so-called morphotropic phase boundary; such phase coexistence is responsible for the large piezoelectric response of PZT at this composition.165,297,304–306 In the context of interfacial ferroelectric devices, the tetragonal phase of PZT is the most interesting on account of the higher Curie temperature and a larger ferroelectric polarization along the tetragonal axis.174,235,301,307–310 The epitaxial growth of thin PZT films has been demonstrated on several substrates using various deposition techniques, as discussed in detail in Ref. 165, and we consider here more recent examples reported in the literature. Due to the volatile Pb element, PZT films are grown less often by MBE311 and instead by rf sputtering or PLD, typically on Nb-doped SrTiO3 substrates or on undoped SrTiO3 with a conducting oxide buffer layer, which provide a good template for the growth of tetragonal PZT (x < 0.5).312 For x = 0.2, the lattice mismatch to SrTiO3 is –1.2% and the films can grow coherently up to a thickness of about 15 nm.313,314 The ferroelectric properties of PZT films have been shown to depend sensitively on the condition of the substrate surface: high quality PbZr0.2Ti0.8O3 films grown by PLD on single terminated, atomically flat SrTiO3(001) substrates using a 20 nm thick SrRuO3 buffer layer were found to grow layer-by-layer, as deduced from atomic force microscopy, and possess a high saturation ferroelectric polarization value of 105 μC/cm2 (attributed to a slightly higher value of the c/a ratio, 1.06, as compared to the bulk value of 1.051).186 TEM data for a 90 nm thick film, reproduced in Figs. 5(a)–5(d), show abrupt, defect-free interfaces; in comparison, a PZT film grown on a less perfect SrTiO3 surface shows the presence of defects, as shown in Fig. 5(e), and is also characterized by a smaller ferroelectric polarization value of 65 μC/cm2.186 PZT films deposited on La0.8Sr0.2MnO3/SrTiO3(001) exhibit larger ferroelectric polarizations, of up to 85 μC/cm2, when the La0.8Sr0.2MnO3 film is grown by MBE on single terminated TiO2 SrTiO3 substrates315,316 as compared to PZT/LSMO films deposited by off-axis rf-sputtering on mixed terminated SrTiO3 substrates, where the ferroelectric polarization is typically of the order of 45 μC/cm2.317 

FIG. 5.

TEM results for PbZr0.2Ti0.8O3/SrRuO3/SrTiO3 heterostructures, showing (a) a cross-sectional image of a defect-free 90 nm thick PZT film and (b) corresponding electron diffraction pattern. (c) High-resolution TEM (HRTEM) image of an 8 nm thick PZT film and (d) Z-contrast STEM image showing the sharp and dislocation-free interface between the PZT and SrRuO3 layers. (e) Cross-sectional TEM image of a defective 120 nm thick PZT film. Reproduced with permission from Vrejoiu et al., Adv. Mater. 18, 1657 (2006). Copyright 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

FIG. 5.

TEM results for PbZr0.2Ti0.8O3/SrRuO3/SrTiO3 heterostructures, showing (a) a cross-sectional image of a defect-free 90 nm thick PZT film and (b) corresponding electron diffraction pattern. (c) High-resolution TEM (HRTEM) image of an 8 nm thick PZT film and (d) Z-contrast STEM image showing the sharp and dislocation-free interface between the PZT and SrRuO3 layers. (e) Cross-sectional TEM image of a defective 120 nm thick PZT film. Reproduced with permission from Vrejoiu et al., Adv. Mater. 18, 1657 (2006). Copyright 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

Close modal

The PbTiO3-based relaxor-ferroelectrics are also materials of high interest. Relaxor-PbTiO3 ferroelectrics, including Pb(Mg1/3Nb2/3)O3-PbTiO3 (PMN-PT) and Pb(Zn1/3Nb2/3)O3-PbTiO3 (PZN-PT), have been extensively studied due to their ultra-high piezoelectric response, with piezoelectric coefficient d33>1500 pC/N and electromechanical coupling k3390%,318,319 which exceed those of high-quality PZT films. Although the origin of the high piezoelectric response remains to be clarified, it makes relaxor-PbTiO3 ferroelectrics ideal materials for electroacoustic transducers and actuators.320 While the majority of the reports on relaxor-PbTiO3 ferroelectrics still focus on single crystals or thick films, thin films of relaxor-PbTiO3 ferroelectrics have also been extensively studied. Magnetron sputtering and PLD have been used to synthesize epitaxial relaxor-PbTiO3 thin films on various substrates, such as SrTiO3,321,322 LaAlO3,293,323,324 MgO,325 (La,Sr)(Al,Ta)O3 (LSAT),326 and Si.327,328 The ultra-high piezoelectric effect of the relaxor-PbTiO3 ferroelectric crystal also opens a new route to study strain engineering. Most studies on strain effects are based on the synthesis of films with different thicknesses (with different strain relaxation states) or on films grown on different substrates to induce a misfit strain. These methods, however, have the drawback that spurious effects related to slightly different growth conditions between different samples cannot be excluded. On the other hand, using relaxor-PbTiO3 ferroelectric crystals as a substrate enables electric control of strain in a single sample, an approach that has been used to study the effect of strain on ferroelectricity329,330 and ferromagnetism,331–334 for instance.

3. BiFeO3

BiFeO3 is a well-known multiferroic material characterized by high ferroelectric and antiferromagnetic (Néel) ordering temperatures of Tce=1123 K and TNm=643 K, respectively,169 making it one of the relatively few room-temperature multiferroic compounds. Under ambient conditions, BiFeO3 has a rhombohedral structure with ferroelectric easy-axis along 111pc and a high spontaneous polarization in excess of ∼100 μC/cm2.176,335–338 Unlike typical displacive ferroelectrics, such as BaTiO3, where off-centering of the Ti cations within the O cage is at the origin of the ferroelectric polarization, the main contribution to the ferroelectric polarization of BiFeO3 arises from lone pair Bi 6s orbitals.339,340 This allows BiFeO3 to have partially filled d shells on the Fe3+ site (and a spin moment) while ferroelectricity arises largely independently from the Bi3+ cations. Despite the independent origin of ferroelectric and magnetic order, the presence of Dzyaloshinskii–Moriya interactions results in a sizable magnetoelectric coupling between the two order parameters,341,342 which has been explored in novel artificial multiferroic heterostrutures for electric field control of magnetism.343–349 The magnetic structure of BiFeO3 is locally G-type antiferromagnetic, but the presence of a significant Dzyaloshinskii–Moriya interaction (DMI) associated with the polar displacement350,351 induces both a spin canting between adjacent spins (resulting in a small local net spin moment) and an incommensurate cycloidal winding of the spins within a cycloidal plane defined by the direction of the ferroelectric direction and the direction of propagation of the cycloid, with a periodicity of about 64 nm,352–354 leading to an effective zero magnetic moment. For a ferroelectric polarization state pointing along [111]pc, the possible propagation directions for the spin cycloid are [1¯10]pc,[011¯]pc, and [101¯]pc. An order of magnitude smaller second DMI interaction, resulting from the alternate rotation of the FeO6 octahedra along the [111]pc direction,351,355 causes the cycloid to be canted about 1o from the cycloidal plane, inducing a spin density wave.354,356–358 Under the effect of epitaxial strain, the spin cycloid can be suppressed or modulated in thin films.354,359–363

The first report of the growth of multiferroic epitaxial BiFeO3 thin films364 generated a strong interest in this system, well documented in the literature.365–370 Here, we mention only the case of BiFeO3 films grown on LaAlO3 crystals, which, due to the large compressive epitaxial strain ( 4.5%), have been found to stabilize in a pseudo-tetragonal phase with a very large c/a ratio of 1.25–1.3239,370,371 characterized by an enhanced spontaneous polarization of over 150 μC/cm2.372 The large polarization value of pseudo-tetragonal BiFeO3 provides great functionality for ferroelectric interface devices.370,373 In addition, with increasing film thickness, the tetragonal BiFeO3 phase partially relaxes to a pseudo-rhombohedral phase, resulting in an effective morphotropic phase boundary (MPB) where the two structures coexist (see Fig. 6).238,374 The MPB exhibits enhanced functionalities, such as enhanced piezoelectric properties and photovoltaic effect.375 The excellent ferroelectric, piezoelectric, and electrooptic370,376 properties of tetragonal BiFeO3 make it of high interest as a Pb-free ferroelectric replacement for PZT.

FIG. 6.

Left: Cross-sectional high angle annular dark field TEM image of the boundaries between pseudo-rhombohedral (R) and pseudo-tetragonal (T) regions indicated by dashed lines of a BiFeO3 film grown on an LaAlO3 substrate. Right: In-plane (black) and out-of-plane (red) lattice constants of BiFeO3 across the R-T boundary. Reprinted with permission from Zeches et al., Science 326, 977 (2009). Copyright 2009 AAAS.

FIG. 6.

Left: Cross-sectional high angle annular dark field TEM image of the boundaries between pseudo-rhombohedral (R) and pseudo-tetragonal (T) regions indicated by dashed lines of a BiFeO3 film grown on an LaAlO3 substrate. Right: In-plane (black) and out-of-plane (red) lattice constants of BiFeO3 across the R-T boundary. Reprinted with permission from Zeches et al., Science 326, 977 (2009). Copyright 2009 AAAS.

Close modal

4. HfO2

The recent discovery of ferroelectricity in pure and doped HfO2 ultrathin films173,377,378 is also of high interest for interfacial ferroelectric devices, given its compatibility with CMOS technology and the robust ferroelectric and dielectric properties at thicknesses in the range from 3 to 30 nm, including high electric coercive fields and high breakdown voltages (see Table I).379,380 Ferroelectricity in HfO2 is associated with a metastable polar orthorhombic Pca21 phase that can be stabilized by interfacial strain, growth orientation, small crystallite grain size, doping (such as with Si, Zr, Al, Y, Gd, Sr, and La) or in combination with the application of an external electric field.173,378,381–384 It is a displacive ferroelectric where the electric dipole moment arises from a coordinated displacement of four of the eight oxygen ions in the orthorhombic unit cell, defining the c direction of the orthorhombic structure and the direction of the ferroelectric polarization; the ferroelectric phase is chiral (upon a point rotation, the polarization reverses but the structure is not that of the mirror reflected structure), with both chiralities possessing the same energy and reversal of the polarization occurring via the coordinated displacement of the same four oxygen cations along the c direction.378,381

Most studies on ferroelectric HfO2 reported in the literature are for polycrystalline films grown by atomic layer deposition (ALD) or by chemical solution deposition, typically on TiN contact layers and requiring an annealing step for film crystallization after metallization,173,378,382,383,385,386 where the ferroelectric orthorhombic phase coexists with a non-polar tetragonal phase. However, a growing number of studies have reported the growth of epitaxial ferroelectric HfO2 films.387–398 One example is the epitaxial growth of 15–20 nm YO1.5-substituted HfO2 films by pulsed laser deposition on [100]-oriented yttrium oxide-stabilized zirconium oxide (YSZ) substrates,387 and on YSZ(110) single crystals using Sn-doped In2O3 (ITO) as a bottom electrode,389 where large ferroelectric polarizations and an estimated critical temperature of 723 K were observed. A cross-sectional annular bright field STEM image of a ca. 20 nm (YO1.5)0.07(HfO2)0.93/YSZ(100) is shown in Fig. 7 together with simulations of the patterns expected for the polar and non-polar phases, showing a good correspondence for the polar structure.387 Most recently, first principles calculations and x-ray diffraction measurements of epitaxial single crystals grown in the [111] orientation show stabilization of another predicted ferroelectric ground state Pnm21 structure.399 Hence, there is promise that better control over the crystalline growth of ferroelectric HfO2 thin films could be achieved; the harnessing of its excellent dielectric and ferroelectric properties would be of high interest for the future development of interfacial ferroelectric devices.379,400–402

FIG. 7.

(a) Cross-sectional annular bright field (ABF)-STEM image of a 20 nm (YO1.5)0.07(HfO2)0.93 epitaxial film grown on YSZ(100). Simulated ABF-STEM images for (b) polar Pca21, (c) non-polar Pbca, and (d) Pbcm structures. Adapted with permission from Shimizu et al., Appl. Phys. Lett. 107, 032910 (2015), Copyright 2015 AIP Publishing LLC.

FIG. 7.

(a) Cross-sectional annular bright field (ABF)-STEM image of a 20 nm (YO1.5)0.07(HfO2)0.93 epitaxial film grown on YSZ(100). Simulated ABF-STEM images for (b) polar Pca21, (c) non-polar Pbca, and (d) Pbcm structures. Adapted with permission from Shimizu et al., Appl. Phys. Lett. 107, 032910 (2015), Copyright 2015 AIP Publishing LLC.

Close modal

A related ferroelectric that has been recently discovered is monolayer ZrO2 grown on silicon.403,404 One of the stable crystal structures of ZrO2 is the fluorite structure, in which an individual atomic ZrO2 plane in the material is polar. In the bulk, this polarization is not switchable, but a combination of first principles calculations and experiment shows that a single ZrO2 plane grown atomically abruptly on silicon has two stable structures with different polarizations. The ferroelectric structure then consists of a continuous monolayer of ZrO2 that forms on silicon when grown using molecular beam epitaxy on an atomically clean Si(001) surface. When the polarization is switched by application of an electric field, the atomic configuration switches and shifts the silicon surface potential by 0.6 V.404 Capacitance–voltage measurements of the gate stack show a large hysteresis with a direction that follows the ferroelectric switching of the ZrO2 monolayer. This new class of materials consisting of a single atomic monolayer thick film represents the lower limit of device scaling.

One important aspect in studying thin ferroelectric films concerns the role of the supporting substrate, whose choice has to be considered judiciously. In addition to chemical stability,245 a first criterion is a relatively good lattice match in order to enable epitaxial growth, but equally important is its impact on the ferroelectric properties of the film due to the elastic and electric constraints that it imposes on the ferroelectric system. By acting as mechanical boundary condition, it can impose a strain on the ferroelectric film either due to the lattice and/or the thermal misfit associated with the growth process (respectively, lattice mismatch and difference in thermal expansion between film and substrate). Strain affects strongly the ferroelectric properties, including the critical temperature, piezoelectric constants, and ferroelectric polarization.12,14,405 For example, the ferroelectric critical temperature and polarization of strained BaTiO3 thin films can be enhanced by up to 500 K and 250%, respectively,218 while the quantum paraelectric SrTiO3 becomes ferroelectric under tensile or compressive strains of about 1%.406–408 The substrate also defines one electrical boundary condition, impacting the electrostatic energy of the system. In the case of an insulating substrate, the presence of a surface bound charge layer at the ferroelectric interface would result in a large electrostatic energy, which the system minimizes by breaking into ferroelectric domains, switching to an in-plane ferroelectric state, or in extreme cases, making the paraelectric state energetically more favorable. In the case of a conducting substrate, the free carriers in the latter will screen the electric field associated with the ferroelectric surface polarization and an electric depolarization field across the ferroelectric will be present if the potential difference to the other surface is different from zero (open circuit condition) or not, if there is no potential difference (short-circuit condition). In addition, the presence of charge carriers in the ferroelectric layer induced by either point defects177,409–412 or band bending at the interface413 can result in inhomogeneous internal electric field profiles across the interface, which can manifest in an imprint field that may favor one orientation of the ferroelectric polarization over the other.414 At a more subtler level, also the particular atomic termination of the substrate may condition the growth of the ferroelectric film and give rise to preferred orientations of the ferroelectric polarization.153,156,160,415,416

1. Elastic boundary conditions: Epitaxial strain

Typically, the growth of single crystalline films requires a single crystal substrate that serves as an atomic template for the oriented crystal growth (epitaxy). In the general case, the equilibrium lattice parameter, crystal symmetry, and chemical bonding of the deposited film material is different from that of the substrate (heteroepitaxy) and the lattice registry imposed by minimization of the bond energy at the interface sets an effective elastic boundary condition in the form of a misfit strain. In particular, the condition for a coherently strained epitaxial film applies insofar as the system can accommodate the associated elastic energy, which increases linearly with film thickness. Above a critical thickness, the build up in elastic energy can be favorably counteracted by mechanisms that act to release strain, including the formation of dislocations,417–422 surface roughness,422–425 and strain-induced atomic diffusion at elevated temperatures.426,427 Hence, epitaxial film growth typically proceeds first through coherent growth, whereby the epitaxial film grows fully strained and in registry with the underlying substrate up to a critical thickness above which strain relaxation processes set in, particularly by the introduction of misfit dislocations. By equating the energy gain through strain release to the strain energy associated with the dislocation core (which involves a disruption of the local crystal structure), expressions for the critical coherence thickness can been deduced to provide estimates of the thickness above which strain starts to be relaxed (which however do not consider kinetic barriers for misfit dislocation formation, such that larger critical coherence thicknesses than those predicted from theory are often observed experimentally).292,428 For ferroelectrics, epitaxial strain has a strong impact on the ferroelectric properties, given the intimate link between lattice structure and ferroelectricity in many ferroelectric systems.12,14,405 The effect can be twofold: for coherent strain, uniform changes in the ferroelectric properties may be expected, including changes in the ferroelectric polarization and critical temperature, while in the incoherent thickness regime, the local stresses associated with the cores of the dislocations and the associated strain gradients (such as through the flexoelectric effect)429–434 can lead to local modulations in the ferroelectric order parameter and to modified ferroelectric properties (Sec. II E).312 

The effect of coherent epitaxial strain on the properties of thin ferroelectric films has been extensively addressed in the literature.12,236,405 In the case of the perovskite ferroelectrics, the effect of biaxial strain is that of modifying the c/a ratio and rotating the oxygen octahedra.22,435 Its effect on the ferroelectric polarization can be estimated to linear order in strain in terms of the improper piezoelectric tensor cαi=Pα/ϵi, where ϵi are the strain components expressed in the Voigt notation and α stands for the cartesian coordinate. For an in-plane biaxial strain and considering only out of plane polarization, one obtains for the change in polarization236 

(1)

where n=ϵ1/ϵ3 is the Poisson ratio. The results for a number of systems (BaTiO3, BiFeO3, PbTiO3, and LiNbO3) are shown in Fig. 8 together with the results of ab initio calculations (symbols).236 It shows that the strongest variation occurs for BaTiO3 and PbTiO3, as a consequence of the large c33 values in combination with relatively low Poisson ratios, while for the other materials, the variation is more modest; one finds also no direct link between crystal structure and polarization susceptibility to in-plane biaxial strain. As expected for the perovskite ferroelectrics, biaxial in-plane compressive strain results in an increase in the c/a ratio and in the ferroelectric polarization, while the opposite is the case for tensile strain. Direct experimental confirmation of enhanced ferroelectric polarization of epitaxially compressively strained films has been reported in the literature, some examples of which are listed in Table II, showing critical ferroelectric temperatures and ferroelectric polarizations well in excess of the bulk value (Table I). For rhombohedral BiFeO3, the change in ferroelectric polarization has been observed to be largely independent of the average in-plane strain (i.e., below and above the critical thickness for coherent growth),428,436 in agreement with the theory predictions,236,437 and a similar behavior was reported for Pb(Zr0.2Ti0.8)O3,438 where the saturation ferroelectric polarization is found to be identical for both a strained 30 nm film (c/a=1.09) and a relaxed 100 nm film (c/a=1.05), an effect that has been ascribed to a suppressed sensitivity of the A-site cations to the epitaxial strain.438 

FIG. 8.

Variation of spontaneous ferroelectric polarization P for several ferroelectrics vs epitaxial strain ϵ1 estimated from expression 1 (lines) and from first principles calculations (symbols). Reproduced with permission from Ederer and Spaldin, Phys. Rev. Lett. 95, 257601 (2005). Copyright 2005 American Physical Society.

FIG. 8.

Variation of spontaneous ferroelectric polarization P for several ferroelectrics vs epitaxial strain ϵ1 estimated from expression 1 (lines) and from first principles calculations (symbols). Reproduced with permission from Ederer and Spaldin, Phys. Rev. Lett. 95, 257601 (2005). Copyright 2005 American Physical Society.

Close modal
TABLE II.

Ferroelectric characteristics of epitaxially strained thin ferroelectric films. t is the ferroelectric film thickness, ϵ1 the in-plane strain (when not available, the room temperature lattice misfit is given instead in square brackets); PLD: pulsed-laser deposition, RFS: rf sputtering; SRO: SrRuO3.

Systemt(nm)ϵ1(%)c / aPs(μC/cm2)Tc(K)GrowthtechniqueReferences
SRO/BaTiO3/SRO/GdScO3(110) 200 –1.0 1.026 50 693 PLD 218  
SRO/BaTiO3/SRO/DyScO3(110) 200 –1.3 1.031 70 953 PLD 218  
PbTiO3/SrTiO3(001) 130 [–0.06] 1.05 ⋯ >973 PLD 292  
Pt/PbTiO3/Nb:SrTiO3(100) 129 [–0.06] 1.238 236 >998 RFS 240  
SRO/PbZr0.2Pb0.8TiO3/SRO/SrTiO3(001) 90 [–0.7] 1.06 105 ⋯ PLD 186  
PbZr0.2Pb0.8TiO3/Nb-SrTiO3(001) 15 –1.2 1.07 ⋯ 953 RFS 313  
Pt/BiFeO3/LSMO/LaAlO3(100) 160 [–1.0] 1.24 130 ± 5 ⋯ PLD 439  
Systemt(nm)ϵ1(%)c / aPs(μC/cm2)Tc(K)GrowthtechniqueReferences
SRO/BaTiO3/SRO/GdScO3(110) 200 –1.0 1.026 50 693 PLD 218  
SRO/BaTiO3/SRO/DyScO3(110) 200 –1.3 1.031 70 953 PLD 218  
PbTiO3/SrTiO3(001) 130 [–0.06] 1.05 ⋯ >973 PLD 292  
Pt/PbTiO3/Nb:SrTiO3(100) 129 [–0.06] 1.238 236 >998 RFS 240  
SRO/PbZr0.2Pb0.8TiO3/SRO/SrTiO3(001) 90 [–0.7] 1.06 105 ⋯ PLD 186  
PbZr0.2Pb0.8TiO3/Nb-SrTiO3(001) 15 –1.2 1.07 ⋯ 953 RFS 313  
Pt/BiFeO3/LSMO/LaAlO3(100) 160 [–1.0] 1.24 130 ± 5 ⋯ PLD 439  

Epitaxial strain can also induce a phase transition to a distinct supertetragonal structure (i.e., with a value of c/a much larger than the bulk value and generally found to exceed 1.2) that may also have a large polarization. Two particular reports are striking, namely, for supertetragonal PbTiO3 and BiFeO3, where ferroelectric polarizations of 236 and 130 μC/cm2, respectively, have been reported,240,439 consistent with first principles calculations.234,440,441

In addition to modifying the ferroelectric polarization, epitaxial strain also affects the ferroelectric critical temperature by hindering the transition to the higher symmetry paraelectric phase, while the order of the phase transition may change from first to second order as a consequence of the reduced electrostrictive coupling in strained ferroelectric films;442–444 intuitively speaking, epitaxial strain prevents the system to lower the crystal symmetry to cubic across the phase transition, resulting in a continuous change in the order parameter. In addition, epitaxial strain can induce the formation of new crystalline phases that otherwise could not be stabilized in bulk.12,218,405,443,445 Phenomenologically, epitaxial strain adds an electrostrictive contribution to the polarization that modifies the onset of ferroelectricity in the film; for a [001]-oriented thin film, the change in critical temperature for an out-of-plane polarized film (tetragonal c phase) is estimated as443 

(2)

while for the onset of in-plane polarization along the 110 direction (orthorhombic aa phase),

(3)

where T0 is the bulk critical temperature, C is the Curie constant, ϵ0=8.854187817×1012 F/m is the electric permittivity of free space, Q are the electrostrictive coefficients, s are the elastic compliances, and um=(asa0)/a0 is the misfit strain, where as is the substrate lattice parameter and a0 the equilibrium lattice parameter of the free standing film. The above expressions predict that, when Q12<0 and Q11+Q12>0, as is the case for BaTiO3 and PbTiO3, the critical temperature increases (linearly) for both signs of the misfit strain.443 Examples of the effect of epitaxial strain in enhancing the critical temperature are given in Table II for several perovskite ferroelectrics, showing a significant increase in Tc with respect to their bulk values (Table I). The enhanced ferroelectric polarization and critical temperature induced by epitaxial strain are beneficial for ferroelectric interfacial devices, since it is conducive to larger field effects over a larger temperature range (as far as the ferroelectric component is concerned). In BiFeO3, the presence of both polar displacements and oxygen octahedra rotations yields an anomalous strain dependence of the Curie temperature. It actually decreases with increasing strain, which is captured by effective Hamiltonian calculations.446 

While the role of a uniform strain is largely beneficial and can be employed to tune the characteristics of ferroelectric thin films, strain relaxation in thin films is expected to have a predominantly negative impact on the ferroelectric properties.312 For example, calculations based on a thermodynamic analysis predict strong variations in the local ferroelectric polarization near the core of the dislocations that lead to strong depolarization fields that suppress ferroelectricity in a region that can extend over several nanometers from the interface.447–449 A telling example is the local reduction of the ferroelectric polarization of a PZT film by up to 48% caused by the strain field of a dislocation inside an adjacent SrTiO3 layer, determined from TEM.450 The presence of such a depolarization layer can lead to the formation of ferroelectric dead layers that absorb a large fraction of the applied electric field and that can act as pinning centers for the ferroelectric domain walls.451 The negative role of misfit dislocations has been identified in several systems, including PZT nanoislands452 and thin PZT films.453–456 

Another mechanism for strain relaxation in thin ferroelectric films is through domain formation, such as domains with either the a-axis or c-axis oriented perpendicular to the surface (a and c domains) in tetragonal ferroelectric systems.457–459 Also for this case, a critical thickness has been estimated above which it is energetically favorable for the system to develop a mixed domain configuration. A manifestation of this process can be found in thick c-oriented PZT films, where a significant a-domain population is often observed experimentally.145,456,460 The presence of a-domains is undesirable, as they lower the switchable out-of-plane ferroelectric polarization of the system and contribute to inhomogeneities that are detrimental to device scaling.458 

2. Electric boundary conditions and size effects

The substrate also sets an electrical boundary condition for the ferroelectric film with important consequences for the equilibrium ferroelectric configuration of the system. This is due to the role of the electrostatic energy, which, in the absence of free charge, is minimized for zero divergence of the ferroelectric polarization (·P=0) and no net bound charge at the boundaries (n·P=0, where n is the surface unit vector). For thin films, this is achieved for a uniform in-plane configuration of the ferroelectric polarization. However, since the direction of the latter is tied to the lattice, the ferroelectric polarization in epitaxial ferroelectric thin films is not free to orient along the direction that minimizes the electrostatic energy. In fact, for field effect devices, a uniform orientation of the ferroelectric polarization along the direction normal to the film plane is generally the most desirable configuration. The latter can be achieved by constraining the polar axis to be oriented along the out-of-plane direction, as in [001]-oriented tetragonal ferroelectric systems or in rhombohedral [001]-oriented BiFeO3 ([111]pc of the pseudo-cubic perovskite structure); also, in rhombohedral BiFeO3 grown on SrTiO3(001), the polarization makes an angle of about 36o from the surface, resulting in a large polarization component along the out-of-plane direction.355,364,365,461 When the ferroelectric polarization cannot be screened by free charges (such as provided by a conducting substrate and/or polar molecules adsorbed at the free surface),413,462,463 the electrostatic energy of the system can be minimized by the formation of oppositely poled ferroelectric domains127,129,464,465 or by reverting to the paraelectric state at the ultrathin limit.466–468 

The geometry of ferroelectric interfacial devices is typically that of a capacitor structure, where the ferroelectric layer is sandwiched between two conducting layers, for example, a gate contact and a channel layer in the case of ferroelectric field effect devices. These layers provide not only for device function but also act to screen the ferroelectric polarization at the respective interface and ensure the required stability for the uniformly polarized ferroelectric state. However, at metallic interfaces, screening occurs over a finite length scale of the order of the characteristic Thomas-Fermi screening length, δTF0.339(rs/a0)1/2 (where rs=(3/4πn)1/3 is the mean distance between carriers, n is the carrier density, and a0=0.529177 Å is the Bohr atomic radius).469 Over this screening length, the electric field associated with the ferroelectric polarization is non-zero and acts as a dielectric layer of thickness465d2δTF that adds to the electrostatic energy and contributes to a series capacitance that absorbs a fraction of the applied electric field. In fact, the depolarizing field associated with the screening length of the metallic contact can be sufficient to destabilize the ferroelectric order of ultrathin films, setting a limit to the minimum thickness required for the onset of ferroelectricity that depends on the electrode material, on the ionic polarizability, and on the ferroelectric band structure.34,466,467 For example, in PbTiO3 and BaTiO3, theoretical ab initio studies have predicted the onset of ferroelectricity at 1 and 6 unit cells, respectively.34,466,467 Experimentally, ferroelectricity has been confirmed in BaTiO3, PbTiO3, Pb(Zr0.2Ti0.8)O3, BiFeO3, HfO, and KNbO3 films down to a few unit cells in thickness (Table III).

TABLE III.

Experimental minimum thickness tc for the onset of ferroelectricity in selected ferroelectric systems.

MaterialElectrodeSubstrateTechniqueT (K)tc (nm)References
BaTiO3 SrRuO3 SrTiO3(001) P-E RT 199 and 470  
BaTiO3 SrRuO3 SrTiO3(001) PFM, STEM RT 1.4 471  
BaTiO3 LaAlO3/SrTiO3 SrTiO3(001) PFM RT 1.0 472  
PbTiO3 ⋯ SrTiO3(001) XRD 550 3 u.c. 127  
PbTiO3 SrRuO3 SrTiO3(001) XRD 700 1.2 462  
PZT Nb-SrTiO Nb-SrTiO3(001) PFM RT 4.0 473  
BiFeO3 Pt, LSMO SrTiO3(001) PFM RT 1.6 474  
HfO2 p+-Ge p+-Ge P-E RT 380  
KNbO3 ⋯ SrTiO3(001) STEM RT 475  
MaterialElectrodeSubstrateTechniqueT (K)tc (nm)References
BaTiO3 SrRuO3 SrTiO3(001) P-E RT 199 and 470  
BaTiO3 SrRuO3 SrTiO3(001) PFM, STEM RT 1.4 471  
BaTiO3 LaAlO3/SrTiO3 SrTiO3(001) PFM RT 1.0 472  
PbTiO3 ⋯ SrTiO3(001) XRD 550 3 u.c. 127  
PbTiO3 SrRuO3 SrTiO3(001) XRD 700 1.2 462  
PZT Nb-SrTiO Nb-SrTiO3(001) PFM RT 4.0 473  
BiFeO3 Pt, LSMO SrTiO3(001) PFM RT 1.6 474  
HfO2 p+-Ge p+-Ge P-E RT 380  
KNbO3 ⋯ SrTiO3(001) STEM RT 475  

For device applications, it is important to be able to apply electric fields, for example, to switch the direction of the ferroelectric polarization. Since most oxide ferroelectrics have relatively modest bandgaps (∼3 eV, see Table I) and a significant density of extrinsic charge carriers (originating typically from point defects as a consequence of the complex defect chemistry of 3d metal oxides, where the metal cations can exist in various oxidation states),412,476–478 the electrical character of the contact (Ohmic or diode-like) will depend on the difference between metal work function ϕ and the electron affinity of the ferroelectric χ (which defines the height of the Schottky barrier, Vbi=ϕχ), the charge carrier density in the ferroelectric (which defines the width of the depletion layer) and the presence of interface states (metal-induced gap states or interface defect states), which may pin the position of the Fermi level of the ferroelectric charge carriers.171,476–481 The electrostatic physics is illustrated in Fig. 9, showing the band alignment for a perfectly insulating ferroelectric in contact with identical metal electrodes compared to that of a defective ferroelectric layer. The situation is similar to that of a metal/semiconductor contact, where the role of the ferroelectric polarization is that of adding a sheet of bound charge at the interface to the electrostatic configuration.476,479,482 To account for the presence of a depletion layer in the metallic contact or a ferroelectric dead layer, a thin dielectric layer of thickness δ is added between the ferroelectric and the metal, which is otherwise treated as ideal. To first approximation, the ferroelectric polarization adds a term to the Schottky built-in potential, Vbi,476,479,482

(4)

where P is the ferroelectric polarization amplitude, δ is the width of the dielectric layer, ϵs is its static dielectric constant, and the ± sign corresponds to positive or negative directions of the ferroelectric polarization. For P50μC/cm2, δ0.2 Å, and ϵs2, the additional contribution to the built-in potential is of the order of 0.5 V, which is considerable. In fact, the experimental determination of the polarization dependence of the Schottky barrier height has been reported for several systems, including BaTiO3, BiFeO3, and Pb(Zr0.2Ti0.8)O3, showing indeed that such large variations in the Schottky barrier height are reached, see Table IV. In some instances, the change in the Schottky barrier may be sufficiently high to drive the contact from diode-like to Ohmic, leading to large changes in the conductivity of the system when switching the direction of the ferroelectric polarization.104,483–486 A similar interface structure will be present at the other side of the ferroelectric layer, possibly with a different conducting material. Hence, the requirement for being able to apply an electric field of sufficient amplitude for switching the ferroelectric polarization is that both interfaces consist of back-to-back Schottky diode junctions. The presence of charge carriers, either intrinsic or due to extrinsic doping arising from point defects, show that ferroelectrics for the most part are not perfect insulators, but instead exhibit charge transport under an applied electric field. In addition to thermally activated transport of charge carriers across the Schottky barrier, and in common to other insulators, tunneling in ultrathin films, electron hopping, or emission from deep traps are also found to contribute to the conductivity;411,487 for ferroelectrics, it has been recently discovered that ferroelectric domain walls tend to exhibit significantly different electronic properties as compared to the bulk material,488,489 including high electric conductivity.490–495 

FIG. 9.

(a) Schematic of the band alignment at the metal contact/ferroelectric interface for a perfectly insulating ferroelectric. (b) For an imperfectly insulating ferroelectric, the presence of mobile charges results in band bending to screen the potential associated with the difference in the Fermi levels of the metal and ferroelectric (including the interfacial bound charge).

FIG. 9.

(a) Schematic of the band alignment at the metal contact/ferroelectric interface for a perfectly insulating ferroelectric. (b) For an imperfectly insulating ferroelectric, the presence of mobile charges results in band bending to screen the potential associated with the difference in the Fermi levels of the metal and ferroelectric (including the interfacial bound charge).

Close modal
TABLE IV.

Variation of the Schottky barrier height for ferroelectric/metal contacts for the two directions of the ferroelectric polarization, pointing into (ϕ+) or away (ϕ) from the metal interface. Δϕ=ϕ+ϕ, and t is the thickness of the ferroelectric. XPS: x-ray photoemission spectroscopy; STS: scanning transmission spectroscopy.

Interfacet (nm)ϕ+ (eV)ϕ (eV)Δϕ (eV)TechniqueReference
BaTiO3/RuO2 bulk 2.85 ± 0.15 1.75 ± 0.15 1.1 ± 0.3 XPS 481  
BaTiO3/Pt bulk 2.90 ± 0.15 2.25 ± 0.15 0.65 ± 0.3 XPS 481  
BaTiO3/Pt 64 0.40 0.85 0.45 XPS 486  
BaTiO3/Nb-SrTiO3 64 0.45 –0.45 XPS 486  
BaTiO3/Cr 0.31 0.34 –0.03 XPS 497  
BaTiO3/Pt 0.42 1.84 –0.38 XPS 497  
BaTiO3/LSMO 2.8 1.35 1.05 0.3 XPS 134  
PZT/LSMO 0.98 0.25 0.7 XPS 498  
BiFeO3/Nb-SrTiO3 120 1.0 ± 0.1 0.2 ± 0.1 0.8 ± 0.2 STS 499  
Interfacet (nm)ϕ+ (eV)ϕ (eV)Δϕ (eV)TechniqueReference
BaTiO3/RuO2 bulk 2.85 ± 0.15 1.75 ± 0.15 1.1 ± 0.3 XPS 481  
BaTiO3/Pt bulk 2.90 ± 0.15 2.25 ± 0.15 0.65 ± 0.3 XPS 481  
BaTiO3/Pt 64 0.40 0.85 0.45 XPS 486  
BaTiO3/Nb-SrTiO3 64 0.45 –0.45 XPS 486  
BaTiO3/Cr 0.31 0.34 –0.03 XPS 497  
BaTiO3/Pt 0.42 1.84 –0.38 XPS 497  
BaTiO3/LSMO 2.8 1.35 1.05 0.3 XPS 134  
PZT/LSMO 0.98 0.25 0.7 XPS 498  
BiFeO3/Nb-SrTiO3 120 1.0 ± 0.1 0.2 ± 0.1 0.8 ± 0.2 STS 499  

Different types of mobile charge carriers may be involved in screening the ferroelectric polarization at the interface.41,414 For example, a detailed STEM-EELS study by Kim et al.496 of the BiFeO3/La0.67Sr0.33MnO3 interface in a region of the sample containing two ferroelectric domains with opposite polarization shows that the domain with the ferroelectric polarization pointing away from the interface presents an out of plane lattice expansion of ∼5% in the BiFeO3 interface region, together with a decrease in the Mn valency and a reduction in the O K edge intensity, which is interpreted as due to screening by oxygen vacancies; in contrast, the other polarization direction shows no such effects and screening is inferred to be purely electronic.496 

In addition to screening by mobile charges discussed above, other mechanisms can also contribute to screen the ferroelectric polarization, including ionic screening. For example, Gerra et al.500 suggested in their theoretical work that additional ionic compensation to the depolarization field can occur at a ferroelectric/electrode interface. By considering a ferroelectric BaTiO3 layer sandwiched between SrRuO3 electrodes as a model system, they show that the ferroelectric ionic displacement of the ferroelectric BaTiO3 layer may penetrate for several atomic layers into the SrRuO3 electrode. The induced ionic displacement inside the electrode acts to further stabilize ferroelectricity by pushing the center of mass of free electrons toward the interface. Subsequent theoretical and experimental work has shown that ionic compensation is active in various ferroelectric/metal interfaces and is crucial in stabilizing ferroelectricity down to the nanoscale.52,467,471,501 Conversely, the ferroelectric induced ionic displacement inside the electrode can result in changes in its electronic properties, leading to an additional mechanism for controlling the correlated state of the conducting channel (see Sec. IV).

Although we have focused on the impact on the ferroelectric properties of materials with decreasing film thickness down to the atomic scale, additional considerations come to the fore when also the lateral dimension of the ferroelectric system is reduced to the nanoscale, in particular for free standing nanoparticle systems, where new domain states such as flux closure states can occur to minimize the electrostatic energy; a critical size may be required to sustain a single domain state and, at smaller dimensions, to sustain a stable ferroelectric state against thermal excitations.502–505 

In addition to the constraints imposed by elastic and electrostatic constraints, atomic bonding effects arising from electron exchange at the interface with the substrate or top layer can have a strong effect on the overall properties of the system that are important to consider. The study of such phenomena often requires probing the local electronic structure at the interface, and in this context, the recent advances in high resolution scanning TEM have been invaluable.28,144 One aspect concerns the chemical bonding between the ferroelectric layer and the adjacent conducting channel, which may impose restrictions on the ferroelectric soft-mode across the heterointerface and affect significantly the ferroelectric properties of the system.52,506 In addition, the asymmetric environment at the interface can induce changes in the electronic state and a concomitant modification of the ferroelectric energy landscape. A striking example of this effect is a preference for a particular orientation of the ferroelectric polarization (imprint effect) that is found to depend on the termination of the interface layer, for example, in BiFeO3/LSMO,415 PZT/LSMO,156,416 and BaTiO3/LSMO.160 In extreme cases, the presence os such in-built bias potential can lead to back-switching of ferroelectric domains and loss of information.146,507 The role of interface termination and of the energy associated with the polar discontinuity508–511 on the ferroelectric state has been studied for several perovskite interfaces.415 For example, in ferroelectric BiFeO3 and metallic La0.7Sr0.3MnO3 heterostructures, the choice of the interface termination sequence, (BiO)+/(MnO2)0.7 or (FeO2)/(La0.7Sr0.3O)+0.7, induces a different charge valence mismatch, +0.3 or –0.3, at the interface, which results in the formation of an interface dipole that affects the polarization switching of the BiFeO3 layer.153,415 The interface dipole can be formed even when no charge valence mismatch is expected. In the BaO/RuO2 interface termination at BaTiO3/SrRuO3, the difference in ionic radii of the cations induce a pinned interface dipole that hinders ferroelectric switching of the system.512,513 The undesirable interface dipole effect can be avoided by changing the BaO/RuO2 termination to TiO2/SrO or by inserting a few atomic layers of SrTiO3.471,514 Such results emphasize the importance of the proper choice of the interface terminating sequence in designing electronic devices based on nanoscale ferroelectricity. Another effect pertains to atomic-scale structural distortions at the interface layers, not only in terms of uniform epitaxial strain, but also associated with modifications in the oxygen octahedral rotations characteristic of some ferroelectric systems, such as BiFeO3. For example, Kim et al.154 have reported the onset of a non-polar BiFeO3 interface layer at the La0.8Sr0.2MnO3 interface of BiFeO3/La0.8Sr0.2MnO3/SrTiO3(001) films, whereby a MnO2 termination of the La0.8Sr0.2MnO3 leads to a suppression of the octahedral rotation angle at the interface extending up to three atomic layers and whose effect propagates over a 3 nm region, while the LaSrO termination leads also to a reduced octahedral rotation compared to the bulk that extends to the whole of the film, as shown in Fig. 10. As a consequence of the presence of the non-polar BiFeO3 interface layer, the piezoelectric properties of the MnO2 terminated film are found to be strongly reduced as compared to the other termination.

FIG. 10.

Octahedral tilt and polarization mapping: bright field STEM images taken under optimal conditions for oxygen column contrast in the [110]pc-oriented BiFeO3 film structures grown on (a) MnO2-terminated LSMO and (b) (La,Sr)O-terminated LSMO. The resulting octahedral tilt maps (insets) show a checkerboard pattern of tilt angles. [(c) and (d)] Line profiles of octahedral tilts averaged over vertical rows of the respective tilt maps in (a) and (b). Solid lines in each graph represent tilt angles in the bulk BiFeO3 and LSMO. Reproduced with permission from Kim et al., Adv. Mater. 25, 2497 (2013). Copyright 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

FIG. 10.

Octahedral tilt and polarization mapping: bright field STEM images taken under optimal conditions for oxygen column contrast in the [110]pc-oriented BiFeO3 film structures grown on (a) MnO2-terminated LSMO and (b) (La,Sr)O-terminated LSMO. The resulting octahedral tilt maps (insets) show a checkerboard pattern of tilt angles. [(c) and (d)] Line profiles of octahedral tilts averaged over vertical rows of the respective tilt maps in (a) and (b). Solid lines in each graph represent tilt angles in the bulk BiFeO3 and LSMO. Reproduced with permission from Kim et al., Adv. Mater. 25, 2497 (2013). Copyright 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

Close modal

We focus in this review on metal oxide materials, for which the lattice parameter is largely determined by the oxygen sublattice and the oxygen ionic radius (in the range from 1.35 to 1.42 Å, increasing with increasing coordination number);515,516 hence, the growth of oxides on oxides is facilitated by both chemical compatibility and the relatively small lattice mismatches. In particular, we have highlighted ferroelectric systems where highly ordered single crystalline films can be obtained on a number of different substrate materials. However, we have also remarked that the oxide growth process is often associated with the creation of defects due to misfit strain relaxation processes (including the onset of dislocations, film roughening, and domain formation), local chemical reactivity at the interface, or associated with the oxygen defect chemistry in oxides, which is largely a direct consequence of the multiple valence states of the 3d transitions metals.476–478 By locally altering the strain field, by disturbing the local electronic structure, or by adding free carriers to the system, the presence of defects can affect strongly (mostly negatively) the properties of thin ferrolectric films, including by depressing the amplitude of the local ferroelectric polarization, increasing the coercive field, lowering the breakdown voltage and inducing current leakage, charge traps, and fatigue.8,451,517,518 Since it may not be possible to fully eliminate the presence of defects, it is important to understand their role on the properties of ferroelectric thin films and their impact on interfacial properties of ferroelectric devices, in order to determine which ones are more deleterious for a given application.

The mechanisms that lead to the presence of defects are directly linked to the growth process itself. The growth temperature of most ferroelectric films, in the range from 500 to 700 °C, tends to be well above the bulk Curie temperature and it is generally assumed that film growth starts from the paraelectric cubic phase (although, as shown in Table II, under misfit strain, ferroelectricity may be present, or start to develop already at such temperatures above a certain critical thickness).292,313,519–521 At this elevated temperature, the lattice misfit is generally different from that at room temperature,312,522 and above a critical thickness, elastic energy is relieved by the onset of misfit dislocations, whose nucleation is expected to be facilitated by the available thermal energy.417–421,458,522 Strain can also make the surface unstable against deformation, whereby elastic energy is reduced by surface roughening at a cost of surface energy.424,425,523,524 As the temperature of the film is reduced to ambient temperature after growth, the density of misfit dislocations may change to reduce the thermal misfit (difference in thermal expansion coefficients) for films above the critical thickness for coherent growth and, as the thermal energy for the movement of misfit dislocations becomes insufficient to reach equilibrium, a residual strain may be present in the system at ambient temperature. Finally, the transition to the ferroelectric state can result in the formation of ferroelectric domains, for example, a-domains in c-oriented tetragonal ferroelectric films, as another channel for misfit strain relaxation.451,457–459,525,526

Because of the relatively high Peierls energy in the perovskites (for SrTiO3 it is estimated to be of the order or 0.5 eV for 110 {110} edge dislocations),527 gliding of dislocations is increasingly suppressed as the temperature is reduced to room temperature, such that the dislocation network observed at low temperature can be assumed to have formed during film growth.451,522 A typical consequence of this kinetic barrier is the observation of strain opposite to that expected from the lattice mismatch at room temperature due to thermal misfit between film and substrate.205,528 The particular misfit dislocation pattern in thin films can be quite complex. For the cubic perovskites, screw and edge dislocations along 110 within {110} glide planes, with a Burgers vector (a/2)110, are energetically more favorable, but dislocations along other directions may also occur at elevated temperatures;529–531 in thin epitaxial films, only dislocations with Burgers vector parallel to the interface act to effectively reduce strain, although experimentally both edge and screw dislocations are observed.205,312,454 For example, in thick, [001]-oriented BaTiO3 films, one observes not only a criss-cross network of misfit dislocations, but also perpendicular screw dislocations and oblique screw dislocations in the {110} glide plane;205,532 these are thought to originate at point defects at the surface that then glide toward the interface, where they dissociate to form interfacial dislocation lines along 100 to relieve epitaxial strain, with the perpendicular component eventually propagating to the surface and eliminated there.205,533 Strategies for minimizing the formation of misfit dislocations include increased film deposition rates,243,534 annealing the film after growth,535 and employing well conditioned single terminated substrates,312 for example. Given the high internal stresses associated with the dislocation core and the direct coupling between internal stress and the ferroelectric order parameter, misfit dislocations result in strong local depolarization fields that modify the amplitude of the local polarization; in thin films with a high density of misfit dislocations, such depolarization fields may result in suppressed ferroelectric polarization (ferroelectric dead layers) and in pinning centers for domain wall motion;145,312,447–449,451,453–455 they can also act as preferential sites for atomic diffusion processes, such as oxygen transport (the impact of point defects is discussed below).536–538 This illustrates the point that different types of defects can interact in complex ways; for instance, strain fields arising from structural defects can lead to modifications to the local composition145,450,539 and promote atomic diffusion,426 while misfit dislocations can drive interdiffusion at the interface427 or lead to surface roughening and provide preferential nucleation sites for film growth.540 Strain gradients, originating from the stress fields issuing from dislocations or from compositional variations with thickness generated during film growth, can also result in depressed properties due to coupling with the ferroelectric polarization via the flexoelectric effect429–432 that acts as an additional local internal electric field and that provides another mechanism for the presence of imprint effects (seen as an horizontal shift in the P-E hysteresis loop) in defective films; however, such effect has been employed recently to bias the direction of the ferroelectric polarization and therefore to control the polarization state by suitable control of the growth process and film thickness.433,434,541,542

The formation of structural polydomains at the transition from the paramagnetic to ferroelectric phase as the film is cooled down to room temperature from the temperature of deposition is another mechanism for reducing misfit strain energy.451,457–459,525,526,543 Although the process here is driven by elastic energy considerations,544 the link between structure and direction of ferroelectric polarization implies that the system will break into a ferroelectric multidomain state as well, for example, a and c domains in the tetragonal ferroelectric perovskites, or a more complex domain structure for the rhombohedral perovskites, such as BiFeO3, where four variant domains can be present in thin films.365,461,545 For the tetragonal perovskites, the presence of a domains leads to a reduction of the switchable ferroelectric polarization,460 since the electric fields required to fully switch an a domain to a c domain, either by coherent rotation or by domain wall displacement (typically strongly pinned at defects such as dislocations),145,152,451 are much higher than those required to switch the polarization along the c direction. For example, in c-oriented PZT films, it is observed that pinning of a domains is associated with pairs of misfit dislocations with Burgers vectors [100] and [001], as illustrated in the TEM results shown in Figs. 11(a)–11(e). Model calculations shown in Fig. 11(f) of the in-plane strain field associated with the dislocation pair shown in Fig. 11(e) show the presence of a large region where the a domain is favored (red color), indicating that pinning of a domains results primarily from the strong strain field produced by the dislocation pair.145 

FIG. 11.

Origin of domain wall pinning. [(a) and (b)] STEM ADF images along the [010] zone axis of two a domains (marked by the solid lines) observed in the PZT(001) film, as schematically shown in (c). [(d) and (e)] STEM ADF images taken at higher magnifications on the areas marked in (c). (f) Calculated spatial variation of the elastic field ϵdis,xx, with the interface between the PZT and substrate set at y = 0; the region where ϵdis,xx>0.04 (red) is the area where the a domain should be stable. Reproduced with permission from Su et al., Appl. Phys. Lett. 99, 102902 (2011). Copyright 2011 AIP Publishing LLC.

FIG. 11.

Origin of domain wall pinning. [(a) and (b)] STEM ADF images along the [010] zone axis of two a domains (marked by the solid lines) observed in the PZT(001) film, as schematically shown in (c). [(d) and (e)] STEM ADF images taken at higher magnifications on the areas marked in (c). (f) Calculated spatial variation of the elastic field ϵdis,xx, with the interface between the PZT and substrate set at y = 0; the region where ϵdis,xx>0.04 (red) is the area where the a domain should be stable. Reproduced with permission from Su et al., Appl. Phys. Lett. 99, 102902 (2011). Copyright 2011 AIP Publishing LLC.

Close modal

One growth strategy for minimizing the nucleation of a polydomain state has been to employ stepped substrate surfaces obtained from cutting the wafer off the main index plane by up to a few degrees (miscut or vicinal surfaces). The role of the stepped surface is that of allowing an asymmetrical strain relaxation along the direction of the steps,546,547 setting a preferential crystallographic direction for strain relaxation and for the growth of systems with lower symmetry than that of the substrate. Such an approach has been used to reduce the number of domain variants of BiFeO3 films grown on SrTiO3(001) from four to two, resulting in an improvement in the ferroelectric switching characteristics of the BiFeO3 film,548,549 or to tune the crystal structure of Pb(Mg1/3Nb2/3)O3-PbTiO3 thin films.550,551 The steps tend also to act as nucleation sites for film growth and can lead to a step flow growth mode and to atomically smooth surfaces;552 however, the regions where the steps meet are sites susceptible to the formation of lattice defects, such as strain induced by residual lattice matching between the film and substrate,553 stacking faults,551 and vertical lattice mismatch as a result of step bunching;554 additionally, the atomic steps interact with dislocations and modify their configuration in the film.555 

Defect chemistry also plays an important role at the growth temperature. Intrinsic defects, such as vacant sites (vacancies), atoms occupying interstitial sites (interstitials), Frenkel defects (atoms that move to occupy an interstitial site leaving behind a vacancy), and Schottky defects (paired anion and cation vacancies) are thermodynamically favored and will be always present in the crystal, since they reduce the Gibbs free energy of the system by increasing the entropy.556 In the oxide perovskites, double ionized oxygen vacancies are one of the most important type of point defects;557 their equilibrium density at a given temperature T and oxygen partial pressure p[O2] can be obtained from the chemical reaction,409,557–560

(5)

in the Kröger–Vink notation,561 where OOx is the lattice oxygen, VO is the oxygen vacancy, and e is the electron, describing the release (capture) of a pair of electrons to the conduction band of SrTiO3 for each oxygen vacancy created (annihilated). The corresponding reaction equilibrium constant K1 is given by

(6)

where [OOx]N0 is approximately constant and close to the number density of oxygen sites (N0), p[O2] the oxygen partial pressure, Ha and ΔS are the enthalpy (activation energy) and entropy of reaction, respectively, and kB is the Boltzmann constant. For SrTiO3, Ha=5.69 eV/vacancy,558 for BaTiO3, Ha=5.96 eV/vacancy.562 One consequence of the presence of oxygen vacancies is that they contribute with charge carriers to the electrical conductivity, σi=μinie (where μi is the charge mobility, e the elementary electric charge, and ni the charge carrier density, where i stands for electrons or hole carriers), with the expression above predicting a variation of [e]=2[VO..]p[O2]1/6. Such variation is in fact observed, for example for SrTiO3 and BaTiO3, but only in the regime of high temperatures and strongly reducing conditions.409,557,558,562–564 More significant in determining the equilibrium concentration of oxygen vacancies is the effect of impurities (extrinsic point defects, typically present in ppm concentrations), which originate mostly from the source material and substitute for the intended atoms or occupy interstitial positions in the lattice (the latter being not favorable in closed packed lattices, such as the perovskite). Most impurities tend to be low valence cations, such as Na, Mg, and Al, which act as acceptors when they replace cations with higher valencies (and as donors in the opposite case).409,557,565,566 Since they tend to be the dominant factor determining the oxygen vacancy concentration, in the reaction equation (5), VO is approximately constant and the electron carrier density now varies as p[O2]1/4, which is observed experimentally over wide temperature and oxygen pressure ranges.409,557,558,562–564 Under oxidizing conditions, excess oxygen fills in part the impurity related vacancies, leading to the introduction of pairs of hole carriers to the valence band (and p-type conductivity) leading to a dependence in hole carrier density pp[O2]+1/4.567 For example, substitutional impurities such as Na, Mg, and Al act as acceptors and make PZT a p-doped ferroelectric (although the much larger mobility of minority electrons still makes the latter responsible for charge transport).177,409–411 In the intermediate regime, compensation of electron and hole charge carriers is reached and charge excitation becomes again intrinsic, i.e., determined by excitations over the bandgap.557,568–570 In particular, by deliberately doping the ferroelectric system with donors, the oxygen vacancy density can be reduced or fully compensated, resulting in softer ferroelectric properties.557,568,571 Systems with volatile elements, such as Pb in PbTiO3, or Bi in BiFeO3, may also have a significant density of cation vacancies at the growth temperature,409,572,573 while in systems where the electrons can be captured by multivalent cations, such as Fe in BiFeO3, the defect chemistry reaction is significantly different and more complex.571,574,575 In PZT, oxygen is released as PbO, such that formation of oxygen vacancies is accompanied by the formation of Pb vacancies.409,576

While oxygen vacancies do not always have a negative impact on the properties of oxide materials (controlling the density of oxygen vacancies may in fact provide another knob to controlling their electronic properties577 and they are desirable for oxygen diffusion and ionic conductivity in solid electrolytes, such as in stabilized zirconia, YSZ578–581) they generally have a deleterious impact on the dielectric properties of ferroelectric oxides.409 They can contribute to the electrical conductivity (in SrTiO3 they form an impurity state in the bandgap, just below the conduction band)573,582 and modify the oxidation state of neighboring cations. Furthermore, oxygen vacancies are mobile, with an activation energy of the order of 0.6–1 eV for the perovskite titanates,583–585 and can be displaced under the action of an applied electric field, or migrate at high temperature to the interface to screen the ferroelectric polarization,586 where the subsequent pileup at one electrode interface and depletion at the other can contribute to resistance degradation, imprint effects, and fatigue in ferroelectric capacitor structures.583,587 It has been inferred that the high density of free carriers in thin PZT ferroelectric layers may be a consequence of the formation of oxygen vacancies induced at the temperature of growth in order to screen the surface ferroelectric polarization (self-doping).412 Interfacial layers, by either acting as a sink, source, or by blocking the flow of oxygen, can affect the equilibrium oxygen vacancy density in the ferroelectric layer and impact its properties;409,572 in particular, by judiciously choosing the interfacial contact layers, improved retention and switching characteristics can be obtained. For example, the use of oxide electrodes in ferroelectric capacitors has been shown to reduce fatigue and imprint effects as compared with metal electrodes, which act to trap the oxygen vacancies within the ferroelectric.409 

Since an oxygen vacancy corresponds to a missing oxygen anion, for example, in the TiO6 octahedron of BaTiO3 or Pb(Zr0.2Ti0.8)O3, oxygen vacancies also give rise to local electric dipoles that point along, against, or perpendicular to the local ferroelectric polarization and which strongly perturb the local electric field, affecting the ferroelectric switching properties (by acting as pinning centers for domain walls, for example) and contributing to imprint effects.571,575,583 For similar reasons, they also cause local distortions in the atomic lattice, resulting in local strains. The impact of the latter will depend on the particular system: one may expect it to be strong in displacive ferroelectrics, such as BaTiO3, while in lone-pair ferroelectric rhombohedral BiFeO3 the effect of oxygen vacancies (and of strain) has been estimated to be minor.437 

Controlling the correlated state of matter is a fundamental topic in condensed matter physics, both because understanding the underlying mechanisms brings insights into the role of electron correlations to the emergence of novel physical phenomena and because those effects can potentially be useful for device applications. In fact, a major difficulty stems from the challenge of accounting fully for electron–electron many body interactions.588,589 Here we focus mostly on transition metal oxides, which are a prototypical class of systems exhibiting strong electron–electron correlations; this is manifested by the strong deviations in their behavior from that expected from a near-free electron system. Typical of these systems are complex phase diagrams, with several equilibrium states as a function of a given physical parameter, such as pressure or doping, as illustrated in the diagram shown in Fig. 12 representative of a wide class of 3d transition metal oxides. Starting from the insulating parent oxide, the addition of charge carriers does not initially modify strongly the electronic state, but at a critical carrier density, the carriers become delocalized, leading to a new electronic conducting ground state; with further increase in the charge carrier density, electron–electron correlations lead to the opening of a bandgap in the hybridized oxygen 2p-transition metal 3d band at half filling, resulting in a typical dome shape in the phase diagram.3 In fact, the general physical properties of strongly correlated systems may be described grosso modo in terms of the charge carrier density and electron bandwidth contributions of the Hubbard Hamiltonian.590 The electronic bandwidth is determined by the cationic crystal field and orbital overlap, while the charge carrier density can be modified through either chemical, ionic, or electrostatic doping. A classic example of the changes in the electronic properties of complex oxides with doping is that of the doped manganites AMnO3, where the average Mn valency can be modulated by varying the amount of di- or trivalent cations occupying the A sites of the perovskite structure via chemical substitution.591–595 One drawback of this approach to modulating the charge carrier density is that it is irreversible and involves local changes in the crystal structure induced by the different atomic size of the substituting element.596 Another approach to modulating the charge carrier is through ionic doping,597–599 based on ionic mass transport into or out of the material, for example oxygen or lithium, both of which have relatively high mobility in oxide materials559,560,600,601 and which can be driven by electric fields (Li+ or charged oxygen vacancies, VO2+). While in general, a change in the oxygen content in an oxide material is undesirable (an exception being in ionic conductors, such as stabilized zirconia, as mentioned earlier), it has been used to advantage for example in memristor devices, where different resistive states can be obtained by drastic modifications in the oxygen atom distribution under electric breakdown conditions,602 or to control the interfacial oxidation and the interfacial magnetic anisotropy of Co/GdOx by an applied electric field.598 Here, we focus on electrostatic doping, which relies on charge screening effects to modulate the charge carrier density at the interface between a dielectric and a conducting material, as in the classic field effect transistor. Unlike semiconductors, however, which are characterized by relatively low intrinsic carrier densities (10101018 cm--3),189 strongly correlated materials have very high charge carrier densities at the Fermi level, of the order of 1021 cm--3, requiring therefore very large induced surface charge polarizations at the dielectric interface which are not possible to achieve with conventional SiOx gates. Instead, large reversible modulation of the carrier density can be achieved in field effect device structures using either high dielectric constant materials or ferroelectrics, whose spontaneous surface bound charge is used to induce large modulations in the charge carrier density at the interface in ferroelectric field effect device structures;4,8,10,603,604 an alternative is to employ electrolytes, i.e., ionic conductors, where the electric field is screened at the electrode interfaces by mobile ions by forming a so-called double-layer of electric charge within a nanometer of the interface and no electric field within the electrolyte.605,606

FIG. 12.

Schematic representation of the operation of the ferroelectric field effect device (top) used to drive the channel system across two equilibrium states in the temperature-charge density phase diagram characteristic of 3d transition metal complex oxides (bottom).

FIG. 12.

Schematic representation of the operation of the ferroelectric field effect device (top) used to drive the channel system across two equilibrium states in the temperature-charge density phase diagram characteristic of 3d transition metal complex oxides (bottom).

Close modal

The original proposal for the ferroelectric field effect transistor58,59,607–609 aimed at combining the binary memory function of the ferroelectric polarization direction with the field effect control of the conductivity in a semiconductor channel in order to achieve a nonvolatile transistor switch, i.e., to extend the classical function of a field effect transistor. An extension of this concept consists of replacing the semiconducting channel layer by a strongly correlated material in order to control electrostatically its electronic state.4,5,85,610,611 The goal is to expand the functionality of such types of device structures by controlling the magnetic state, orbital or charge order, superconductivity, or the metal to insulator transition. A schematic of the ferroelectric field effect device structure and of the basic approach to controlling the electronic state of the channel layer is shown in Fig. 12: by choosing a chemical doping in the phase diagram of a strongly correlated system near the boundary between two different ground states, the system is driven across the boundary (horizontal arrow in the figure) through electrostatic carrier modulation, induced by charge carrier depletion or accumulation at the ferroelectric interface.3,317 In these device structures, the role of the ferroelectric layer is to provide the required change in the interfacial carrier density through its surface bound charge as the direction of the ferroelectric polarization is reversed by application of a voltage to the gate contact. The practical requisites have been discussed in Sec. II and include, for optimal effects, a low density of defects; an abrupt interface with a low charge trap density and minimal interdiffusion; large ferroelectric polarizations; and Schottky contacts to enable application of electric fields for ferroelectric switching. Alternatively, high dielectric materials such as SrTiO3 can be employed as a gate dielectric, which are capable of providing large surface polarizations at relatively modest applied electric fields, although, in contrast to ferroelectrics, the effect will be present only as long as the electric field is applied to the system. Based on such device structures, control of several physical properties of correlated materials has been achieved and will be discussed in Secs. III A–III E. These include control of the superconducting state in the cuprates and of the Mott insulator to metal transition in Mott insulators, and of spin, charge and orbital order in several metal oxide systems.

Superconductivity is one classic example of the effect of electron–electron and electron–phonon interactions inducing surprising physical phenomena in the solid state, such as zero resistivity and the expulsion of magnetic field lines known as the Meissner effect. In fact, while a good understanding of type I superconductors has been achieved through the Bardeen–Cooper–Schrieffer (BCS) model,612,613 a complete physical understanding of type II, high-Tc, or unconventional superconductivity, occurring at temperatures where the BCS theory would preclude superconductivity, has still not been reached yet.614–616 Given the intimate link between superconductivity and charge carrier density, it is not surprising that control of the superconducting state through electrostatic doping was attempted early, with the first reports dating back to the 1960s demonstrating small modulations in the superconducting critical temperature in metals, of the order of 0.1–1 mK.617–619 In addition to offering one knob for controlling the physics governing the electron pairing mechanism, and a perspective for a better understanding of superconducting phenomena by avoiding compositional structural modifications that come with chemical doping,3,620 electrostatic control of superconductivity is also of potential interest for device applications, such as superconducting FETs, where in principle, switching the system between the superconducting and normal states could result in very high on/off resistance ratios.621–625 From the BCS expression for the critical temperature,612,613,626,627

(7)

where kB is the Boltzmann constant, ω and V0 are the parameters of the BCS effective interaction (with ω/kB of the order of the Debye temperature), and N0=g(EF)/2 is the (free electron) density of electronic levels for a single spin population in the normal metal, it follows that

(8)

predicting a proportional increase (decrease) in the critical temperature with electron accumulation (depletion), consistent with earlier observations.617–619 

In superconductors, in addition to the screening length over which electric fields can extend into the system, two other important characteristic length scales are the London penetration depth, λL, over which magnetic fields and electric currents can be present in the superconductor, and the coherence length ξ, over which changes in the order parameter occur. The response to magnetic fields depends strongly on whether the London penetration depth is smaller or larger than the coherence length, ξ, separating superconductors into type I and II, respectively.627 For electric fields, the relevant comparison is between the Thomas-Fermi screening length and the coherence length perpendicular to the surface, ξ,624,628 with stronger field effects expected when δTFξ. Since in metals the electron density is high, the Thomas-Fermi screening length and the change in the carrier density within ξ are small, such that the changes in Tc are also correspondingly small. One way to optimize the field effect is to reduce the film thickness down to a minimum where superconductivity is still observed (of the order of ξ) and ideally to a value comparable to the Thomas–Fermi screening length. Using such an approach, a gradual transition from the normal to the superconducting state could be observed in a 10 Å thick bismuth film grown on SrTiO3(001) upon accumulation of charge carriers through electrostatic doping using the substrate as the gate dielectric.629 In this example it was argued that, in addition to the change in the carrier density, the electronic screening (described by V0) was modulated by the increase in electron density. Reducing the lateral dimensions of the system can also result in larger and new field effects, as illustrated by recent field effect experiments carried out in Al and Ti planar wire structures fabricated on Si substrates, showing an ambipolar modulation of the critical current that is not expected from a conventional modulation in the electron density.625,630 The results mentioned thus far are for polycrystalline metallic films of type I superconductors, indicating that field effect control of the superconducting state is robust, even if both the size of the effect and the critical temperatures tend to be small in these systems.

To achieve larger perturbations in the superconducting state using field effects, systems with larger screening lengths and shorter perpendicular coherence lengths are the most suited.631,632 In this context, the discovery of high temperature superconductivity in the cuprates633 made much larger field effect modulations of superconductivity possible.632 The cuprates consist of electron- or hole-doped layered perovskites with phase diagrams similar to those schematically shown in Fig. 12, exhibiting an antiferromagnetic phase at low doping that competes with the superconducting phase at higher doping, with a critical temperature that peaks at the so-called optimal doping. Between the antiferromagnetic and superconducting region, a pseudo-gap phase emerges for hole doped systems.590,634,635 These systems offer several advantages for field effect modulation, including much lower carrier densities as compared to metallic superconductors, a much smaller coherence length along the crystalline c direction (superconductivity being essentially two-dimensional, occurring in the CuO2 planes of the layered structure) and, importantly, much higher critical temperatures of up to 135 K at ambient pressure.636 In addition, the presence of boundaries in the doping phase diagram separating superconducting from normal and antiferromagnetic phases suggests that direct changes in the ground state of the system may be achieved by crossing one such boundary using electrostatic doping.632 Indeed, much larger modulations in superconductivity induced by field effects were reported for films of cuprate superconductors, first by using a gate dielectric to modulate the charge carrier density,620,637–645 and later by using ferroelectric gates,507,646–654 electrolytes,655,656 and also ionic doping.657 In Table V, we list several examples of systems where the variation in Tc has been obtained through electrostatic doping, together with a figure of merit β that gives a measure of the relative change in Tc for a relative change in carrier density via Eq. (8).

TABLE V.

Electrostatic control of the superconducting critical temperature (ordered chronologically and by type of gate, dielectric or ferroelectric).

Systemt (u.c.)Tc(K)ΔTc(K)n(×1021 cm−3)ΔQ(μC/cm2)βReferences
Kapton/YBa2Cu3O7 30 89.9 +1.1×104 4.1 0.13 0.007 637  
STO/YBa2Cu3O7−x 2.3a –2 –40 640  
STO/YBa2Cu3O7−x 4.7a 20 –2 –40 640  
BTO/Sm0.7Ca0.3Ba2Cu3O7−x 50 40 +1 0.6c 8.0 1.5 641  
Bi2Sr2CaCu2O8+x/BSTO 21 24 –2 1.8 –25 642  
YBa2Cu3O7−x/BSTO 10 68 –1 2.3 –25 0.2 642  
STO/PrBa2Cu3O7−x/STO 2.3a 17.4 +2.8 (2.3) 644  
Al2O3/Nd1.2Ba1.8Cu3Oy 9.3a 18 IST 0.56 ⋯ 620 and 645  
YBa2Cu3O7−x/BTO 100 83 0.5 14 646 and 647  
YBa2Cu3O7−x/LNOb 100 60 40 140 30 646 and 647  
PZT/GdBa2Cu3O7−x 50 20 0.4 649  
PZT/STO/GdBa2Cu3O7−x 8.8 35.2 0.65 46 0.2 650  
PZT/NdBa2Cu3O7−x 10 60 1.25 20 0.2 651  
PZT/NdBa2Cu3O7−x 78.4 1.5 20 0.1 652  
BFO/YBa2Cu3O7−x 3.5a 40 30 0.9 130 0.3 653  
BFO/YBa2Cu3O7−x 4.7a 40 15 0.88 130 0.2 654  
BFO/YBa2Cu3O7−x 3.5a (55) 28 (0.9) 140 0.2 660  
BFO/YBa2Cu3O7−x 4.7a 38 120 0.1 507  
PZT/Sr(Ti0.98Nb0.02)O3 26 0.266 0.06 0.05 12 0.4 661 and 662  
Systemt (u.c.)Tc(K)ΔTc(K)n(×1021 cm−3)ΔQ(μC/cm2)βReferences
Kapton/YBa2Cu3O7 30 89.9 +1.1×104 4.1 0.13 0.007 637  
STO/YBa2Cu3O7−x 2.3a –2 –40 640  
STO/YBa2Cu3O7−x 4.7a 20 –2 –40 640  
BTO/Sm0.7Ca0.3Ba2Cu3O7−x 50 40 +1 0.6c 8.0 1.5 641  
Bi2Sr2CaCu2O8+x/BSTO 21 24 –2 1.8 –25 642  
YBa2Cu3O7−x/BSTO 10 68 –1 2.3 –25 0.2 642  
STO/PrBa2Cu3O7−x/STO 2.3a 17.4 +2.8 (2.3) 644  
Al2O3/Nd1.2Ba1.8Cu3Oy 9.3a 18 IST 0.56 ⋯ 620 and 645  
YBa2Cu3O7−x/BTO 100 83 0.5 14 646 and 647  
YBa2Cu3O7−x/LNOb 100 60 40 140 30 646 and 647  
PZT/GdBa2Cu3O7−x 50 20 0.4 649  
PZT/STO/GdBa2Cu3O7−x 8.8 35.2 0.65 46 0.2 650  
PZT/NdBa2Cu3O7−x 10 60 1.25 20 0.2 651  
PZT/NdBa2Cu3O7−x 78.4 1.5 20 0.1 652  
BFO/YBa2Cu3O7−x 3.5a 40 30 0.9 130 0.3 653  
BFO/YBa2Cu3O7−x 4.7a 40 15 0.88 130 0.2 654  
BFO/YBa2Cu3O7−x 3.5a (55) 28 (0.9) 140 0.2 660  
BFO/YBa2Cu3O7−x 4.7a 38 120 0.1 507  
PZT/Sr(Ti0.98Nb0.02)O3 26 0.266 0.06 0.05 12 0.4 661 and 662  
a

Thickness (t) converted to unit cells (u.c.) using the lattice parameter of YBaCu3Ox, c = 11.68 Å.658Tc is the critical temperature, ΔTc the change in Tc for the change in surface charge density ΔQ, and n is the average carrier density (values in parenthesis are estimated from the expected number of hole carriers per CuO2 unit). IST: insulator to superconductor transition.

b

Deduced from films grown on the two opposite faces of the LiNbO3 crystal.

c

Estimated from Ref. 659.

By using dielectric gates to electrostatically control the charge density, large modulations in the critical temperature and/or critical current of high Tc superconductors have been observed. Although the effect is non-persistent, this approach has provided interesting insights into the physics of high-temperature superconductors. For example, using a 7 μm thick Kapton dielectric gate, Fiory et al.637 studied the variation of the effective mass and hole carrier density of YBa2Cu3O7 films, to show that the hole carrier density is constant in temperature (except in a region near Tc) and to obtain an estimate of the mass of the Cooper pair, m*=(5.5±0.5)me at T < 70 K.663,664 The first result showed that the hole carrier density cannot be estimated from the Hall coefficient directly, which instead exhibits a marked variation with temperature in the cuprates.507,651,657,660,665 Further, it suggested that the critical temperature of the surface layer (probed by modulating the excitation current) is the same as that of the bulk of the film (30 nm thick). Larger effects could be reached using high dielectric constant oxides, including SrTiO3 and BaxSr1−xTiO3, with changes in Tc of a few K. Since the oxygen stoichiometry determines hole doping and Tc in YBa2Cu3O7−x,666 an important aspect is whether mobile oxygen vacancies are at the origin of the electric field modulation of superconductivity, a question that was answered in the negative in the study by Frey et al.642 by showing no significant differences in the field effect response between YBa2Cu3O7−x and a well oxidized superconductor Bi2Sr2CaCu2O8+x, ruling out a major contribution from electric field induced oxygen diffusion.643 Dielectric gates have also been employed to control the onset of superconductivity induced by field effects at the LaAlO3/SrTiO3 interface.667–673 In this system, the carrier density could be modulated to map a large region of the temperature-doping space parameter, which exhibits a dome-like shape superconducting state within the normal state and indicating the presence of a quantum critical point.667 Rather than a result of a simple change in the sheet carrier density, the main impact has been ascribed to changes in the electron charge mobility as a result of compression or expansion of the electron gas due to band bending.674,675 An alternative approach to probing the phase space electrostatically is by ionic doping, such as in Bi2Sr2CaCu2O8+x,657 achieved by growing the film on a lime-glass substrate and by changing the carrier density through the diffusion of Na+ cations induced by the applied electric field. Here also, the effect of lattice distortions that arise with chemical doping is avoided.

Larger and nonvolatile modulations in Tc are found in ferroelectric gated cuprate superconductors, of up to 30 K in YBa2Cu3O7−x,653 and where, in some instances, switching between the normal metallic and superconducting state could be achieved.620,645 One strategy to minimize the film thickness has been to employ a suitable buffer layer, such as PrBa2Cu3O7,649,654 a semiconductor that acts to decouple the superconductor from the SrTiO3 substrate and adds a first layer of CuO chains and metallicity that are thought to promote superconductivity in the superconducting layer.676,677 An example is provided in Fig. 13 for BiFeO3/3 u.c. YBa2Cu3O7−x, showing that the critical temperature can be shifted by 30 K between the two polarization states of BiFeO3.653 This example also shows that the change in Tc decreases with increasing film thickness, as expected for a field effect and for the screening length of YBa2Cu3O7−x, of about 1 nm.624 One other interesting prospect for ferroelectric field effect devices ensues from the possibility of locally controlling the ferroelectric domain pattern678,679 to create nanoscale superconducting devices, for example, to create weak links in device structures.660 Examples of other systems where ferroelectric control of the superconducting state has been reported include Nb-doped SrTiO3, where on/off switching of superconductivity has been demonstrated using PZT as a ferroelectric gate.661,662

FIG. 13.

(a) Temperature variation of the resistance of a BiFeO3/3 u.c. YBa2Cu3O7−x (BFO/YBCO) heterostructure (normalized to the resistance at 150 K and measured at J = 1.7 kA/cm2) for two neighboring areas in which the BiFeO3 polarization points along opposite directions, as shown in the figure insets. (b) Superconducting critical temperature Tc vs charge carrier density n for samples with different YBa2Cu3O7−x thicknesses (in u.c.), for the depletion (empty symbols) and accumulation (solid symbols) states. Inset: variation of Tc with n. Reproduced with permission from Crassous et al., Phys. Rev. Lett. 107, 247002 (2011). Copyright 2011 American Physical Society.

FIG. 13.

(a) Temperature variation of the resistance of a BiFeO3/3 u.c. YBa2Cu3O7−x (BFO/YBCO) heterostructure (normalized to the resistance at 150 K and measured at J = 1.7 kA/cm2) for two neighboring areas in which the BiFeO3 polarization points along opposite directions, as shown in the figure insets. (b) Superconducting critical temperature Tc vs charge carrier density n for samples with different YBa2Cu3O7−x thicknesses (in u.c.), for the depletion (empty symbols) and accumulation (solid symbols) states. Inset: variation of Tc with n. Reproduced with permission from Crassous et al., Phys. Rev. Lett. 107, 247002 (2011). Copyright 2011 American Physical Society.

Close modal

The metal–insulator transition refers to a change in the conducting state of the system, between metallic and insulating regimes, driven by a control parameter, such as temperature, pressure, charge doping, magnetic field, or disorder.590,680–685 Several physical mechanisms can be responsible for such a transition, including changes in band overlap (Wilson metal–insulator transition), disorder (Anderson metal–insulator transition), charge instabilities (Peierls metal–insulator transition) and electron correlations (Mott metal–insulator transition).686 Although we focus here on electric field control of the correlated state, we discuss briefly for completeness the other metal–insulator transition mechanisms. The physics underlying metal-insulator transitions is complex and touches on several fundamental concepts in solid state physics, including the role of electron correlations, disorder, electronic band structure, and underlies important magnetic and transport phenomena such as superconductivity and colossal magnetoresistence.590 

One point of interest in using the metal–insulator transition for logic switches is that of achieving smaller subthreshold swings, S=dVg/d(logId) (gate voltage Vg needed to reduce the drain current Id by one decade), to achieve lower switching energies than is currently possible with MOSFETs, where S(1+Cd/Ci)(kBT/q)ln10 is essentially limited by the thermal excitation of the charge carriers to (kBT/q)ln1060 mV/dec at room temperature (Cd and Ci are the capacitance of the depletion and oxide layers, respectively, T the temperature and q the elementary charge value).189,687 Hence, one motivation for Mott FET transistors is that by replacing the physical mechanism driving the charge carrier density limited by Boltzmann statistics to a faster and more abrupt process based on a phase transition, one may achieve smaller subthreshold swings.687 In this context, it is worth mentioning a new device concept based on “negative capacitance” that seeks to reduce the m=1+Cd/Ci factor by designing the gate structure such that an effective negative value for Cd is obtained.688–690 

1. Wilson metal–insulator transition

One mechanism for the onset of the metal–insulator transition is due to band overlap (Wilson metal–insulator transition), such as in divalent metals where metallicity is due to overlap between the first and second zones and where an insulating state is expect to set in when the atomic distance is increased to the point where reduced orbital overlap leads to electron localization and conduction becomes possible only through hopping.683,691 Transitions of this kind occur in several materials under high pressure, where the reduced lattice parameter (larger orbital overlap) leads to band overlap and to the onset of an insulator to metal transition, such as in iodine,692–695 selenium,693,694 and oxygen.696–699 A more familiar example of (electrostatic) control of band overlap in band insulators is in semiconductors, such as in the MOSFET transistor, where band bending under a gate voltage changes the channel layer from a back-to-back diode junction to a conducting inversion layer, where electron confinement gives rise to a two-dimensional electron gas.189,700 Another example of control of the conduction state in band insulators is that of the LaAlO3/SrTiO3(001) system, where at a critical LaAlO3 thickness of about 4 u.c. the system transitions from insulator to metallic by the onset of a 2D electron gas state that sets in at the interface701 (which can be further driven into a superconducting state by gating).667–673 (Various mechanisms have been advanced for explaining the onset of the 2D electron gas state at the interface, including band bending, polar catastrophe, interfacial strain effects, and oxygen vacancies.)702 Field effect devices employing SrTiO3687,703–707 and KTaO3708,709 as the channel layer have also been shown to display large changes in resistivity as a function of applied gate voltage, with on/off current ratios of up to 105. In one instance, a transition from metallic to insulator state in n-doped SrTiO3 with applied gate voltage has been observed, where a CaHfO3 layer is used as the gate insulator (epitaxial at the interface but amorphous above 4 unit cells thickness).706 This effect is observed as a change in the temperature variation of the channel conductance, from a resistive to a conductive regime, at a bias voltage of 1.3 V; the conduction mechanism is explained in terms of thermal carrier excitation with an activation energy barrier that is found to collapse (i.e., drop to zero) for bias voltages above 1.3 V, leading to a metallic state (the conductivity in the insulating state being attributed to the presence of in-gap states). Interestingly, from a scaling analysis of the resistivity data, it was concluded that the system behaves as three dimensional.

Another example that could be described as a Wilson insulator to metal transition is found to occur at the PZT/LaNiO3 interface, where changes in the conductivity of LaNiO3, a metallic oxide, are found to be much larger than expected based on the charge carrier modulation in the thin LaNiO3 channel layer (3–4 unit cells).710 The results of ab initio calculations, shown in Fig. 14, suggest that the PbO interface layer of PZT in contact with the TiO2 layer of LaNiO3 is subject to an upward band shift when going from the depletion to the accumulation state that makes it cross the Fermi level, leading to the appearance of a conductive PbO state and to a new conducting channel at the interface.

FIG. 14.

Band structure of PbTiO3/LaNiO3 strained to the theoretical in-plane lattice constant of LaAlO3 for (a) accumulation and (b) depletion, where zero of energy is the Fermi level in each case and a (1 × 1) interfacial unit cell is employed. Red-colored bands correspond to dominant LaNiO3 characters, while blue-colored bands indicate strong contributions from the interfacial PbO layer. The top of the PbO-dominated bands shift from about –1.2 eV in depletion to approximately 0.5 eV in accumulation. (c) The layer-resolved LDOS at the interface for accumulation and (d) depletion. The PbTiO3 is insulating in depletion and in accumulation away from the interface, but exhibits a finite LDOS at the Fermi level at the interface in the latter case. Reproduced with permission from Marshall et al., Phys. Rev. Appl. 2, 051001 (2014). Copyright 2014 American Physical Society.

FIG. 14.

Band structure of PbTiO3/LaNiO3 strained to the theoretical in-plane lattice constant of LaAlO3 for (a) accumulation and (b) depletion, where zero of energy is the Fermi level in each case and a (1 × 1) interfacial unit cell is employed. Red-colored bands correspond to dominant LaNiO3 characters, while blue-colored bands indicate strong contributions from the interfacial PbO layer. The top of the PbO-dominated bands shift from about –1.2 eV in depletion to approximately 0.5 eV in accumulation. (c) The layer-resolved LDOS at the interface for accumulation and (d) depletion. The PbTiO3 is insulating in depletion and in accumulation away from the interface, but exhibits a finite LDOS at the Fermi level at the interface in the latter case. Reproduced with permission from Marshall et al., Phys. Rev. Appl. 2, 051001 (2014). Copyright 2014 American Physical Society.

Close modal

2. Anderson metal-insulator transition

The presence of disorder, such as atomic disorder, can also lead to the onset of an insulating state (Anderson metal-insulator transition), whereby deviations from the periodic lattice potential result in localization of the electronic states near the Fermi energy and to charge transport through hopping between states separated by varying energy barriers.680,711 In weakly disordered Anderson insulators, a metallic state can be induced by increasing the charge density such that the Fermi energy rises to above the localized states to reach the metallic states (at the so-called mobility edge)681,711,712 by electrostatic doping, for example, a process that has been extensively investigated in disordered semiconductors.711 Disorder-induced localization effects can also play an important role in the transport properties of correlated oxides, for example due to the short-range disorder associated with the atomic substitution in mixed valent materials, such as La and Sr site disorder in La0.5Sr0.5MnO3, which has been associated with the onset of a disorder-induced insulating state in thin films at low temperatures.713 

3. Peierls metal–insulator transition

Charge instabilities induced by electron–phonon interactions in low dimensional systems can lead to changes in the atomic and band structure and also result in a transition from a conducting to an insulating state (Peierls metal-insulator transition).687,714,715 Peierls insulators are typically strongly anisotropic materials characterized by exotic charge transport properties, in that charge can be transported by the charge density wave above an applied voltage threshold.715 To our knowledge, electrostatic control of the Peierls metal-insulator transition has not been reported in the literature.

4. Mott metal–insulator transition

In contrast to early expectations of electron band theory, many systems containing an odd number of charge carriers per atom are insulators, such as NiO, a behavior that was explained in terms of the role played by electron correlations (electron–electron Coulomb interaction), which favors electron localization (singly occupied electron states) and a gap opening in an otherwise half-filled electron band, resulting an insulating state (Mott insulators).590,683 In fact, until recently it was thought that the failure of spin dependent local density functional theory (DFT) to correctly predict the opening of band gaps in many complex materials was associated with the shortcomings in accounting for the electron correlation energy in a mean field approximation; however, recent results indicate that the opening of band gaps in DFT ensues once the early constraints imposed to the simulated system (such as cell size, atomic relaxation, breaks in symmetry) are relaxed, indicating that the exchange functionals used are sufficient to describe the band structure of complex oxides.716–719 Nevertheless, a common approach to describe the effect of electron correlations in the opening of such a bandgap in systems with half-filled bands has been to use the Hubbard model Hamiltonian, where two competing terms, one favoring electron hopping and another favoring localization, induce a gap opening in the density of states with increasing strength of the intra-site energy U. In this picture, two limit cases are distinguished for metal oxides:590,720 in one case, typically for light 3d transition metals, the oxygen derived p band lies well below the 3d band and the effect of electron correlations is that of splitting the original 3d band into a fully occupied lower Hubbard band and an empty upper Hubbard band, separated by an energy gap of the order of the intra-site energy U (Mott-Hubbard insulators). A second case occurs in oxides with strong p-d hybridization, such as in the late transition metal oxides, where the wavefunction overlap between cations and the oxygen p-orbitals is strong and U is larger than the energy separation between the d and p bands; the effect of electron correlations in such case is to split the d-band into a filled lower Hubbard band below the (filled) p band and an empty upper Hubbard band separated from the p-derived band by an energy gap Δ ( charge transfer insulators). This division underlines the fact that in Mott-Hubbard insulators charge excitation occurs through direct hopping between cation sites while in the charge transfer insulators, charge excitation occurs primarily through the oxygen anions.590 In Table VI, we list the electronic band parameters of several oxide materials following this picture, mostly derived from x-ray spectroscopy measurements, running from the early to the later transition 3d metals, showing a gradual decrease in Δ and an increase in U as one goes down the table (increasing 3d atomic number).

TABLE VI.

Properties of selected perovskite strongly correlated oxide materials. BI: band insulator; CT: charge transfer insulator; MH: Mott–Hubbard insulator; NM: non-magnetic; AFM: antiferromagnetic order; NO: not ordered. Tc is the critical temperature for the onset of the magnetic order, Δ is the p-d charge transfer energy, U is the d-d Coulomb repulsion energy, and Eg is the direct (optical) bandgap. Example of typical A-site dopants are given in the last column.

SystemTypeSpin orderTc (K)Δ (eV)U (eV)Eg (eV)Dopants
SrTiO3 BI NM ⋯ 4.0l 4.5l 3.75a La, Nb 
LaTiO3 MH AFM 140b 6.0m 4b 0.2b Sr, Y 
LaVO3 MH AFM 143d 6c 4.5c 1.1e Ca, Sr 
LaCrO3 CT AFM 286g 5.2m 5.2m 3.4e Ca, Sr 
LaMnO3 CT AFM 140 4c 5.5c 1.1e Ca, Sr, Ba, Pb 
LaFeO3 CT AFM 740h 2.5n 7.5n 2.1e Ca, Sr, Ba, Bi 
LaCoO3 CT NO ⋯ 2.0p 5.5p 0.3e Sr 
LaNiO3 NO  1.0q 7.0q 0e 
SrCuO2j CT AFM 442i   3.3f Ca, La, Nd 
La2CuO4 CT AFM 325s 0.3–0.4t 5–6.5t 2.0k Sr, Ba 
YBa2Cu3O7−δ CT AFM 400 0.5u 5–6.5u 1.5v 
SystemTypeSpin orderTc (K)Δ (eV)U (eV)Eg (eV)Dopants
SrTiO3 BI NM ⋯ 4.0l 4.5l 3.75a La, Nb 
LaTiO3 MH AFM 140b 6.0m 4b 0.2b Sr, Y 
LaVO3 MH AFM 143d 6c 4.5c 1.1e Ca, Sr 
LaCrO3 CT AFM 286g 5.2m 5.2m 3.4e Ca, Sr 
LaMnO3 CT AFM 140 4c 5.5c 1.1e Ca, Sr, Ba, Pb 
LaFeO3 CT AFM 740h 2.5n 7.5n 2.1e Ca, Sr, Ba, Bi 
LaCoO3 CT NO ⋯ 2.0p 5.5p 0.3e Sr 
LaNiO3 NO  1.0q 7.0q 0e 
SrCuO2j CT AFM 442i   3.3f Ca, La, Nd 
La2CuO4 CT AFM 325s 0.3–0.4t 5–6.5t 2.0k Sr, Ba 
YBa2Cu3O7−δ CT AFM 400 0.5u 5–6.5u 1.5v 
a

Reference 721.

b

Reference 722.

c

Reference 723.

d

Reference 724.

e

Reference 725.

f

Reference 726.

g

Reference 727.

h

Reference 728.

i

Reference 729.

j

Infinite-layer, high pressure, tetragonal phase of SrCuO2.

k

Reference 730.

l

Reference 731.

m

Reference 732.

n

Reference 733.

p

Reference 734.

q

Reference 735.

s

Reference 736.

t

Referenecs 737 and 738.

u

References 737 and 738.

v

For semiconducting δ = 1, Refs. 739 and 740.

At constant temperature, two approaches can be envisaged for inducing an insulator to metal transition in the Mott insulators, namely, through bandwidth modulation or through band filling.590 Bandwidth modulation can be achieved by changing the atomic spacing to control the orbital overlap through mechanical or chemical pressure (by substituting isovalent cations within the compound) and thereby affect the strength of electron correlations; in the doped manganites, it can be controlled by applying a magnetic field, where electron hoping is then modified through the double exchange mechanism,741,742 a process underlying the so-called “colossal” magnetoresistance (CMR) effect.743–747Band filling consists in directly modulating the charge carrier density, for example, through chemical, electrostatic, or ionic doping. In the case of chemical doping, it is found that the insulating state is stable against relatively large doping contents, an effect that is ascribed to carrier localization induced by static random potential and/or electron-lattice interaction;590 doping levels of the order of 0.1%–10% are typically required for a transition into a metallic state to occur. This implies that large charge carrier modulations are required to bring a Mott insulator to the metallic state. Doping by chemical substitution has the disadvantage of being irreversible and of introducing local distortions in the atomic lattice that can further perturb the electronic state. An alternative approach is to use field effects to modulate the carrier density, which is simultaneously reversible and does not introduce local distortions to the lattice. In this approach to controlling the Mott metal insulator transition, the channel layer consists of a Mott insulator chosen with a composition on the phase diagram close to the insulator to metal transition. For maximum effect, the channel layer is made as thin as possible, ideally of the order of the Thomas–Fermi screening length of the metallic phase. One important consideration here is that, due to the epitaxial strain or other interface effects, the phase diagram of the Mott insulator may deviate strongly from that of the bulk. One example is La1−xSrxMnO3, where epitaxial strain strongly affects the transport and magnetic properties in ultrathin films.748,749

One motivation for the electrostatic control of the Mott transition has been as an alternative to CMOS field effect transistors, where the main advantages include high on/off resistance ratios (expected at the metal–insulator transition), good scaling characteristics, and the inherent fast dynamics (since the process is purely electronically driven).750–753 Prototype Mott transition field effect transistors based on cuprates and using SrTiO3 as the gate dielectric were reported by Newns et al.,750,754 where control of the conductivity of the channel layer with gate voltage and a quasi-linear variation of the drain current with the drain–source voltage were demonstrated. Unlike in superconducting FETs, the aim here is to control the metal–insulator transition and the measurements were carried out at room temperature. In another example, electrostatic doping of Sm2CuO4 was achieved by using the band bending characteristics when in contact with a Nb-doped SrTiO3 interface,755 an approach that has been extended to other materials systems.756 In addition, control of the resistive state of Mott insulators using ionic liquids as the gate dielectric has been reported in the manganites, including control of the electronic state742 and of the conduction percolation path.757 

A large number of studies has been carried out on ferroelectric Mott devices based on several channel systems, including cuprates,758,759 manganites,603,760–766 nickelates,710,767,768 and cobaltates.764 Several aspects need consideration. A first aspect is that the resistivity of the doped Mott insulator near the insulator to metallic transition is usually far lower than that of the undoped system. For example, for La1−xSrxMnO3, the 0 K resistivity for x = 0.175, in the metallic regime, is 1.0×103 Ω cm, while for x = 0.15, in the insulating state, it is 2.0×103 Ω cm, which in turn is orders of magnitude lower than that for x = 0.769 As a consequence, screening lengths will be small, requiring small channel layer thicknesses for effective field effect modulation. However, control of the electronic properties of strongly correlated oxides at small thicknesses is challenging, since their properties tend to deviate strongly from the bulk for ultrathin films due to their sensitivity to strain and defects.749,767 Another aspect relates to changes in other properties accompanying the metal-insulator transition, such as changes in the magnetic and orbital state and phase segregation as a result of competing ground states.593,770–772 For example, early ferroelectric field effect transistors employed thick manganite layers, 30–50 nm thick,603,773 and the large changes in resistivity found, up to 300%, were interpreted within a percolative phase separation picture.773 In another study, on 8.2 nm La1−xCaxMnO3 films using SrTiO3 as a gate dielectric, an ambipolar effect with the applied electric field was observed,774 an effect that was attributed to the development of a pseudogap in the density of states, providing another illustration of the complex behavior of this class of complex oxide materials.

Since large changes in the charge carrier density are required to control Mott insulators, it is advantageous to employ ferroelectric gates with large polarizations. Promising in this regard is BiFeO3 in its supertetragonal phase, characterized by very large spontaneous ferroelectric polarizations in excess of 100 μC/cm2 at room temperature,239,372,439 and which has been employed for the control of the electric conductivity of CaMnO3 and Ce-doped CaMnO3.765,766 For CaMnO3, large changes in conductivity are obtained, of 400% at room temperature and a tenfold change at 200 K, although the system remains in the insulating state for the film thicknesses investigated, in the range from 6 to 40 unit cells.765 With Ce-doping of 0.04, a metal to insulator transition is observed at around 100 K.765 However, ferroelectric field modulation of the resistivity was found to be much less efficient in this system, which was attributed to incomplete ferroelectric polarization reversal at the interface, as determined from local measurements of the ferroelectric polarization with HR-STEM.766 

Control of the conductivity of the rare-earth perovskite nickelates has also generated interest. These materials have electrical properties that depend sensitively on the lattice structure and which are strongly affected by the overlap of Ni 3d and O 2p orbitals, as determined by the Ni–O bond length and Ni-O-Ni bond angle.775,776 Due to the strong electron-lattice coupling, one route to controlling the electrical properties of the nickelates is by modulating the Ni–O–Ni bond angle using strain777 and the surface termination.778,779 However, such approaches are limited to static control of the Ni–O–Ni bond angle. In this context, a recent report on the ferroelectric PbTiO3/LaNiO3 interface suggested a novel way to achieve active control of the electrical properties of the nickelates via interface structure modulation.780 In this case, switching the ferroelectric polarization of the PbTiO3 layer modifies the Ni–O bond length and Ni–O–Ni bond angles at the interfacial LaNiO3 layer through ferroelectrically induced structural distortions; the latter contribute to the conductivity change of the LaNiO3 channel by a modification of the mobility, unlike in the canonical field effect.710 The result shows that the ferroelectric interface structure modulation can provide a novel way to controlling functionality in oxide electronics.

Strategies for optimizing the changes in the conduction state of Mott insulators have been proposed, such as by combining different materials in bilayer stacks in order to better engineer the interface charge modulation. For example, it is observed that changes in the conductivity of Sm0.5Nd0.5NiO3, which exhibits a metal to insulator transition as a thin film grown on LaAlO3,767,768 can be enhanced from a 10% to a 300% effect when a thin LSMO layer is added at the substrate interface.768 In that study, x-ray spectroscopy measurements supported by ab initio calculations showed that between the interfacial Mn and Ni layers a charge transfer from Mn to Ni of about 0.1 electron per 2D unit cell takes place, which has the effect of making the system more susceptible to the ferroelectric charge modulation.

In the examples discussed above, large changes in the resistivity of the doped Mott insulators are observed, much higher than expected based on the simple modulation of the charge carrier density, a signature of the role of electron correlations; however, the state of the system tends to remain either insulating or conducting, an indication of the difficulty in controlling the Mott insulator to metallic transition. However, for optimized device structures, a genuine control of the conduction state can be achieved. An example is shown in Fig. 15 for PZT/La0.8Sr0.2MnO3, where the La0.8Sr0.2MnO3 film thickness is set at the onset of conductivity (3.8 nm or 10 u.c.) and a change between insulating and conducting states is observed as the system is switched between depletion and accumulation states, respectively.761 

FIG. 15.

Resistivity vs temperature curves for PZT/3.8 nm La0.8Sr0.2MnO3 heterostructure for the two polarization states of PZT (“off”/“on” is the depletion/accumulation state). The inset shows a lin-log plot of the data. Reproduced with permission from Hong et al., Appl. Phys. Lett. 86, 142501 (2005). Copyright 2005 AIP Publishing LLC.

FIG. 15.

Resistivity vs temperature curves for PZT/3.8 nm La0.8Sr0.2MnO3 heterostructure for the two polarization states of PZT (“off”/“on” is the depletion/accumulation state). The inset shows a lin-log plot of the data. Reproduced with permission from Hong et al., Appl. Phys. Lett. 86, 142501 (2005). Copyright 2005 AIP Publishing LLC.

Close modal

Orbital order refers to the cooperative onset of a preferential occupation of different orbital states which otherwise are degenerate in energy.683,781–786 When those states carry no orbital moment, such as the eg states, the degeneracy can be lifted by the Jahn-Teller effect (which posits that orbital degeneracy makes the structure of the molecular complex unstable against distortions),787 while in the case of states carrying orbital moment, such as the t2g states, the degeneracy can alternatively be lifted by the spin–orbit coupling.781 Long-range orbital order arises from a preferential spatial overlap of neighboring orbitals and is dictated by the sign of the exchange integrals, similar to the situation with the spin Hamiltonian, such that an orbital Hamiltonian can be written in terms of orbital pseudo-spins.781–783 Hence, orbital order corresponds to an additional fundamental degree of freedom in the solid state, with its fundamental collective excitation modes,788 and whose control constitutes another knob to controlling the electronic state of the system, including the magnetic state to which it is often intimately linked. One example is the orbital arrangement in LaMnO3, where the single electron at the eg level occupies alternately d3z2r2 or dx2y2 orbitals to form a complex ordered orbital structure that is also linked to the antiferromagnetic spin structure.788 Also important in the context of the orbital polarization is the role of strain, and of epitaxial strain in particular, since distortions in the crystal structure can profoundly modify the local crystal field and the position of the atomic energy levels, and therefore, the orbital arrangement.

Although orbital order has been identified in many systems, its control remains challenging. A first hurdle relates to probing orbital order, which typically requires techniques that are sensitive to local orbital arrangement, such as resonant x-ray scattering788,789 or x-ray absorption spectroscopy using the x-ray linear dichroic effect,790 although it can also be probed indirectly, such as with polarized visible light.791 For example, control of the orbital domains in La1/2Sr3/2MnO4 using electric current has been reported, where Joule heating results in melting of charge-orbital ordered domains and the subsequent reduction of the current favoring one particular domain over the other.791 Electric-field control of the orbital order has been investigated theoretically using first principles calculations in perovskite vanadate792 and PbTiO3/LaTiO3 superlattices.793 In these instances, the polar distortion (induced in the case of the vanadates by a trilinear coupling with antiferrodistortive motions and/or Jahn–Teller distortions) couples with the non-polar ordered Jahn–Teller distortion modes (leading to orbital order), which in turn couple to the spin configuration via the superexchange interaction.792,793 In principle, such coupling makes electric field control of orbital and spin ordering in those systems possible.

A related concept, particularly relevant for thin films and interfaces, is that of orbital polarization, which refers to a preferential orbital occupation in the system.27 For example, epitaxial strain in thin films can induce a modification in the energy position of the eg orbital states in the manganites and a change in the relative orbital occupancy, i.e., in orbital polarization.749,794–796 Control of orbital polarization at the interface between ferroelectrics and complex oxides has been reported for LSMO/BiFeO3,797,798 LSMO/BaTiO3,160,799 and LSMO/PZT.800 For the case of LSMO/BiFeO3,797,798 a modification of the orbital occupancy of the 3d states at the interface was inferred from x-ray linear dichroic effect measurements due to hybridization between the Mn and Fe orbitals and leading to the presence of a large interfacial magnetic moment in BiFeO3, exchange-coupled to the La0.8Sr0.2MnO3 magnetization; however, in these studies, the effect of switching the ferroelectric polarization of BiFeO3 was not addressed, nor the extent of the modified orbital polarization. These latter aspects were studied by Chen et al.799 and Preziosi et al.800 for La0.8Sr0.2MnO3/BaTiO3 and La0.825Sr0.175MnO3/PZT, respectively. In the first study, the change in orbital polarization was investigated by first principles calculations, predicting the occurrence of a large change in orbital polarization when going from the accumulation state (which favors occupation of the dx2y2 states) to the depletion state (which favors d3z2r2 states) over a region extending to 3 unit cells from the ferroelectric interface [Fig. 16(a)]. The calculations also predict that the orbital polarization depends only modestly on the ferroelectric polarization amplitude and that the main effect results from a displacement in the atomic position of the Mn cations induced by the ferroelectric polar displacements [Figs. 16(b) and 16(c)], a result that was confirmed experimentally by TEM measurements of the same system [Fig. 16(d)].799 Hence, the change in orbital occupation is determined by distortions in the Mn octahedron that, in turn, modulate the energy position of the two eg bonding states. These results were corroborated by x-ray linear dichroism measurements reported for La0.825Sr0.175MnO3/PZT, where a change in orbital polarization of 1.5% averaged over the whole LSMO film is observed (but expected to be significantly higher in the interface region).800 Multiplet calculations suggest as well an inversion on the order of the dx2y2 and d3z2r2 states when switching the ferroelectric polarization (where the latter are favored for the depletion state), pointing also here to a modulation in the Mn octahedron structure (see Fig. 17). Changes in the orbital moment of Mn in La0.67Sr0.33MnO3/BaTiO3 structures where the direction of the ferroelectric polarization was pre-set by controlling the substrate termination, were also observed and quantified by using XMCD sum rules801–805 at the Mn L2,3-edge, providing another demonstration of ferroelectric control of the orbital population.160 

FIG. 16.

[(a)–(c)] Ab initio results showing, respectively, the variation in orbital polarization, c/a ratio, and Mn cation displacement δ as a function of atomic position in an La0.8S0.2MnO3 (LSMO) film from the BaTiO3 interface for the two directions of the BaTiO3 ferroelectric polarization. (d) Scanning transmission electron microscopy results for the La0.8Sr0.2MnO3/BaTiO3 interface (accumulation state), demonstrating the presence of a cation displacement at the interface region. Adapted with permission from Chen et al., Nano Lett. 14, 4965 (2014). Copyright 2014 American Chemical Society.

FIG. 16.

[(a)–(c)] Ab initio results showing, respectively, the variation in orbital polarization, c/a ratio, and Mn cation displacement δ as a function of atomic position in an La0.8S0.2MnO3 (LSMO) film from the BaTiO3 interface for the two directions of the BaTiO3 ferroelectric polarization. (d) Scanning transmission electron microscopy results for the La0.8Sr0.2MnO3/BaTiO3 interface (accumulation state), demonstrating the presence of a cation displacement at the interface region. Adapted with permission from Chen et al., Nano Lett. 14, 4965 (2014). Copyright 2014 American Chemical Society.

Close modal
FIG. 17.

Experimental and multiplet calculations of the x-ray linear dichroism measured in La0.825Sr0.175MnO3/PZT for the two directions of the PZT ferroelectric polarization. The bottom panels show the schematic ionic model with the changes in the relative energy position of the 3d eg states in the two cases. Reproduced with permission from Preziosi et al., Phys. Rev. Lett. 115, 157401 (2015). Copyright 2015 American Physical Society.

FIG. 17.

Experimental and multiplet calculations of the x-ray linear dichroism measured in La0.825Sr0.175MnO3/PZT for the two directions of the PZT ferroelectric polarization. The bottom panels show the schematic ionic model with the changes in the relative energy position of the 3d eg states in the two cases. Reproduced with permission from Preziosi et al., Phys. Rev. Lett. 115, 157401 (2015). Copyright 2015 American Physical Society.

Close modal

In systems with mixed cation valency, a reduction in Coulombic energy can be achieved at half or 1/8 doping through ordered charge localization (charge-ordered state), manifested by an electrically insulating state below a critical temperature. The paradigmatic system where charge ordering was first described is magnetite (Verwey transition),806–810 while more recently, charge ordering has been reported in another ferrite system, LuFe2O4,811 in the doped manganites such as La1−xCaxMnO3,812–815 PrCaMnO3,816–818 and BiSrCaMnO3,819–821 and in the cuprates, where charge order is thought to be a ubiquitous phenomenon (here most often seen as ordered charge segregation).615,822 Probing and establishing charge order in a system is, in general, challenging; early work on the manganites cited above relied on neutron scattering measurements, which are not directly sensitive to charge, but rather to lattice modulations induced by doping,823 while interpretation of resonant x-ray scattering measurements, which are in principle highly sensitive to charge order, have not been without controversy824,825 and still largely rely on models to interpret the experimental results.826 For example, a pure ionic picture for charge ordering in the manganites would imply that the x-ray absorption spectrum should consist of a superposition of Mn3+ and Mn4+, while the observed spectrum is found to correspond to that of a mixed valency state;827–832 instead, x-ray spectroscopy and neutron scattering measurements suggest the presence of two Mn sites with different magnetic moments and a difference in valency (charge disproportionation) of ∼0.2 electrons,825,826,828,830,833,834 much less than the expected difference of 1 electron between Mn3+ and Mn4+. A similarly complex picture has emerged for magnetite, where the initial ionic model for charge ordering proposed by Verwey806–809 was found not to be consistent with the results of detailed x-ray diffraction and x-ray resonant scattering measurements.835–837 In fact, our understanding of magnetite and of the Verwey transition has improved significantly over the last decade.835,836,838–841 The current understanding is that, at the Verwey temperature, a structural transition from cubic to monoclinic occurs, characterized by a complex lattice distortion that leads to an increase by 2×2×2 of the high temperature cubic unit cell together with a charge localization process whereby the double-exchange active octahedral Fe cations cluster in a superstructure of trimeron complexes with charge and orbital order.835,836,841 The iron valence state varies by 0.6 electrons from +2.35 to +2.95 in contrast to a mixed 3+ and 2+ valence state. This variation is for octahedral sites belonging to the trimeron complex.835 

Given the challenge of probing charge order directly, it is perhaps not surprising that only a few studies have attempted to control charge order using external excitations. Examples include the ultrafast excitation of La1−xCaxMnO3 thin films using high intensity 100 fs laser pulses, where the superlattice reflection associated with orbital/charge order was used to follow the temporal evolution of the system after the excitation;842 under such strong excitation conditions, melting of the orbital/charge order is found to occur on a timescale faster than 200 fs. Another example is the report of a large birefringence effect that is ascribed to charge order in magnetite, controlled by cooling the system under an applied magnetic field.843 In this study, an intense ultra-fast laser pulse was used to achieve a rapid heating of the system to above the Verwey temperature, with the magnetic field setting a preferential orientation for the monoclinic c axis,844 and consequently, of charge order; the fast time response of polarization rotation angle was also used to study the dynamics of charge-order melting, found to be faster than 0.8 ps.843 

Electric field control of magnetism in the solid state has generated much attention recently,37,845–855 motivated by the interest in understanding better the physical processes that lead to a coupling between spin and charge degrees of freedom and also by its potential for applications in next generation electronic devices.344,856–858 In magnetoelectric and in multiferroic materials with magnetic and ferroelectric order, such coupling is intrinsic to the system and intimately linked to the specific origin of ferroelectric polarization.33,36,37,40,845,847,852,859–861 Although exciting materials to study from a basic science level, the potential for applications is severely limited by their scarcity in nature (a consequence of the required time reversal and inversion broken symmetries together with dielectric behavior for ferroelectricity)32,33,862,863 and the generally small value of the effective magnetoelectric coupling or critical temperatures, with BiFeO3 being one outstanding exception.365,369,864–866 In order to circumvent such limitations, an alternative route to achieving a strong coupling between magnetic and ferroelectric order parameters consists in designing artificial multiferroic systems obtained by interfacing magnetic with ferroelectric materials.846,848,850,854,867–869 To reach large magnetoelectric couplings, both the interface characteristics and the coupling mechanism need to be engineered judiciously. Several interfacial magnetoelectric coupling mechanisms have been devised according to the main interaction process exploited: strain, spin exchange, or charge.848,850 Strain-mediated coupling relies on the elastic coupling to modify the magnetic state via magnetoelastic and piezoelectric effects,870–873 while spin exchange-mediated magnetoelectric coupling relies on magnetic exchange interactions, for example, between a ferromagnet and an antiferromagnetic multiferroic via the exchange-bias effect.343,345,348,349,848,874–877 More directly related to field effects is the charge-mediated magnetoelectric coupling, whereby the magnetic state of the system is modified by changing the charge carrier density by using charge screening at the ferroelectric boundary to accumulate or deplete charge carriers at the metallic interface. We consider next materials whose charge screening length lie within three distinct regimes: (i) metallic ferromagnets, characterized by very short Thomas-Fermi screening lengths due to the very large electron charge density at the Fermi level; (ii) 3d correlated systems, with screening lengths and electron mean free paths of the order of the unit cell, such as mixed valence 3d transition metal oxides; and (iii) magnetic semiconductors, where charge screening occurs over the characteristic Debye length LD=ϵsϵ0kBT/(q2n),189 characterized by extended charge screening lengths (ϵs is the static dielectric constant, q the carrier charge, and n the charge carrier density), typically in the nanometer range (ca. 2–20 nm for carrier densities in the range from 1018 to 1016 cm–3 at room temperature).

1. Metallic ferromagnets

In metallic ferromagnets, such as the 3d transition metals, the exchange interaction responsible for magnetic order originates from the Pauli exclusion principle and the condition for antisymmetry of the electronic wave function, which directly links the exchange integral term (electron–electron Coulomb energy) to the spin state of the system.626,878 The Pauli exclusion principle effectively acts to increase electron–electron separation, reducing the Coulomb energy but at a cost of kinetic energy; hence, only in systems where the density of states at the Fermi energy is high it becomes advantageous for a spin-polarized state to emerge, such as in late 3d transition metal elements. The high charge carrier density together with quasi-free electron behavior leads to extremely short Thomas-Fermi screening lengths, of below 1 Å. As a consequence, electrostatic control of magnetism, such as of the exchange interaction or the magnetic moment, which entails a shift in the relative occupancy of the spin up and down subbands,879 may be expected to be modest except for the thinnest films. Such modifications in the interfacial magnetic moment have been estimated for bcc Fe(001), Ni(001), and hcp Co(0001) when subjected to a static electric field applied across a vacuum880 or dielectric barrier.881,882 However, modulation of the charge carrier density at the interface can modify strongly the surface magnetic anisotropy via spin–orbit coupling, as a result of a change in the relative electron occupancy of the interfacial orbital states.880,882–888 The impact of the surface magnetocrystalline anisotropy can be very strong in 3d transition metal ferromagnetic films and nanoparticles, often dominating the magnetic energy.889–892 Effectively, the control of the surface magnetic anisotropy through charge screening has been demonstrated recently in a number of systems, including Fe,884 FePt,893 Co40Fe40B20,894–896 Co,897,898 Ni,899 Fe0.5Pd0.5,900 Fe80Co20,901–903 and in Fe80Co20/MgO/Fe904,905 and Fe40Co40B20/MgO/Fe40Co40B20 tunnel junctions.906,907 A motivation in this context is the electric field control of the relative magnetization alignment of ferromagnetic layers in spin-valve device structures and of the electrical resistivity through the giant (GMR) or tunneling magnetoresistance (TMR) effects for applications in nonvolatile random access memories (MRAM) and logic switches.344,858,908

Additionally, chemical bonding (or hybridization) at the interface can also play an important role. For instance, density functional theory calculations for Fe/BaTiO3(001) show a large surface magnetoelectric response whose origin is largely attributed to changes in the chemical bonding at the interface, i.e., to modulations in the Ti 3d, Fe 3d and O 2p orbital overlap upon reversal of the ferroelectric polarization direction due to the displacement of the Ti atoms, which additionally become magnetically polarized.909–914 Similar effects have been calculated for Co/PZT,915 and also for oxides such as LSMO/BaTiO3,160,799 suggesting that such a process may be quite general at ferroelectric interfaces. In fact, magnetic polarization of the Ti interface layer in contact with ferromagnetic systems has been observed in several systems, including Fe/BaTiO3(001),916 LSMO/BaTiO3,160,917 and Co/PZT,918 turning the interface region of the ferroelectric also ferromagnetic, i.e., multiferroic.

2. Mixed valence transition metal oxides

In the 3d transition metal oxides, the magnetic exchange interaction between cations is indirect, mediated by the oxygen anions, either through virtual excitations of spin up-down pairs to neighboring cations in the case of superexchange interaction (which tends to favor antiferromagnetic coupling)919 or through electron hopping in the case of double exchange interaction, which favors ferromagnetic coupling.741,920–922 Although indirect, magnetic interactions in many oxides can be strong and result in high magnetic critical temperatures, for example, up to 858 K for the ferrimagnetic spinels NiFe2O4 and Fe3O4.840,923,924 In comparison to magnetic metals and magnetic semiconductors, complex oxides tend to exhibit stronger electron–electron correlations and a more diverse electronic behavior as a function of charge carrier density, where magnetism competes or co-exists with other types of order, such as orbital order or superconductivity, leading to complex phase diagrams.590,593,784,925–927 Typically, the effect of doping in the 3d metal oxides, for example via chemical substitution, is that of modifying the average valence state of the 3d cations and modifying the charge carrier density. The main approach to electrostatic control of magnetism in such mixed-valence systems is to choose a chemical doping that places the system at a phase boundary between two magnetic ground states of the system and to drive it across the boundary by electrostatic modulation of the charge carrier density. Also here it is important to recall that the phase diagram of thin films can deviate significantly from that of the bulk material, for example, due to epitaxial strain, as illustrated by the changes in the magnetic and conducting state of La1−xSrxMnO3 films near 0.5 doping with strain shown in Fig. 18.748,749

FIG. 18.

Diagram of the magnetic state of the LSMO films grown on LSAT (circles) and SrTiO3 (triangles) as a function of lattice mismatch (ϵ) and doping. The dotted line separates ferromagnetic from antiferromagnetic ground states. Inset shows the variation of the magnetic moment, μ, vs ϵ (dashed line is guide to the eye). Reproduced with permission from Vaz et al., Phys. Rev. B 90, 024414 (2014). Copyright 2014 American Physical Society.

FIG. 18.

Diagram of the magnetic state of the LSMO films grown on LSAT (circles) and SrTiO3 (triangles) as a function of lattice mismatch (ϵ) and doping. The dotted line separates ferromagnetic from antiferromagnetic ground states. Inset shows the variation of the magnetic moment, μ, vs ϵ (dashed line is guide to the eye). Reproduced with permission from Vaz et al., Phys. Rev. B 90, 024414 (2014). Copyright 2014 American Physical Society.

Close modal

Among the mixed valence materials, 3d metal oxides crystallizing in the perovskite structure have generated notable interest due to their diverse and important chemical and physical properties.928–931 Within this class of materials, the manganites have been intensively investigated,932,933 particularly since the observation of the CMR effect,743–747 effectively, a magnetic field-driven insulator to metal transition. The phase diagrams of the manganites are characterized by several phase transitions as a function of chemical doping, for example, for La1−xSrxMnO3, from an insulating orbital ordered antiferromagnetic (x = 0), to insulating ferromagnetic (0<x<0.2), metallic ferromagnetic (0.2<x<0.5), metallic antiferromagnetic (x = 0.5) to antiferromagnetic insulating (x > 0.5).769,934,935 In many manganite perovskites, the equilibrium phases near the boundaries are quasi-degenerate in energy and coexist,593,770–772 suggesting that the equilibrium state of the system can be tilted toward one or the other state with only modest perturbations. The charge carrier density at the boundaries separating such phase transitions is high, typically in the range of 1019–1022 cm–3, requiring a correspondingly large charge carrier modulation for controlling the equilibrium state.3 In addition, the system should be as thin as possible for maximal field effect, ideally, of the order of the charge screening length. Experimentally, it is known that the magnetic and electrical properties of thin films of the manganites grown on SrTiO3 are strongly depressed below a thickness of 1–3 nm,862,936–944 while epitaxial strain is also known to strongly affect the magnetic and electrical properties (Fig. 18).748,749 Hence, the strategy for modulating the magnetic properties of the mixed valence manganites using the ferroelectric field effect has been to setting the thickness of the manganite layer at the transition between insulator to metallic state; for example, for La1−xSrxMnO3/SrTiO3(001) at x = 0.2 and x = 0.3, this occurs at a film thickness of about 10 unit cells.760,940,945 Using this approach, demonstration of electric field control of the magnetic properties of the doped manganites in ferroelectric field effect structures could be achieved, including modulations in the critical temperature,760,761 anisotropic magnetoresistance596 and in the magnetic polarization.946 

An important breakthrough was the experimental report of a large modulation of the coercivity, magnetic critical temperature, and magnetic moment of a 4 nm La0.8Sr0.2MnO3/PZT heterostructure as a function of the applied electric field measured by magneto-optic Kerr effect (MOKE) magnetometry, demonstrating the presence of a large magnetoelectric coupling in this artificial multiferroic system that was both hysteretic and nonvolatile.317 The results are in qualitative agreement with what is expected from the bulk phase diagram of La1−xSrxMnO3,769,934,935 including a higher critical temperature and lower magnetic moment with increasing doping. Subsequent x-ray absorption near-edge spectroscopy (XANES) measurements showed that switching the ferroelectric polarization resulted in a shift in the absorption edge of the Mn K-edge by about 0.3 eV, interpreted as a change in the valence state of the interfacial Mn layer at the PZT interface when switching between the accumulation and depletion states.315,945 Such measurements demonstrated the electronic origin of the magnetoelectric effect in PZT/LSMO and allowed a quantitative comparison of the changes in the magnetic moment expected from a simple modulation of the Mn valence state (of 0.1 μB/Mn) and that measured experimentally (0.76 μB/Mn), showing that an additional change in the spin configuration at the interface takes place, from a ferromagnetic coupling in the depletion state to an antiferromagnetic coupling in the accumulation state, and attributed to a change in the relative weights of the superexchange and double exchange energy terms as a function of the 3dz2 orbital occupancy induced by screening. The presence of such a magnetic reconstruction at the interface is supported by the results of first principles calculations.947–953 Also significant, the hole accumulation state shows a much higher critical temperature than the depletion state (by about 20 K), such that over a large temperature window on/off switching of magnetism driven by the orientation of the ferroelectric polarization is realized.954,955 Subsequent studies reported similar modulations of the magnetization in La1−xSrxMnO3-ferroelectric heterostructures at various doping levels and employing different ferroelectric systems, examples of which are summarized in Table VII, showing that the magnetoelectric effect is not very sensitive to the exact amplitude of the ferroelectric polarization and that charge modulations in the range between 25 and 100 μC/cm2 at the LSMO interface are sufficient to modify the interfacial magnetic state. However, a number of discrepant results for the magnetic and electronic behavior has been observed, such as the case for BaTiO3/LSMO/SrTiO3(001) at x = 0.3 doping, where an opposite change in the saturation moment with the direction of the ferroelectric polarization is observed and no change in the critical temperature,956 an effect attributed to a modulation in the relative population of the ferromagnetic metallic and antiferromagnetic insulating phases coexisting in the system. One advantage of La1−xSrxMnO3 at x = 0.3 doping resides in its relatively high bulk Curie temperature, of about 370 K, such that room temperature demonstration of a large change in the magnetic moment would be of high practical interest.

TABLE VII.

Magnetoelectric coupling in manganite–ferroelectric heterostructures. LSMO: La1-xSrxMnO3, STO: SrTiO3, BFO: BiFeO3, BTO: BaTiO3.

SystemxtAMnO3(u.c.)PsFE (μC/cm2)Tcacc (K)ΔTc (K)msdep (μB/Mn)msdep/msaccT (K)Obs.Reference
PZT/LSMO/Nb:STO(001) 0.12 18.5a 25 250 ∼2.28a 1.02 50 SQUID 958  
PZT/LSMO/STO(001) 0.175 13a 100 263 15 3.63a 1.03 25 SQUID 959  
LSMO/PZT/LSMO/STO(001) 0.175 15a 80 ⋯ ⋯ 3.6 1.21 4.2 XAS 800  
PZT/LSMO/STO(001) 0.2 10 45 197 20 0.88 1.22 100 MOKE 317  
PZT/LSMO/STO(001) 0.2 10 85 200 20 3.30 1.30 100 MOKE 315  
LSMO/PZT/LSMO/STO(001) 0.2 33 80 ⋯ ⋯ 3.76 8.91 120 PNR 960  
BFO/LSMO/STO(001) 0.2 15a ⋯ 275 ∼10 3.20 1.11 10 SQUID 961  
BFO/LSMO/STO(001)b 0.3 13 65 340 3.39 1.10 10 SQUID 962  
BFO/LSMO/STO(001)c 0.3 13 65 320 3.01 1.24 10 SQUID 962  
BTO/LSMO/STO(001) 0.33 64a ⋯ 345 2.93 0.94 200 SQUID 956  
BTO/LSMO/STO(001)d 0.33 50 170 60 1.40 1.48 10 SQUID 160  
PZT/LSMO/STO(001) 0.33 13a ⋯ ⋯ ⋯ 0.5f 0.67 298 PNR 963  
BFO/LSMO/STO(001) 0.33 15a ⋯ 300 ∼5 2.56 1.14 10 SQUID 961  
PZT/LSMO/STO(001)e 0.33 5a ∼80 123 10 1.22a 1.17 10 SQUID 964  
BFO/LSMO/STO(001) 0.5 15a ⋯ 280 ∼0 1.60 1.25 10 SQUID 961  
SystemxtAMnO3(u.c.)PsFE (μC/cm2)Tcacc (K)ΔTc (K)msdep (μB/Mn)msdep/msaccT (K)Obs.Reference
PZT/LSMO/Nb:STO(001) 0.12 18.5a 25 250 ∼2.28a 1.02 50 SQUID 958  
PZT/LSMO/STO(001) 0.175 13a 100 263 15 3.63a 1.03 25 SQUID 959  
LSMO/PZT/LSMO/STO(001) 0.175 15a 80 ⋯ ⋯ 3.6 1.21 4.2 XAS 800  
PZT/LSMO/STO(001) 0.2 10 45 197 20 0.88 1.22 100 MOKE 317  
PZT/LSMO/STO(001) 0.2 10 85 200 20 3.30 1.30 100 MOKE 315  
LSMO/PZT/LSMO/STO(001) 0.2 33 80 ⋯ ⋯ 3.76 8.91 120 PNR 960  
BFO/LSMO/STO(001) 0.2 15a ⋯ 275 ∼10 3.20 1.11 10 SQUID 961  
BFO/LSMO/STO(001)b 0.3 13 65 340 3.39 1.10 10 SQUID 962  
BFO/LSMO/STO(001)c 0.3 13 65 320 3.01 1.24 10 SQUID 962  
BTO/LSMO/STO(001) 0.33 64a ⋯ 345 2.93 0.94 200 SQUID 956  
BTO/LSMO/STO(001)d 0.33 50 170 60 1.40 1.48 10 SQUID 160  
PZT/LSMO/STO(001) 0.33 13a ⋯ ⋯ ⋯ 0.5f 0.67 298 PNR 963  
BFO/LSMO/STO(001) 0.33 15a ⋯ 300 ∼5 2.56 1.14 10 SQUID 961  
PZT/LSMO/STO(001)e 0.33 5a ∼80 123 10 1.22a 1.17 10 SQUID 964  
BFO/LSMO/STO(001) 0.5 15a ⋯ 280 ∼0 1.60 1.25 10 SQUID 961  
a

Estimated using 1 u.c. = 3.9 Å.

b

With a BiO-MnO2 interface stacking.

c

With a FeO2-(La,Sr)O interface stacking.

d

Different ferroelectric polarization states achieved by changing substrate termination.

e

Different ferroelectric polarization states achieved by using the flexoelectric effect.

f

Using the conversion ratio for the neutron magnetic scattering length of 106μB2 for LSMO from Ref. 957.

The exact interfacial spin configuration in the different types of heterostructures discussed above still remains to be fully elucidated, in particular, since several possibilities for the ground state of the system in the accumulation state have been predicted theoretically.2,948,949,951,952 It is clear that to determine the interfacial spin configuration in such heterostructures requires local probes of the spin configuration. For example, Ma et al.965 probed the interfacial magnetic signal of PZT/LSMO/LaAlO3(001) at x = 0.3 doping with second harmonic signal generation to show that the second harmonic signal is absent in the accumulation state (no net magnetization, suggesting an interfacial antiferromagnetic state) and present in the depletion state (corresponding to a uniform ferromagnetic state across the interface), consistent with the results for PZT/LSMO/SrTiO3(001) at x = 0.2 described above (although the strain state of the LSMO film may differ in the two cases).315,317 Another local probe of magnetism is polarized neutron reflectivity (PNR), which enables one to measure the magnetization profile of buried films thanks to the high penetration depth of neutron beams and the intrinsic magnetic moment of the neutron.966–970 The study by Meyer et al.960 on various LSMO/ M/LSMO/SrTiO3(001) (x = 0.2) heterostructures, with M = PZT or LaAlO3, could identify strongly varying magnetic moments in the LSMO films as the layer structure grows in complexity (i.e., from the single La0.8Sr0.2MnO3 layer to the full multilayer stack), showing that adding a capping layer to the LSMO enhances the magnetic moment in the bulk of the film and that the presence of a ferroelectric interface enhances the magnetic moment in the charge-depleted interface and reduces it in the charge-accumulated interface (both in absolute terms and also with respect to the bulk film magnetic moment); the control sample where PZT is replaced by a non-ferroelectric LaAlO3 film still shows a reduction of the moment of the top (uncapped) LSMO film but no moment enhancement at the LaAlO3/La0.8Sr0.2MnO3 interface, as shown in Fig. 19. The results of a very recent study by Liu et al.964 combining electrical, magnetic, spectroscopic, and neutron transport (PNR) characterization are also included in Table VII for PZT/LSMO (x = 0.33), where the direction of the PZT ferroelectric polarization was configured via the flexoelectric effect. In this case, different LSMO films were prepared with the PZT polarization pointing into or away from the LSMO interface; the LSMO films at 2 nm thickness were found to exhibit the strongest magnetolectric effects and consistent with the results for 0.2 doping; in particular, the films at this thickness are found to exhibit two magnetic transitions, the lower ones associated with the presence of the ferroelectric layer. By using PNR to evaluate the magnetic profile of the two 2 nm LSMO films with different polarization orientations, the authors find the presence of an oppositely aligned interfacial magnetic layer for the accumulation state and a slightly enhanced magnetic moment for the depletion state, in agreement with the model proposed by Vaz et al.315 and in agreement with DFT calculations.948,950–953 The XAS results indicate that while the depletion state favors the occupation of 3z2r2 states, the accumulation state favors occupation of x2y2 states and an antiferromagnetic coupling with the dominant x2y2 states in the bulk of the films, favored by epitaxial strain.964 In another PNR study of a PZT/LSMO (x = 0.33) bilayer structure,963 the PNR spectra were taken for the two polarization states of the PZT, where the ferroelectric polarization was switched by using a Pt electrode deposited on the top surface of the PZT film. The change in the magnetization profile shows a large reduction of the interfacial moment of LSMO when going from the accumulation to the depletion state, i.e., opposite to that found for LSMO at x = 0.2 doping. The measurements were carried out while the electric field was applied to the PZT layer and although one could expect the strain state of the PZT to be identical for both states of the ferroelectric polarization, x-ray diffraction measurements of the out of plane lattice parameter of the PZT layer indicate an asymmetric behavior for positive and negative voltages, such that the observed magnetic changes are explained as a combination of charge and strain effects; the results of ab initio calculations support the conclusion that strain is the leading factor responsible for the observed experimental behavior.963 

FIG. 19.

Polarized neutron reflectivity spectra and magnetic depth profiles of [(a)–(c)] LSMO/PZT/LSMO/STO(001) and [(d)–(f)] LSMO/LAO/LSMO/STO(001) heterostructures obtained from fits to PNR data. Reproduced with permission from Meyer et al., Phys. Rev. B 94, 174432 (2016). Copyright 2016 American Physical Society.

FIG. 19.

Polarized neutron reflectivity spectra and magnetic depth profiles of [(a)–(c)] LSMO/PZT/LSMO/STO(001) and [(d)–(f)] LSMO/LAO/LSMO/STO(001) heterostructures obtained from fits to PNR data. Reproduced with permission from Meyer et al., Phys. Rev. B 94, 174432 (2016). Copyright 2016 American Physical Society.

Close modal

In a detailed study of the spin and orbital state of 5 u.c. La2/3Sr1/3MnO3/BaTiO3 heterostructures by XAS and HR-STEM, Wang et al.160 have observed a dependence of both magnetism and orbital polarization in both LSMO and BaTiO3 interfacial layers with the direction of the ferroelectric polarization. Samples with different directions of the ferroelectric polarization were prepared by engineering the interface of the SrTiO3(001) substrate, by either using a SrO termination, which favors the state with the polarization pointing away from the SrTiO3 interface, or using a TiO2 surface termination, which favors the opposite polarization direction. The results of magnetic x-ray absorption spectroscopy at the Mn L-edge show a stronger magnetic signal for the state pointing away from the La1−xSrxMnO3, corresponding to depletion of hole carriers from the La1−xSrxMnO3 interface, compared to the accumulation state; in addition, a similar increase in the magnetic signal amplitude is observed at the Ti L-edge, albeit with a change in signal for the depletion state, as shown in Fig. 20. Interestingly, using the XMCD sum rules to estimate the spin and orbital components of the magnetic moment, they find that the orbital moment is significant for the depletion state, but zero for the accumulation state. These results were explained in terms of a change in the magnetic configuration from ferro- to antiferromagnetic when going from the depletion to accumulation state, in the latter case, supported by the observation of a small exchange bias in the system (of up to 20 Oe at 10 K, which is not observed for a similar La1−xSrxMnO3 film grown on SrTiO3 without the BaTiO3 film).

FIG. 20.

Normalized x-ray absorption spectroscopy and x-ray magnetic circular dichroism of LSMO/BTO heterostructures at 80 K and ±1 T at the Mn (top panels) and Ti (bottom panels) L2,3 edge, for accumulation (left panels) and depletion (right panels) of hole carriers from the LSMO interface. Adapted with permission from Wang et al., Nano Lett. 19, 3057 (2019). Copyright 2019 American Chemical Society.

FIG. 20.

Normalized x-ray absorption spectroscopy and x-ray magnetic circular dichroism of LSMO/BTO heterostructures at 80 K and ±1 T at the Mn (top panels) and Ti (bottom panels) L2,3 edge, for accumulation (left panels) and depletion (right panels) of hole carriers from the LSMO interface. Adapted with permission from Wang et al., Nano Lett. 19, 3057 (2019). Copyright 2019 American Chemical Society.

Close modal

Additional important insights into the electronic and magnetic modifications at manganite-ferroelectric interfaces have been gained as well by local atomic characterization using high-resolution scanning transmission electron microscopy (HR-STEM), particularly in combination with atomically resolved electron energy loss spectroscopy (EELS), by following the evolution of the energy loss spectra as a function of distance from the interface (although interpretation of such variations in terms of changes in valency is challenging).28,144,971,972 In addition, other important information, such as local atomic spacing and oxygen octahedra rotations, can be further related to other properties, such as crystalline structure, oxygen stoichiometry, and ferroelectric polarization;973,974 further, by combining STEM and EELS, the specific interface stacking at the interface from one material to the other can be determined.162 For example, while XANES measurements demonstrated a change in the valence state in the LSMO/PZT interface, as discussed above, they do not allow one to determine precisely the charge distribution across the interface.315,945 Such layer-resolved local measurements of the Mn valence state across ferromagnetic/manganite interfaces using STEM-EELS have been reported meanwhile for PZT/LSMO,147,156–158 BaTiO3/LSMO,159,160 and BiFeO3/CaCeMnO3.161 STEM-EELS studies of LSMO/PZT/LSMO,SRO/STO(001) heterostructures, where the two bottom PZT interfaces (LSMO or SRO) are used for setting the direction of the PZT polarization, indicated a drop in Mn valency at the top LSMO (x = 0.3)/PZT interface layer when the ferroelectric polarization points to that interface (corresponding to depletion of charge carriers), with an opposite behavior when the PZT polarization points in the other direction;156 changes in the critical temperature and magnetic moment are consistent with the results at 0.2 doping. In a follow up study,147 both interfaces of a symmetric LSMO/PZT/LSMO/SrTiO3(001) structure were characterized for one direction of the ferroelectric polarization (although the STEM results show that the interfaces are structurally slightly different, with the top LSMO/PZT interface appearing more diffuse when compared to the atomically sharp bottom interface). The STEM-EELS results suggest a drop in the Mn valency over a distance of about 1 nm from both PZT interfaces, a surprising result especially for the bottom interface, where an increase in the valence state is expected based on the direction of the ferroelectric polarization;28 complementary polarized neutron reflectometry measurements carried out to probe the magnetic depth profile of the heterostructure give a much larger magnetic moment also at the bottom interface, a behavior opposite to that found for PZT/LSMO at 0.2 doping315,317,945 but consistent with the results for BaTiO3/LSMO/SrTiO3(001) at x = 0.3 doping.956 STEM-EELS studies of BaTiO3/LSMO superlattices for several values of the LSMO thickness (4, 10, and 32 u.c.) by Guo et al.159 also found a strong decrease in the Mn valency near both BaTiO3 interfaces. A recent STEM-EELS study of the PZT/12 u.c. LSMO (x = 0.2)/SrTiO3(001) interface structure by Meng et al.158 has determined the change in the charge transfer across the various interfaces, estimated from the shift in the plasma peaks of valence EELS. The results indicate a strong variation in the valence charge, extending through the whole LSMO layer, 4 nm into the PZT, and about 2 nm into the SrTiO3 substrate interface.

An additional advantage of STEM is the capability of probing the atomic structure in real space and, in particular, of providing important insights into the electronic characteristics of interfaces and of manganite-ferroelectric interfaces. This is important, since several studies have shown that ferroelectric structural distortions can cross the interface between the ferroelectric and the conducting channel and continue for a few atomic layers into the channel region.159,500,501,799,800,951,975 For strongly correlated systems, whose physical properties are strongly coupled to the lattice structure, a ferroelectrically induced structural distortion can be expected to induce large modifications of the electronic and magnetic properties. Indeed, a recent STEM study pointed to the important role of changes in the local interfacial atomic structure induced by the polar cations of the ferroelectric in modifying the interfacial orbital occupancy of LSMO via local strain for the accumulation state (see Fig. 16)799 while a direct confirmation of the change of the orbital occupancy with the direction of the ferroelectric polarization was confirmed in LSMO/PZT (x = 0.175)800 and BaTiO3/LSMO (x = 1/3)800 by measuring the changes in the linear dichroic response of the system at the Mn L-edge, ascribed to modifications in the interfacial atomic structure that favors occupation of 3d x2y2 orbitals for the accumulation state and 3d z2 orbitals for the depletion state (results that were discussed in Sec. III C in the context of electric field control of the orbital polarization, see Fig. 17).

While ferroelectric field effect control of magnetism in the oxide manganites has received the bulk of the attention, field effects in other oxide systems have also been investigated, including magnetite (Fe3O4) and strontium ruthenate (SrRuO3). Magnetite is a conducting mixed valence oxide, where, in an ionic picture, electron transport occurs through electron hopping between Fe2+ and Fe3+ ions on the octahedral sites; since the ions in this sublattice are coupled ferromagnetically, magnetite is expected to be fully spin-polarized.840 In principle, as in the doped manganites, changes in the electron charge density should result in changes in the magnetic moment and in the magnetic properties and, effectively, such an effect has been demonstrated via ionic (Li+) doping.597–599 However, while control of magnetism by interfacing with a ferroelectric has been extensively reported for strain-mediated coupling, aiming at taking advantage of the high magnetoelastic coupling coefficients of magnetite,976,977 electrostatic control of its magnetic properties has been studied less often.978–980 For example, Wu et al.979 have prepared a 5 nm thick, [001]-oriented Fe3O4 film sandwiched between a 200 nm thick SrTiO3 on top and a bottom Au/PZT bilayer, which enabled them to study the electric field response to both strain (via Au/PZT) and electric field (via the SrTiO3): they found that, while strain modifies the shape of the hysteresis loop (i.e., the magnetic anisotropy), charge modulation affects mainly the saturation magnetization.

Ferromagnetic strontium ruthenate (SrRuO3) is another important material where structural distortions can impact strongly its physical properties. Recent reports have shown that ultrathin SrRuO3 films and SrRuO3-based heterostructures can be a fertile playground for studying novel skyrmion spin structures. Skyrmions are winding spin textures akin to magnetic bubbles981,982 but endowed with chirality in systems with broken inversion symmetry by the Dzyaloshinskii–Moriya interaction (DMI).983 They have been observed in low symmetry bulk crystals984,985 and in asymmetric metallic multilayers between 3d transition metal ferromagnetic and heavy metals with large spin–orbit coupling.986–991 The latter approach to inducing skyrmion textures has been extended to SrRuO3,992–995 a 4d transition metal magnetic oxide with an intrinsic spin–orbit coupling of 0.1–0.15 eV.996 In particular, the work by Wang et al.997 showed that, in a BaTiO3/SrRuO3 heterostructure, a ferroelectrically-driven non-centrosymmetric structural distortion in the SrRuO3 layer can trigger a DMI sufficiently large to induce the formation of skyrmion spin structures. Furthermore, they showed that the skyrmion density and stability can be controlled by switching the ferroelectric polarization due to the asymmetric distribution of the latter across the heterostructure. The ability to ferroelectrically control skyrmions in BaTiO3/SrRuO3 heterostructures opens the possibility to realize novel spintronic devices based on ferroelectric materials. It is also worth noting that the ferroelectrically-driven broken inversion symmetry not only leads to the emergence of skyrmions, but also enhances the ferromagnetic properties of SrRuO3.992,998

3. Magnetic semiconductors

Magnetic semiconductors, such as EuO or the dilute magnetic semiconductor (Ga,Mn)As, are characterized by the presence of a bandgap at the Fermi level and a relatively low density of thermally activated charge carriers (intrinsic or from dopants). The relevant screening length in these semiconducting systems is the Debye length, which, depending on the carrier density, is typically in the nm range or larger. The origin of magnetism varies with the particular system. In dilute magnetic semiconductors, such as (In,Mn)As, (Ga,Mn)As, and MnxGe1−x, magnetism is driven by the p-d exchange interaction between delocalized hole carriers (from the Mn dopant) and the localized spins of the Mn atoms.999–1006 For sufficiently high hole carrier density (of the order of a few %), a long-range ferromagnetic state develops. In particular, control over the magnetic exchange interaction can be obtained by modulating the carrier density electrostatically, as demonstrated, for example, by gating (In,Mn)As,1007,1008 MnxGe1−x,1009–1011 (Zn,Mn)Se,1012 and (Ga,Mn)As.1013–1019 Given the much lower charge carrier density and the large extent of the depletion layer, large modulations of the charge carrier density as provided by ferroelectric gates are not essential and the process is not interfacial in nature, although a high quality interface is still required in order to minimize charge traps. However, the non-volatility provided by a ferroelectric gate could still be of much interest as is the case for ferroelectric FETs.

The reversible control of magnetism has been demonstrated also in the dilute magnetic insulator TiO2 (anatase) doped with Co,1020 where both a change in the saturation magnetization and coercive field is observed in PZT/Co:TiO2/SrRuO3 heterostructures as a function of the PZT ferroelectric polarization direction. Magnetism in the insulator Co:TiO2 is not mediated by charge carriers, as in the dilute magnetic semiconductors; instead, a bound magnetic polaron percolation picture1021–1023 and/or a defect (F-center) state percolation model1024 have been suggested to drive ferromagnetic order in this system.1020 

Ferroelectric tunnel junctions (FTJs) were conceptualized in the early 1970s by Leo Esaki,1025 although experiments on samples using established ferroelectric films thin enough to act as tunnel barriers were carried out only 30 years later.67 The reasons for the time lag include the difficulty to achieve good tunnel barriers (solid-state quantum mechanical tunneling could be achieved early on, for example in superconducting tunnel junctions,1026,1027 but magnetic tunnel junctions, for instance, were only developed in the 1990s1028) and to stabilize ferroelectricity in films just a few nm thick (which was only realized in the 2000s).127,466,1029,1030 However, Esaki already anticipated that FTJs would exhibit what is now called tunnel electroresistance (TER), a dependence of the tunnel resistance (or current) on the ferroelectric polarization direction, an effect that can be explored for nonvolatile memory devices. Importantly, FTJs present an essential advantage over ferroelectric random access memories (FeRAMs): reading the logic state (encoded by the junction's resistance) does not alter the state, resulting in simplified architectures.

The first clear demonstration of TER in FTJs was provided in 2009 by Garcia et al.67 Working with ultrathin epitaxial layers of BaTiO3 grown onto La2/3Sr1/3MnO3 bottom electrodes, they used a scanning probe microscope to electrically write micrometer-scale ferroelectric domains with the polarization pointing either up or down, image these domains by piezoresponse force microscopy (PFM), and map the local resistance for both polarization directions. These resistance maps are displayed in Fig. 21(a) for three BaTiO3 barrier thicknesses. While for the lowest thickness (1 nm) only a weak resistance contrast is visible, at 2 nm and 3 nm, the resistance shows a strong dependence on the polarization direction, with a contrast reaching 75000% at room temperature. Importantly, both the tunnel resistance and the TER were found to increase systematically with the barrier thickness, indicating that the BaTiO3 films are good tunnel barriers and that the changes in resistance are linked to the tunneling process.

FIG. 21.

(a) Conductive AFM resistance maps of BTO/LSMO across 1 × 4 μm2 poled domains for (from left to right) 1–3 nm-thick BTO barriers. (b) Tunnel barrier potential profiles modified by imperfect screening in the electrodes (good metal on the right side of the barrier with a short screening length δ2; poor metal or semiconductor on the left with a long screening length δ1). The dashed purple line shows the average barrier height ϕ+ or ϕ for both polarization directions. From Ref. 67.

FIG. 21.

(a) Conductive AFM resistance maps of BTO/LSMO across 1 × 4 μm2 poled domains for (from left to right) 1–3 nm-thick BTO barriers. (b) Tunnel barrier potential profiles modified by imperfect screening in the electrodes (good metal on the right side of the barrier with a short screening length δ2; poor metal or semiconductor on the left with a long screening length δ1). The dashed purple line shows the average barrier height ϕ+ or ϕ for both polarization directions. From Ref. 67.

Close modal

Three possible physical mechanisms responsible for the TER had been discussed earlier by Tsymbal and Kohlstedt.6 One is based on piezoelectricity, causing the barrier thickness to change under the poling voltage and affecting the tunnel resistance; a second mechanism is related to changes of the density of states at the barrier/electrode interfaces; a third process, which appears to usually dominate the TER response, is electrostatic and was initially proposed by Zhuravlev et al.1031 It is described schematically in Fig. 21(b). At the interface between the ferroelectric barrier and the electrodes, polarization charges must be screened; if the electrode is a good metal, with a short screening length [δ2 in Fig. 21(b)], the screening is efficient and the electrostatic potential at the interface is only weakly modified; however, if the electrode is a semiconductor or a bad metal with a longer screening length [δ1 in Fig. 21(b)], the potential is substantially modified. As a result, in an FTJ combining a good metal and a bad metal sandwiching the tunnel barrier, the barrier potential profile is asymmetric, with an average potential barrier height [purple dashed line in Fig. 21(b)] that depends on the polarization direction. Because the tunneling current depends exponentially on the square root of the barrier height, switching the ferroelectric polarization results in a large change of the tunnel resistance, producing the TER effect.

In this picture, the TER is therefore an interface-driven effect and FTJs qualify as interfacial ferroelectric devices. Among the vast literature on FTJs,1032,1033 several studies aimed at addressing these interfacial effects in detail. Two examples are presented in Fig. 22. Figures 22(a)–22(d) showing the resistance at low dc voltage vs the amplitude of short (10 ns) voltage pulses applied to solid-state FTJs based on tetragonal BiFeO3 ferroelectric barriers grown on a bottom electrode of (Ca,Ce)MnO3.1034 A hysteretic dependence is obtained, reminiscent of the ferroelectric switching (as checked by measuring piezoelectric loops on the same device). The TER amplitude is found to depend strongly on the nature of the metal used as the top electrode, and in fact roughly scaling with the metal work function [Fig. 22(g)]. This reflects the dependence of the TER on the (average) tunnel barrier height, which increases for high work function metals such as Ir. A second example, displayed in Fig. 22(e), shows a similar resistance vs pulse voltage cycle for Pt/BaTiO3/Nb-SrTiO3 junctions.1035 Here, the Nb-doped SrTiO3 bottom electrode has a low carrier density and therefore a large screening length. The amplitude of the TER increases with decreasing Nb doping level, i.e., with increasing the screening length, as shown in Fig. 22(f). As explained above, a strong difference of screening lengths between the top and bottom electrodes should lead to a strong barrier asymmetry and to a larger TER, in agreement with Zhuravlev's model.1031 

FIG. 22.

[(a)–(d)] Representative resistance hysteresis cycles for M/BiFeO3/(Ca,Ce)MnO3 junctions with different top electrodes M = W, Co, Ni, and Ir, starting from the high resistive state. (e) Resistance hysteresis loops of Pt/BaTiO3/Nb-SrTiO3 junctions. The loop composed of dark squares (green circles) is obtained by using the writing pulse train “w” between +3.6 and –3.6 V shown schematically in the bottom-left (top-right) inset and measured with read pulses “r” of +0.1 V following each write pulse. (f) On/off ratio in Pt/BaTiO3/Nb-SrTiO3 FTJs for different Nb concentrations. (g) Off/on ratio as a function of the work function of the metal used as a top electrode in (a)–(d). [(a)–(d)] Boyn et al., APL Mater. 3, 061101 (2015). Copyright 2015 Author(s), licensed under a Creative Commons Attribution (CC BY) license. [(e) and (f)] Reproduced with permission from Wen et al., Nat. Mater. 12, 617 (2013). Copyright 1998 Macmillan Publishers Limited.

FIG. 22.

[(a)–(d)] Representative resistance hysteresis cycles for M/BiFeO3/(Ca,Ce)MnO3 junctions with different top electrodes M = W, Co, Ni, and Ir, starting from the high resistive state. (e) Resistance hysteresis loops of Pt/BaTiO3/Nb-SrTiO3 junctions. The loop composed of dark squares (green circles) is obtained by using the writing pulse train “w” between +3.6 and –3.6 V shown schematically in the bottom-left (top-right) inset and measured with read pulses “r” of +0.1 V following each write pulse. (f) On/off ratio in Pt/BaTiO3/Nb-SrTiO3 FTJs for different Nb concentrations. (g) Off/on ratio as a function of the work function of the metal used as a top electrode in (a)–(d). [(a)–(d)] Boyn et al., APL Mater. 3, 061101 (2015). Copyright 2015 Author(s), licensed under a Creative Commons Attribution (CC BY) license. [(e) and (f)] Reproduced with permission from Wen et al., Nat. Mater. 12, 617 (2013). Copyright 1998 Macmillan Publishers Limited.

Close modal

Over the last decade, progress in FTJ research has led to several substantial advances toward device applications. TER values in excess of 108% (off/on ratio >106),975 scaling down to sub 100 nm sizes1036 and compatibility with Si have been demonstrated.1037 Significantly, FTJs based on ferroelectric phases based on HfO2 have also been reported,1038–1040 paving the way toward the integration of FTJs into CMOS-based processes.

Beyond their possible use as binary nonvolatile memory devices, FTJs have also emerged as interesting candidates for analog-based, neuromorphic computation. Indeed, and as visible in the loops shown in Fig. 22, FTJs can possess intermediate levels of resistance, in addition to the two saturation values.373,1041 In fact, there is a quasi-continuum of resistance states between the on and off states that can be stabilized by applying voltage pulses of intermediate amplitudes373,1041 or trains of pulses of moderate amplitude.77 Importantly, these resistance states are remanent and the resistance of the device only changes when the applied voltage exceeds specific threshold values. These two ingredients qualify FTJs as memristors,1042,1043 electronic analogs of the synapses in the brain whose information retention (non-volatility) and plasticity properties are key to the learning process.76 The junction conductance plays the role of the synaptic weight in biologic synapses, which is modified as the synapses receive voltage spikes from their neighboring neurons. Compared to biological synapses, FTJs have a higher operation speeds (in the 10–100 ns range vs ms for biological synapses) and very small size (down to nanometer scale for FTJs). These properties allow one to use FTJs to implement specific learning rules observed in biological systems. One is spike-timing dependent plasticity (STDP).76 Through STDP, the synaptic weight varies according to the time difference Δt between pulses received by the synapses from the pre- and post-synaptic neurons they connect. When the pre-synaptic neuronal spike reaches the synapse before the post-synaptic neuronal spike (Δt>0), the synaptic weight typically increases, while it decreases for Δt<0. STDP can be implemented in memristors by mimicking the neuronal spikes by voltage waveforms having specific shapes. In the example of Fig. 23(a), one voltage waveform is comprised of a rectangular positive wavefront and a negative trapezoidal long slope. Within a single pulse the voltage never exceeds the memristor's threshold for switching. When two waveforms are combined to mimic the reception of neuronal spikes coming from the pre-synaptic and the post-synaptic neurons, they add up so that if |Δt|>0 is sufficiently short, the threshold voltage will be exceeded and the memristor conductance will change (increasing or decreasing depending on the sign of Δt). However, if Δt is long, the waveforms do not overlap, the switching threshold is not reached and the conductance remains unchanged.

FIG. 23.

(a) Measurements of STDP in a Co/BiFeO3/(Ca,Ce)MnO3 ferroelectric memristor. Modulation of the device conductance (ΔG) as a function of the delay (Δt) between pre- and post-synaptic spikes. Seven datasets were collected on the same device showing the reproducibility of the effect. The total length of each pre- and post-synaptic spike is 600 ns.77 (b) STDP measurements in a HZO-based FTJ. The top panel shows the shape of the pre- and post-synaptic spikes. The red lines are exponential fits to the data. (c) Training evolution of pattern “F” in the 6 × 6 HZO-based FTJs array. Random false training pulses were involved during the 2nd and 5th training steps to verify its fault tolerance. (d) Conductance maps of patterns “F,” “D,” and “U” after training for 10, 25, and 38 cycles, respectively, and the corresponding statistical distribution of synapse conductance. (a) Reproduced with permission from Boyn et al., Nat. Commun. 8, 14736 (2017). Copyright 2017 Author(s), licensed under a Creative Commons Attribution (CC BY) license. [(b)–(d)] Adapted with permission from Chen et al., Nanoscale 10, 15826 (2018). Copyright 2018 Royal Society of Chemistry.

FIG. 23.

(a) Measurements of STDP in a Co/BiFeO3/(Ca,Ce)MnO3 ferroelectric memristor. Modulation of the device conductance (ΔG) as a function of the delay (Δt) between pre- and post-synaptic spikes. Seven datasets were collected on the same device showing the reproducibility of the effect. The total length of each pre- and post-synaptic spike is 600 ns.77 (b) STDP measurements in a HZO-based FTJ. The top panel shows the shape of the pre- and post-synaptic spikes. The red lines are exponential fits to the data. (c) Training evolution of pattern “F” in the 6 × 6 HZO-based FTJs array. Random false training pulses were involved during the 2nd and 5th training steps to verify its fault tolerance. (d) Conductance maps of patterns “F,” “D,” and “U” after training for 10, 25, and 38 cycles, respectively, and the corresponding statistical distribution of synapse conductance. (a) Reproduced with permission from Boyn et al., Nat. Commun. 8, 14736 (2017). Copyright 2017 Author(s), licensed under a Creative Commons Attribution (CC BY) license. [(b)–(d)] Adapted with permission from Chen et al., Nanoscale 10, 15826 (2018). Copyright 2018 Royal Society of Chemistry.

Close modal

Figures 23(a) and 23(b) show examples of the realization of STDP in FTJs based on BiFeO377 and Hf0.5Zr0.5O2 (HZO),78 respectively. In both cases a clear dependence of the junction conductance with Δt is observed. Remarkably, STDP in such FTJs can be harnessed to perform pattern recognition. This is illustrated in Figs. 23(c) and 23(d) for a 6 × 6 FTJ array used to encode the character ‘F’. The array is trained by sending sequences of half bias voltage pulses, i.e., half the voltage needed to switch a cell, so that only cells at the intersection (fully selected cells) between lines and rows are actually switched, corresponding to the pattern to be learned. While initially all FTJs are in a low conductance state, after the first sequence the conductance of the fully selected cells is clearly updated, as visible in the 6 × 6 conductance map displayed in Fig. 23(c). After subsequent training sequences, the conductance contrast between these cells and the others increases, although some errors can occur in the process (see 2nd training). However, after 10 training cycles, the conductance maps accurately reflect the expected pattern: the character ‘F’ has been learned by the array. Figure 23(d) shows the conductance maps of patterns “F,” “D” and “U” after training for 10, 25 and 38 cycles, respectively.

Ferroelectrics can accumulate or deplete charge at the interface with an adjacent material, while polar atomic displacements at such interfaces can alter the atomic distances and orbital overlap in the adjacent material. Both mechanisms may lead to polarization direction dependent modifications of the local density of states. If the adjacent material is a ferromagnet with a spin-split density of states, these modifications need not be the same for the spin up and spin down states and therefore switching the ferroelectric will yield a change of the interfacial spin-polarization Pσ (the normalized difference between the density of states for spin up and spin down states at the Fermi energy). Such a dependence can be harnessed in spintronics devices, whose response usually directly depends on the spin polarization. This is the case for magnetic tunnel junctions in which the tunnel magnetoresistance (TMR) is directly linked to the spin polarization value through Jullière's formula1044TMR=2Pσ1Pσ2/(1Pσ1Pσ2) where Pσ1 and Pσ2 are the spin polarizations of the two ferromagnetic electrodes interfacing the barrier. Hence, changing Pσ by ferroelectric switching should yield a change in the TMR value, through an effect coined tunnel electromagnetoresistance (TEMR).1045 

Focusing mostly on Fe/BaTiO3 interfaces, several groups predicted changes in the amplitude of the magnetic moments carried by Fe and Ti and of the interfacial spin polarization, depending on the ferroelectric polarization direction, as shown in Fig. 24(a).909,910 To probe these effects experimentally, Garcia et al. fabricated Fe/BaTiO3/LSMO nanoscale tunnel junctions [Fig. 25(a)].1045 The junctions showed negative TMR at 4 K, indicating that the Fe/BaTiO3 interface tunnel spin-polarization is negative, consistent with ab initio calculations.909,910,913 Switching the ferroelectric polarization of the BaTiO3 barrier resulted in changes of the negative TMR: the TMR was large (small) when the ferroelectric polarization points toward the Fe (LSMO) layer.1045 The relative variation of the TMR with the ferroelectric polarization corresponds to a TEMR effect of up to 450%.

FIG. 24.

(a) Majority (top) and minority (bottom) spin density of states of Fe at the Fe/BaTiO3 interface for two opposite ferroelectric polarization directions. The gray shaded plot is the density of states away from the interface. (b) Sketch of the electrically induced magnetic reconstruction at the interface La0.5A0.5 MnO3/BaTiO3 (A = Ca, Sr, or Ba). The arrangement of magnetic moments at Mn sites (small arrows) changes from ferromagnetic (FM) to antiferromagnetic type A when the ferroelectric bias (large arrows) of the BaTiO3 is reversed. (a) Adapted with permission from Duan et al., Phys. Rev. Lett. 97, 047201 (2006). Copyright 2006 American Physical Society. (b) Reproduced with permission from Burton and Tsymbal, Phys. Rev. B 80, 174406 (2009). Copyright 2009 American Physical Society.

FIG. 24.

(a) Majority (top) and minority (bottom) spin density of states of Fe at the Fe/BaTiO3 interface for two opposite ferroelectric polarization directions. The gray shaded plot is the density of states away from the interface. (b) Sketch of the electrically induced magnetic reconstruction at the interface La0.5A0.5 MnO3/BaTiO3 (A = Ca, Sr, or Ba). The arrangement of magnetic moments at Mn sites (small arrows) changes from ferromagnetic (FM) to antiferromagnetic type A when the ferroelectric bias (large arrows) of the BaTiO3 is reversed. (a) Adapted with permission from Duan et al., Phys. Rev. Lett. 97, 047201 (2006). Copyright 2006 American Physical Society. (b) Reproduced with permission from Burton and Tsymbal, Phys. Rev. B 80, 174406 (2009). Copyright 2009 American Physical Society.

Close modal
FIG. 25.

(a) TMR curves of a Fe/BaTiO3/LSMO tunnel junction measured at electrical remanence and 4 K after switching the ferroelectric polarization up or down. The large variation of the TMR reflects changes in the interfacial Fe spin polarization due to ferroelectric polarization reversal. (b) TMR at 50 K in the as-grown state of a Co/PZT/LSMO junction (black squares) and after polarization switching (red circles). The ferroelectric polarization states of the barrier as well as the magnetization directions in the electrodes are shown by arrows. (a) Reproduced with permission from Garcia et al., Science 327, 1106 (2010). Copyright 2010 AAAS. (b) Reproduced with permission from Pantel et al., Nat. Mater. 11, 289 (2012). Copyright 2012 Macmillan Publishers Limited.

FIG. 25.

(a) TMR curves of a Fe/BaTiO3/LSMO tunnel junction measured at electrical remanence and 4 K after switching the ferroelectric polarization up or down. The large variation of the TMR reflects changes in the interfacial Fe spin polarization due to ferroelectric polarization reversal. (b) TMR at 50 K in the as-grown state of a Co/PZT/LSMO junction (black squares) and after polarization switching (red circles). The ferroelectric polarization states of the barrier as well as the magnetization directions in the electrodes are shown by arrows. (a) Reproduced with permission from Garcia et al., Science 327, 1106 (2010). Copyright 2010 AAAS. (b) Reproduced with permission from Pantel et al., Nat. Mater. 11, 289 (2012). Copyright 2012 Macmillan Publishers Limited.

Close modal

These results were then extended to Co/BaTiO3/LSMO junctions, with comparable TEMR effects.916 Pantel et al.1046 subsequently showed that it is even possible to reverse the sign of the TMR by switching the ferroelectric polarization in Co/PZT/LSMO junctions [Fig. 25(b)]. In these experiments, the ferroelectric polarization direction in the PZT tunnel barrier controls the sign of the current spin-polarization, providing new pathways for electrically controlled spintronic devices.

Specific interface engineering can also yield interesting and large TEMR effects. Yin et al.1047 introduced an interfacial La0.5Ca0.5MnO3 layer in LSMO/BaTiO3/LSMO junctions and observed large modifications of the TMR. As illustrated in Fig. 24(b), the polarization-induced metal/insulator phase transition in La0.5Ca0.5MnO3 is accompanied by a transition between ferromagnetic and antiferromagnetic states. At 80 K, the TMR is about 100% when the ferroelectric polarization points toward La0.5Ca0.5MnO3, corresponding to its ferromagnetic state, but is close to zero when pointing in the opposite direction, corresponding to an antiferromagnetic state. Driving an interfacial magnetic phase transition by switching the ferroelectric polarization of the tunnel barrier thus appears to be an efficient way to control the spin-polarization of the tunnel current (with a TEMR reaching 500% in these devices).

Recently, new ideas have arisen that take advantage of ferroelectrically controlled interfacial effects for spintronics in the context of the renewed interest in Rashba spin–orbit coupling (RSOC).1048,1049 RSOC typically occurs at surfaces and interfaces where spatial symmetry breaking results in a built-in electric potential along the out-of-plane direction z. The corresponding Rashba Hamiltonian can be expressed as HR=αR(k×σ)·z, where σ is the vector of the Pauli spin matrices, k the momentum and αR the Rashba coefficient, which is proportional to the electric field strength and the spin–orbit interaction. In a Rashba system the momentum and spin degrees of freedom are locked, as illustrated in Figs. 26(a)–26(c) and the spin degeneracy of the 2D band structure is lifted.

FIG. 26.

(a) The momentum-dependent spin split sub-bands in a Rashba system. (b) Energy cut showing the two Fermi contours with chiral and opposite spin textures. (c) Rashba split energy bands. (d) Direct Edelstein effect: a charge current jc injected along –x induces a shift Δk of both Fermi contours, resulting in a spin accumulation δsδs polarized along the y axis due to the inequivalence of the two contours. (e) Inverse Edelstein effect: when a spin current is injected (e.g., by spin pumping) with the spin polarization along the y axis, the spin population is altered in the steady state causing a displacement in momentum space of the two inequivalent Fermi surfaces (red and blue lines) by ±Δk. This results in a net charge current generated perpendicular to the spin current and to its spin polarization.

FIG. 26.

(a) The momentum-dependent spin split sub-bands in a Rashba system. (b) Energy cut showing the two Fermi contours with chiral and opposite spin textures. (c) Rashba split energy bands. (d) Direct Edelstein effect: a charge current jc injected along –x induces a shift Δk of both Fermi contours, resulting in a spin accumulation δsδs polarized along the y axis due to the inequivalence of the two contours. (e) Inverse Edelstein effect: when a spin current is injected (e.g., by spin pumping) with the spin polarization along the y axis, the spin population is altered in the steady state causing a displacement in momentum space of the two inequivalent Fermi surfaces (red and blue lines) by ±Δk. This results in a net charge current generated perpendicular to the spin current and to its spin polarization.

Close modal

RSOC is appealing for spintronics because it allows one to interconvert spin and charge currents and represents a promising pathway toward ferromagnet-free spintronics (by using RSOC systems as sources and detectors of spin currents, tasks conventionally performed with ferromagnets).1050 When a charge current flows in a Rashba system, it can generate a transverse spin accumulation; this effect is called the Edelstein (or Rashba–Edelstein) effect,1051 as illustrated in Fig. 26(d). This spin accumulation can diffuse into an adjacent conducting material through the interface and generate a pure 3D spin current without net charge current (that is, a pure flow of spin angular momentum, with opposite flows for electrons of opposite spins). Conversely, the injection of a spin current in a Rashba system generates a net charge current (inverse Edelstein effect, IEE),1052 see Fig. 26(e). The efficiency of the IEE can be estimated by the figure of merit λIEE given by the ratio between the 2D charge current generated, jc2D, and the 3D spin current injected, js3D. Note that since one current is 2D and the other is 3D, λIEE has the dimension of length. In the simplest case of circular Fermi contours, λIEE is proportional to the Rashba coefficient αR and the relaxation time τ, so that λIEE=(αRτ)/. One way to optimize the spin-to-charge conversion is therefore to find systems with substantial Rashba SOC and long relaxation times. Interestingly, bulk materials with broken inversion symmetry may also exhibit a Rashba effect, three-dimensional in nature (3D Rashba), as was first identified in BiTeI.1053 The polar nature of ferroelectrics makes them prime candidates to harbor bulk RSOC, with the bonus that it could also be switchable by an electric field. The last few years have seen efforts toward the identification of single-phase Rashba ferroelectrics,1054 mostly focusing on GeTe, a low bandgap ferroelectric semiconductor with a TC of 700 K. GeTe has been predicted to be a bulk Rashba material,1055 where the ferroelectric polarization switching causes a full reversal of the spin texture of the Rashba split Fermi contours. Until recently, because of high leakage, indications of ferroelectricity have only been provided in thin films using piezoresponse force microscopy (PFM),1056 but recent results show a strong electroresistance effect reflecting polarization switching.1057 The surface band structure of GeTe has been mapped by angle resolved photoemission spectroscopy (ARPES), revealing a strong Rashba splitting that depends on the ferroelectric polarization state (in two different samples,1058 or on the same in situ poled sample1059). Very recent results have demonstrated ferroelectric control of spin-charge conversion with GeTe at room temperature.1057 In parallel, a switchable Rashba SOC has been predicted in more classical compounds from the perovskite family including BiAlO3,1060 strained KTaO3,1061 strained SrBiO3,1062 as well as in other oxide compounds such as Aurivilius phases.1063 

Charge accumulation/depletion and polarization-dependent atomic reconstruction at ferroelectric interfaces also offer opportunities to control interfacial RSOC by electric fields in a nonvolatile way [cf. Figs. 27(a)–27(c)]. The first prediction of ferroelectric control of Rashba SOC was by Mirhosseini et al.1065 in Bi/BaTiO3. The dependence of the Rashba coefficient was modest, but Zhong et al.1066 later calculated a large (0.1–0.7 eV Å) and fully switchable Rashba coefficient in various types of perovskite interfaces combining BaTiO3 with BaRuO3, BaIrO3 or BaOsO3. Recently, Song et al.1067 predicted a large effect as well in PbTiO3/BiInO3 interfaces, see Fig. 28.

FIG. 27.

(a) Sketch of a ferroelectric Rashba architecture combining a ferroelectric material (green) and a material with spin–orbit coupling (purple). Upon switching the polarization, electrons accumulate (left) or deplete (right) in the SOC material, creating an electric field whose sign depends on the polarization direction. (b) Corresponding Rashba-split chiral Fermi contours with spin-momentum locking. The chirality of the contours switches upon switching ferroelectric polarization. (c) Inverse Edelstein effect in a Rashba interface. When a spin current is injected (e.g., by spin pumping) with a spin polarization along the y axis, the spin population is altered, causing a displacement in momentum space of the two inequivalent Fermi surfaces (red and blue lines) by ±Δk. This results in a net charge current, generated perpendicular to the spin current and to its spin polarization. The sign of the generated current depends on the chirality of the Fermi contours and is, therefore, reversed upon switching the ferroelectric polarization. (d) Nonvolatile device operated by ferroelectricity and Rashba SOC. A charge current JC is generated by the conversion of a spin current JS through the inverse Edelstein effect. The sign of JC changes with the direction of the ferroelectric polarization. Noël et al., Nature 580, 483 (2020).1064 Copyright 2020 Authors, under exclusive license to Springer Nature Limited.

FIG. 27.

(a) Sketch of a ferroelectric Rashba architecture combining a ferroelectric material (green) and a material with spin–orbit coupling (purple). Upon switching the polarization, electrons accumulate (left) or deplete (right) in the SOC material, creating an electric field whose sign depends on the polarization direction. (b) Corresponding Rashba-split chiral Fermi contours with spin-momentum locking. The chirality of the contours switches upon switching ferroelectric polarization. (c) Inverse Edelstein effect in a Rashba interface. When a spin current is injected (e.g., by spin pumping) with a spin polarization along the y axis, the spin population is altered, causing a displacement in momentum space of the two inequivalent Fermi surfaces (red and blue lines) by ±Δk. This results in a net charge current, generated perpendicular to the spin current and to its spin polarization. The sign of the generated current depends on the chirality of the Fermi contours and is, therefore, reversed upon switching the ferroelectric polarization. (d) Nonvolatile device operated by ferroelectricity and Rashba SOC. A charge current JC is generated by the conversion of a spin current JS through the inverse Edelstein effect. The sign of JC changes with the direction of the ferroelectric polarization. Noël et al., Nature 580, 483 (2020).1064 Copyright 2020 Authors, under exclusive license to Springer Nature Limited.

Close modal
FIG. 28.

Band structure of symmetric PbTiO3/BiInO3 heterostructure with ferroelectric polarization along (a) [001] and (b) [001¯]. The blue circles represent Bi p orbitals. Rashba bands around the Γ point below the Fermi level are magnified in (c) and (d). Reproduced with permission from Song et al., arXiv:1909.03727 (2019), Copyright 2019 Authors, licensed under a Creative Commons Attribution (CC BY) license.

FIG. 28.

Band structure of symmetric PbTiO3/BiInO3 heterostructure with ferroelectric polarization along (a) [001] and (b) [001¯]. The blue circles represent Bi p orbitals. Rashba bands around the Γ point below the Fermi level are magnified in (c) and (d). Reproduced with permission from Song et al., arXiv:1909.03727 (2019), Copyright 2019 Authors, licensed under a Creative Commons Attribution (CC BY) license.

Close modal

Two-dimensional electron gases (2DEGs) based on SrTiO3701 are known to display a sizeable RSOC670 that has been harnessed to achieve very efficient spin-charge1068,1069 and charge-spin conversion.1070,1071 Taking advantage of the possibility to induce a ferroelectric-like state in SrTiO3 by applying large electric fields,1072 Noöl et al. have demonstrated a nonvolatile electrical control of giant spin-charge conversion in SrTiO3 2DEGs.1064 

Achieving ferroelectric control of RSOC offers many types of device opportunities for spintronics, one being illustrated in Fig. 27(d). In this device, a ferromagnet with a fixed magnetization direction is used to inject a spin current into a Rashba state at the interface between a ferroelectric and a material containing a heavy element (e.g., the PbTiO3/BiInO3 interface mentioned above).1067 The spin current injection proceeds by tunneling across the ferroelectric (as in the FTJs described in Sec. IV), which must be sufficiently thin. Depending on the ferroelectric polarization direction, the chirality of the spin contours in the Rashba state will switch, resulting in a change in sign of the effective Rashba coefficient and a reversal in sign of the produced charge current. This illustrates how ferroelectricity could control the inverse Edelstein effect. Reciprocally, the direct Edelstein effect could be controlled ferroelectrically in a similar way, leading to the generation of spin currents with up or down-spin polarization from a given charge current of fixed sign and used to manipulate spin textures in a neighboring layer.

In this overview of the recent developments of ferroelectric interfacial phenomena and devices, we covered both the basic aspects of thin film ferroelectricity and the several device structures where the ferroelectric polarization has been employed to control the interfacial charge carrier density to modulate the correlated state of matter or the charge tunneling transport. We highlighted the recent progress achieved in the growth of high quality, single-crystalline ferroelectric thin films and in the fundamental understanding of the factors defining and limiting the ferroelectric state in the ultrathin limit. In particular, we emphasized the importance of the elastic and electric boundary conditions in determining the ferroelectric properties of thin films, exemplified for the canonical ferroelectric perovskite systems BaTiO3, PbTiO3, Pb(Zr0.2Ti0.8)O3, BiFeO3, and a more recent addition to the ferroelectric family, HfO2, with its proven compatibility with Si CMOS, high bandgap, and high dielectric breakdown voltage, making it a promising materials system for device applications. While we have described only a subset of the large family of ferroelectric materials, our choice of systems illustrate the various phenomena that typically occur at the ultrathin film limit, including the role of epitaxial strain in modifying the ferroelectric properties (and in some cases of stabilizing supertetragonal phases with very high ferroelectric polarizations, which are of high interest for device structures), and of the electrical boundary conditions, since the relatively small bandgap of most ferroelectrics makes consideration of band alignment and Schottky barriers critical for device function. Although we considered here only the basic aspects associated with the static band alignment, it is clear that a working device must fulfill the conditions that allow for switching the ferroelectric polarization, including high endurance and stability, which depend in large measure on the presence and evolution of defects present in the ferroelectric film and largely formed during the growth process, which we also discussed in detail. Indeed, although much progress has been achieved, certain aspects still require improvement; for example, imprint effects, where built-in electric fields favor one direction of the ferroelectric polarization, lead to time instability of the ferroelectric state and depolarization; in a worst case scenario, the opposite polarization state may not be reached at remanence, resulting in reduced field effects or to loss of information.507 

Importantly, in the context of ferroelectric interfacial devices, ferroelectricity has now been demonstrated in ferroelectric films a few unit cells in thickness, settling a long controversy over the intrinsic or extrinsic nature of the so-called ferroelectric dead layer observed in early reports of ferroelectricity in thin films and paving the way for the realization of ferroelectric tunneling devices where very large changes in the conduction state are achieved as a function of the direction of the ferroelectric polarization, and of low voltage switching, high polarization ferroelectric field effect device structures.

In discussing ferroelectric interfacial phenomena, we have largely focused on controlling the correlated state in complex oxide materials using oxide ferroelectric systems. One advantage of such a materials' combination is the compatibility as far as the oxide growth process is concerned: the stability of oxides to ambient conditions and the wide range of physical properties that have been identified in such material systems. Specifically, we have reviewed the control of superconductivity and of the insulator to metal transition, where field effect control of the conducting state aimed at achieving devices with high on/off resistance ratios and lower switching energies as an alternative to CMOS switches. Although in some instances large changes in the resistivity ratio have been obtained, a full control over the order parameter (wave function amplitude in the case of superconductivity, translation symmetry for the metal insulator transition) has proved challenging thus far. Electric field control of orbital and charge order has also been achieved in some instances, although the difficulty in these cases lies in probing the changes in the ordered state, since they require probes that are sensitive to the orbital or valence state. In comparison, we argue that more rapid progress has been achieved in controlling the interfacial magnetic state with electric fields, given the many probes available for measuring magnetic moments and the clearer physical processes linking spin order to charge (even if the details of such processes may be complex in nature). The most clear example is the case of the manganite perovskites, where the dependence of the magnetic properties with charge are known from the bulk phase diagrams, but where the presence of epitaxial strain in thin films and bonding effects at the interface result in more complex and richer magnetic and electronic behavior. Unveiling the details of the coupling mechanism between charge and spin order and a better understanding of how that process evolves at the different magnetic phase boundaries across the large doped manganite family still remains to be carried out.

Also, the rapid development of ferroelectric tunnel junctions has been impressive, from the theoretical and experimental demonstration to applications in device structures, which we briefly reviewed here. Given the challenge of demonstrating ferroelectricity in such ultrathin ferroelectric tunnel barriers, this endeavor has required a painstaking process of ensuring that the required properties are present at the various levels of the device structure. For example, given the large tunneling current across such structures, observation of the displacement current associated with the ferroelectric switching is not possible, while the large electric field applied to such ferroelectric barriers possibly suggests that a large fraction of the applied electric field drops at interfacial dielectric layers within the structure.

Clearly, the approaches reviewed here for oxide materials can, in principle, be extended for other family of correlated systems and ferroelectric materials. Critical are good quality interfaces, good ferroelectric properties down to the ultrathin limit, and correlated systems whose properties show a large sensitivity to the charge density. The current progress in this area shows that controlling the correlated state of matter at ferroelectric interfaces provides not only a means to better understand at a fundamental level the role of electron interactions in determining the electronic properties of materials but also as a means of achieving new functionalities that may be harnessed for electronic devices capable of complex functions, such as low power logic switches, complex logic operations, nanometer scale logic memories, and neuron-like memory elements.

We are very thankful to the authors that kindly allowed reproduction of their work here. C.A.F.V. acknowledges support from the Swiss National Science Foundation under Grant No. 200021_184684. C.H.A., Y.S., and F.J.W. acknowledge support from ONR under Grant No. N000144-19-1-2104. M.B. acknowledges support from the ERC Advanced under Grant No. 833973 “FRESCO.” K.M.R. acknowledges support from ONR under Grant No. N000144-19-1-2073.

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
P.
Zubko
,
S.
Gariglio
,
M.
Gabay
,
P.
Ghosez
, and
J.-M.
Triscone
,
Ann. Rev. Condens. Matter Phys.
2
,
141
(
2011
).
2.
C. A. F.
Vaz
,
F. J.
Walker
,
C. H.
Ahn
, and
S. S.
Ismail-Beigi
,
J. Phys.: Condens. Matter
27
,
123001
(
2015
).
3.
C. H.
Ahn
,
J.-M.
Triscone
, and
J.
Mannhart
,
Nature
424
,
1015
(
2003
).
4.
C. H.
Ahn
,
A.
Bhattacharya
,
M. D.
Ventra
,
J. N.
Eckstein
,
C. D.
Frisbie
,
M. E.
Gershenson
,
A. M.
Goldman
,
I. H.
Inoue
,
J.
Mannhart
,
A. J.
Millis
,
A. F.
Morpurgo
,
D.
Natelson
, and
J.-M.
Triscone
,
Rev. Mod. Phys.
78
,
1185
(
2006
).
5.
A. M.
Goldman
,
Annu. Rev. Mater. Res.
44
,
45
(
2014
).
6.
E. Y.
Tsymbal
and
H.
Kohlstedt
,
Science
313
,
181
(
2006
).
7.
E. Y.
Tsymbal
,
A.
Gruverman
,
V.
Garcia
,
M.
Bibes
, and
A.
Barthélémy
,
MRS Bull.
37
,
138
(
2012
).
8.
M.
Dawber
,
K. M.
Rabe
, and
J. F.
Scott
,
Rev. Mod. Phys.
77
,
1083
(
2005
).
9.
P.
Ghosez
and
J.
Junquera
, “
First-principles modeling of ferroelectric oxide nanostructures
,”
Handbook of Theoretical and Computational Nanotechnology
, edited by
M.
Rieth
and
W.
Schommers
(
American Scientific Publisher
,
Stevenson Ranch, CA
,
2006
), p.
623
.
10.
N.
Setter
,
D.
Damjanovic
,
L.
Eng
,
G.
Fox
,
S.
Gevorgian
,
S.
Hong
,
A.
Kingon
,
H.
Kohlstedt
,
N. Y.
Park
,
G. B.
Stephenson
,
I.
Stolitchnov
,
A. K.
Taganstev
,
D. V.
Taylor
,
T.
Yamada
, and
S.
Streiffer
,
J. Appl. Phys.
100
,
051606
(
2006
).
11.
I. A.
Kornev
,
H.
Fu
, and
L.
Bellaiche
,
J. Mater. Sci.
41
,
137
(
2006
).
12.
D. G.
Schlom
,
L.-Q.
Chen
,
C.-B.
Eom
,
K. M.
Rabe
,
S. K.
Streiffer
, and
J.-M.
Triscone
,
Annu. Rev. Mater. Res.
37
,
589
(
2007
).
13.
Y.
Wang
,
W.
Chen
,
B.
Wang
, and
Y.
Zheng
,
Material
7
,
6377
(
2014
).
14.
L. W.
Martin
and
A. M.
Rappe
,
Nat. Rev. Mater.
2
,
16087
(
2017
).
15.
E.
Bousquet
,
M.
Dawber
,
N.
Stucki
,
C.
Lichtensteiger
,
P.
Hermet
,
S.
Gariglio
,
J.-M.
Triscone
, and
P.
Ghosez
,
Nature
452
,
732
(
2008
).
16.
M.
Dawber
,
N.
Stucki
,
C.
Lichtensteiger
,
S.
Gariglio
, and
J.-M.
Triscone
,
J. Phys.: Condens. Matter
20
,
264015
(
2008
).
17.
M.
Finazzi
,
L.
Duò
, and
F.
Ciccacci
,
Surf. Sci. Rep.
64
,
139
(
2009
).
18.
J.
Mannhart
and
D. G.
Schlom
,
Science
327
,
1607
(
2010
).
19.
L. F.
Kourkoutis
,
H. L.
Xin
,
T.
Higuchi
,
Y.
Hotta
,
J. H.
Lee
,
Y.
Hikita
,
D. G.
Schlom
,
H. Y.
Hwang
, and
D. A.
Muller
,
Philos. Mag.
90
,
4731
(
2010
).
20.
M.
Bibes
,
J. E.
Villegas
, and
A.
Barthélémy
,
Adv. Phys.
60
,
5
(
2011
).
21.
M.
Opel
,
S.
Geprägs
,
E. P.
Menzel
,
A.
Nielsen
,
D.
Reisinger
,
K.-W.
Nielsen
,
A.
Brandlmaier
,
F. D.
Czeschka
,
M.
Althammer
,
M.
Weiler
,
S. T. B.
Goennenwein
,
J.
Simon
,
M.
Svete
,
W.
Yu
,
S.-M.
Hühne
,
W.
Mader
, and
R.
Gross
,
Phys. Status Solidi A
208
,
232
(
2011
).
22.
J. M.
Rondinelli
and
N. A.
Spaldin
,
Adv. Mater.
23
,
3363
(
2011
).
23.
H. Y.
Hwang
,
Y.
Iwasa
,
M.
Kawasaki
,
B.
Keimer
,
N.
Nagaosa
, and
Y.
Tokura
,
Nat. Mater.
11
,
103
(
2012
).
24.
J.
Chakhalian
,
A. J.
Millis
, and
J.
Rondinelli
,
Nat. Mater.
11
,
92
(
2012
).
25.
M.
Fiebig
,
Philos. Trans. R. Soc. A
370
,
4972
(
2012
).
26.
J. M.
Rondinelli
,
S. J.
May
, and
J. W.
Freeland
,
MRS Bull.
37
,
261
(
2012
).
27.
J.
Chakhalian
,
J. W.
Freeland
,
A. J.
Millis
,
C.
Panagopoulos
, and
J. M.
Rondinelli
,
Rev. Mod. Phys.
86
,
1189
(
2014
).
28.
H.
Guo
,
M.
Saghayezhian
,
Z.
Wang
,
Y.
Zhu
,
J.
Zhang
, and
W.
Plummer
,
Front. Phys.
15
,
13401
(
2020
).
29.
M. E.
Lines
and
A. M.
Glass
,
Principles and Applications of Ferroelectrics and Related Materials
(
Clarendon Press
,
Oxford
,
1977
).
30.
K. M.
Rabe
,
M.
Dawber
,
C.
Lichtensteiger
,
C. H.
Ahn
, and
J.-M.
Triscone
, “
Modern physics of ferroelectrics: Essential background
,” in
Physics of Ferroelectrics
, Volume 105 of Topics in Applied Physics, edited by
K.
Rabe
,
C. H.
Ahn
, and
J.-M.
Triscone
(
Springer-Verlag
,
Berlin
,
2007
), p.
1
.
31.
R. E.
Cohen
,
J. Phys. Chem. Solids
61
,
139
(
2000
).
32.
N. A.
Hill
,
J. Phys. Chem. B
104
,
6694
(
2000
).
33.
D. I.
Khomskii
,
J. Magn. Magn. Mater.
306
,
1
(
2006
).
34.
K. M.
Rabe
and
P.
Ghosez
, “
First-principles studies of ferroelectric oxides
,” in
Physics of Ferroelectrics
, Volume 105 of Topics in Applied Physics, edited by
K.
Rabe
,
C. H.
Ahn
, and
J.-M.
Triscone
(
Springer-Verlag
,
Berlin
,
2007
), p.
117
.
35.
N. A.
Spaldin
, “
Analogies and differences between ferroelectrics and ferromagnets
,” in
Physics of Ferroelectrics
, Volume 105 of Topics in Applied Physics, edited by
K.
Rabe
,
C. H.
Ahn
, and
J.-M.
Triscone
(
Springer-Verlag
,
Berlin
,
2007
), p.
175
.
36.
37.
S.-W.
Cheong
and
M.
Mostovoy
,
Nat. Mater.
6
,
13
(
2007
).
38.
C. N. R.
Rao
and
C. R.
Serrao
,
J. Mater. Chem.
17
,
4931
(
2007
).
39.
J.
van den Brink
and
D. I.
Khomskii
,
J. Phys.: Condens. Matter
20
,
434217
(
2008
).
40.
41.
C. M.
Teodorescu
,
Phys. Chem. Chem. Phys.
23
,
4085
(
2021
).
42.
R. D.
King-Smith
and
D.
Vanderbilt
,
Phys. Rev. B
47
,
1651
(
1993
).
43.
D.
Vanderbilt
and
R. D.
King-Smith
,
Phys. Rev. B
48
,
4442
(
1993
).
44.
45.
46.
R.
Resta
and
D.
Vanderbilt
, “
Theory of polarisation: A modern approach
,” in
Physics of Ferroelectrics
, Volume 105 of Topics in Applied Physics, edited by
K.
Rabe
,
C. H.
Ahn
, and
J.-M.
Triscone
(
Springer-Verlag
,
Berlin
,
2007
), p.
31
.
47.
N. A.
Spaldin
,
J. Solid State Chem.
195
,
2
(
2012
).
48.
M. V.
Berry
,
Proc. R. Soc. London, A
392
,
45
(
1984
).
49.
D. J.
Kim
,
J. Y.
Jo
,
Y. S.
Kim
,
Y. J.
Chang
,
J. S.
Lee
,
J.-G.
Yoon
,
T. K.
Song
, and
T. W.
Noh
,
Phys. Rev. Lett.
95
,
237602
(
2005
).
50.
M.
Stengel
and
N. A.
Spaldin
,
Nature
443
,
679
(
2006
).
51.
L.-W.
Chang
,
M.
Alexe
,
J. F.
Scott
, and
J. M.
Gregg
,
Adv. Mater.
21
,
4911
(
2009
).
52.
M.
Stengel
,
D.
Vanderbilt
, and
N. A.
Spaldin
,
Nat. Mater.
8
,
392
(
2009
).
53.
Y.
Wang
,
M. K.
Niranjan
,
K.
Janicka
,
J. P.
Velev
,
M. Y.
Zhuravlev
,
S. S.
Jaswal
, and
E. Y.
Tsymbal1
,
Phys. Rev. B
82
,
094114
(
2010
).
54.
P.
Sun
,
Y.-Z.
Wu
,
T.-Y.
Cai
, and
S.
Ju
,
Appl. Phys. Lett.
99
,
052901
(
2011
).
55.
Q.
Yang
,
J.
Cao
,
Y.
Zhou
,
L.
Sun
, and
X.
Lou
,
Acta Mater.
112
,
216
(
2016
).
56.
C. M.
Teodorescu
,
L.
Pintilie
,
N. G.
Apostol
,
R. M.
Costescu
,
G. A.
Lungu
,
L.
Hrib
,
L.
Trupină
,
L. C.
Tănase
,
I. C.
Bucur
, and
A. E.
Bocîrnea
,
Phys. Rev. B
96
,
115438
(
2017
).
57.
G. A.
Boni
,
C. F.
Chirila
,
L.
Hrib
,
R.
Negrea
,
L. D.
Filip
,
I.
Pintilie
, and
L.
Pintilie
,
Sci. Rep.
9
,
14698
(
2019
).
58.
W. L.
Brown
, U.S. patent 2,791,759 (7 May
1957
).
59.
D. H.
Looney
, U.S. patent 2,791,758 (7 May
1957
).
61.
J. R.
Contreras
,
H.
Kohlstedt
,
U.
Poppe
,
R.
Waser
,
C.
Buchal
, and
N. A.
Pertsev
,
Appl. Phys. Lett.
83
,
4595
(
2003
).
62.
M. Ye.
Zhuravlev
,
S. S.
Jaswal
,
E. Y.
Tsymbal
, and
R. F.
Sabirianov
,
Appl. Phys. Lett.
87
,
222114
(
2005
).
63.
H.
Kohlstedt
,
N. A.
Pertsev
,
J. R.
Contreras
, and
R.
Waser
,
Phys. Rev. B
72
,
125341
(
2005
).
64.
J. P.
Velev
,
C.-G.
Duan
,
K. D.
Belashchenko
,
S. S.
Jaswal
, and
E. Y.
Tsymbal
,
Phys. Rev. Lett.
98
,
137201
(
2007
).
65.
P.
Maksymovych
,
S.
Jesse
,
P.
Yu
,
R.
Ramesh
,
A. P.
Baddorf
, and
S. V.
Kalinin
,
Science
324
,
1421
(
2009
).
66.
M. Ye.
Zhuravlev
,
Y.
Wang
,
S.
Maekawa
, and
E. Y.
Tsymbal
,
Appl. Phys. Lett.
95
,
052902
(
2009
).
67.
V.
Garcia
,
S.
Fusil
,
K.
Bouzehouane
,
S.
Enouz-Vedrenne
,
N. D.
Mathur
,
A.
Barthélémy
, and
M.
Bibes
,
Nature
460
,
81
(
2009
).
68.
A.
Gruverman
,
D.
Wu
,
H.
Lu
,
Y.
Wang
,
H. W.
Jang
,
C. M.
Folkman
,
M. Ye.
Zhuravlev
,
D.
Felker
,
M.
Rzchowski
,
C.-B.
Eom
, and
E. Y.
Tsymbal
,
Nano Lett.
9
,
3539
(
2009
).
69.
R.
Meservey
and
P. M.
Tedrow
,
Phys. Rep.
238
,
173
(
1994
).
70.
W. J.
Gallagher
and
S. S. P.
Parkin
,
IBM J. Res. Dev.
50
,
5
(
2006
).
71.
M.
Lorenz
,
M. S. R.
Rao
,
T.
Venkatesan
,
E.
Fortunato
,
P.
Barquinha
,
R.
Branquinho
,
D.
Salgueiro
,
R.
Martins
,
E.
Carlos
,
A.
Liu
,
F. K.
Shan
,
M.
Grundmann
,
H.
Boschker
,
J.
Mukherjee
,
M.
Priyadarshini
,
N.
DasGupta
,
D. J.
Rogers
,
F. H.
Teherani
,
E. V.
Sandana
,
P.
Bove
,
K.
Rietwyk
,
A.
Zaban
,
A.
Veziridis
,
A.
Weidenkaff
,
M.
Muralidhar
,
M.
Murakami
,
S.
Abel
,
J.
Fompeyrine
,
J.
Zuniga-Perez
,
R.
Ramesh
,
N. A.
Spaldin
,
S.
Ostanin
,
V.
Borisov
,
I.
Mertig
,
V.
Lazenka
,
G.
Srinivasan
,
W.
Prellier
,
M.
Uchida
,
M.
Kawasaki
,
R.
Pentcheva
,
P.
Gegenwart
,
F. M.
Granozio
,
J.
Fontcuberta
, and
N.
Pryds
,
J. Phys. D: Appl. Phys.
49
,
433001
(
2016
).
72.
A.
Kalaboukhov
and
H.
Boschker
, “
Nanoscale patterning of complex-oxide materials
,”
Appl. Surf. Sci.
482
,
17
(
2019
).
73.
G.
Cheng
and
J.
Levy
, “
Possible future quantum technologies based on correlated nanoelectronics
,”
Appl. Surf. Sci.
482
,
44
(
2019
).
74.
J.
Grollier
, “
Oxide nano-electronics for neuromorphic computing
,”
Appl. Surf. Sci.
482
,
42
(
2019
).
75.
R.
Dittmann
, “
Resistive switching oxides for data storage
,”
Appl. Surf. Sci.
482
,
30
(
2019
).
76.
S.
Saïghi
,
C. G.
Mayr
,
T.
Serrano-Gotarredona
,
H.
Schmidt
,
G.
Lecerf
,
J.
Tomas
,
J.
Grollier
,
S.
Boyn
,
A. F.
Vincent
,
D.
Querlioz
,
S.
La Barbera
,
F.
Alibart
,
D.
Vuillaume
,
O.
Bichler
,
C.
Gamrat
, and
B.
Linares-Barranco
,
Front. Neurosci.
9
,
51
(
2015
).
77.
S.
Boyn
,
J.
Grollier
,
G.
Lecerf
,
B.
Xu
,
N.
Locatelli
,
S.
Fusil
,
S.
Girod
,
C.
Carrétéro
,
K.
Garcia
,
S.
Xavier
,
J.
Tomas
,
L.
Bellaiche
,
M.
Bibes
,
A.
Barthélémy
,
S.
Saïghi
, and
V.
Garcia
,
Nat. Commun.
8
,
14736
(
2017
).
78.
L.
Chen
,
T.-Y.
Wang
,
Y.-W.
Dai
,
M.-Y.
Cha
,
H.
Zhu
,
Q.-Q.
Sun
,
S.-J.
Ding
,
P.
Zhou
,
L.
Chua
, and
D. W.
Zhang
,
Nanoscale
10
,
15826
(
2018
).
79.
K.
Galatsis
,
C.
Ahn
,
I.
Krivorotov
,
P.
Kim
,
R.
Lake
,
K. L.
Wang
, and
J. P.
Chang
,
IEEE J. Exploratory Solid-State Comput. Devices Circuits
1
,
19
(
2015
).
80.
A. H. C.
Neto
,
F.
Guinea
,
N. M. R.
Peres
,
K. S.
Novoselov
, and
A. K.
Geim
,
Rev. Mod. Phys.
81
,
109
(
2009
).
81.
V.
Meunier
,
A. G.
Souza Filho
,
E. B.
Barros
, and
M. S.
Dresselhaus
,
Rev. Mod. Phys.
88
,
025005
(
2016
).
82.
W.
Zhao
,
R. M.
Ribeiro
, and
G.
Eda
,
Acc. Chem. Res.
48
,
91
(
2015
).
83.
Y.
Li
,
C.
Chen
,
W.
Li
,
X.
Mao
,
H.
Liu
,
J.
Xiang
,
A.
Nie
,
Z.
Liu
,
W.
Zhu
, and
H.
Zeng
,
Adv. Electron. Mater.
6
,
2000061
(
2020
).
84.
R.
Samal
,
G.
Sanyal
,
B.
Chakraborty
, and
C. S.
Rout
,
J. Mater. Chem. A
9
,
2560
(
2021
).
85.
X.
Hong
,
J. Phys.: Condens. Matter
28
,
103003
(
2016
).
86.
Y.
Sun
,
R.
Wang
, and
K.
Liu
,
Appl. Phys. Rev.
4
,
011301
(
2017
).
87.
C.
Zhou
and
Y.
Chai
,
Adv. Electron. Mater.
3
,
1600400
(
2017
).
88.
J.
Wang
and
W.
Hu
,
Chin. Phys. B
26
,
037106
(
2017
).
89.
W.
Zhu
,
T.
Low
,
H.
Wang
,
P.
Ye
, and
X.
Duan
,
2D Mater.
6
,
032004
(
2019
).
90.
K. T.
Kang
,
J.
Park
,
D.
Suh
, and
W. S.
Choi
,
Adv. Mater.
31
,
1803732
(
2019
).
91.
H.
Ryu
,
K.
Xu
,
D.
Li
,
X.
Hong
, and
W.
Zhu
,
Appl. Phys. Lett.
117
,
080503
(
2020
).
92.
J.
Chu
,
Y.
Wang
,
X.
Wang
,
K.
Hu
,
G.
Rao
,
C.
Gong
,
C.
Wu
,
H.
Hong
,
X.
Wang
,
K.
Liu
,
C.
Gao
, and
J.
Xiong
,
Adv. Mater.
33
,
2004469
(
2021
).
93.
A.
Lipatov
,
A.
Fursina
,
T. H.
Vo
,
P.
Sharma
,
A.
Gruverman
, and
A.
Sinitskii
,
Adv. Electron. Mater.
3
,
1700020
(
2017
).
94.
S.
Yan
,
H.
Huang
,
Z.
Xie
,
G.
Ye
,
X.-X.
Li
,
T.
Taniguchi
,
K.
Watanabe
,
Z.
Han
,
X.
Chen
,
J.
Wang
, and
J.-H.
Chen
,
ACS Appl. Mater. Interfaces
11
,
42358
(
2019
).
95.
P.-C.
Shen
,
C.
Lin
,
H.
Wang
,
K. H.
Teo
, and
J.
Kong
,
Appl. Phys. Lett.
116
,
033501
(
2020
).
96.
E. V.
Milov
and
B. A.
Strukov
,
Phys. Solid State
43
,
513
(
2001
).
97.
N.
Ikeda
,
H.
Ohsumi
,
K.
Ohwada
,
K.
Ishii
,
T.
Inami
,
K.
Kakurai
,
Y.
Murakami
,
K.
Yoshii
,
S.
Mori
,
Y.
Horibe
, and
H.
Kitô
,
Nature
436
,
1136
(
2005
).
98.
R.
Takahashi
and
M.
Lippmaa
,
Jpn. J. Appl. Phys.
57
,
Part 1
,
0902A1
(
2018
).
99.
J.
Robertson
and
C. W.
Chen
,
Appl. Phys. Lett.
74
,
1168
(
1999
).
100.
B.
Nagaraj
,
S.
Aggarwal
, and
R.
Ramesh
,
J. Appl. Phys.
90
,
375
(
2001
).
101.
J. J.
Lee
and
S. B.
Desu
,
Ferroelectric Lett.
20
,
27
(
1995
).
102.
S. K.
Dey
,
J.-J.
Lee
, and
P.
Alluri
,
Jpn. J. Appl. Phys.
34
,
Part 1
,
3142
(
1995
).
103.
L.
Pintilie
,
I.
Vrejoiu
,
D.
Hesse
, and
M.
Alexe
,
J. Appl. Phys.
104
,
114101
(
2008
).
104.
I.
Pintilie
,
C. M.
Teodorescu
,
C.
Ghica
,
C.
Chirila
,
A. G.
Boni
,
L.
Hrib
,
I.
Pasuk
,
R.
Negrea
,
N.
Apostol
, and
L.
Pintilie
,
ACS Appl. Mater. Interfaces
6
,
2929
(
2014
).
105.
U.
Rabe
,
M.
Kopycinska
,
S.
Hirsekorn
,
J. M.
Saldaña
,
G. A.
Schneider
, and
W.
Arnold
,
J. Phys. D: Appl. Phys.
35
,
2621
(
2002
).
106.
A.
Gruverman
and
A.
Kholkin
,
Rep. Prog. Phys.
69
,
2443
(
2006
).
107.
S. V.
Kalinin
,
A. N.
Morozovska
,
L. Q.
Chen
, and
B. J.
Rodriguez
,
Rep. Prog. Phys.
73
,
056502
(
2010
).
108.
E.
Soergel
,
J. Phys. D: Appl. Phys.
44
,
464003
(
2011
).
109.
N.
Balke
,
P.
Maksymovych
,
S.
Jesse
,
A.
Herklotz
,
A.
Tselev
,
C.-B.
Eom
,
I. I.
Kravchenko
, and
S. V.
Kalinin
,
ACS Nano
9
,
6484
(
2015
).
110.
A. G.
Agronin
,
Y.
Rosenwaks
, and
G. I.
Rosenman
,
Nano Lett.
3
,
169
(
2003
).
111.
T.
Jungk
,
Á.
Hoffmann
, and
E.
Soergel
,
Appl. Phys. Lett.
89
,
042901
(
2006
).
112.
T.
Jungk
,
Á.
Hoffmann
, and
E.
Soergel
,
Appl. Phys. Lett.
91
,
253511
(
2007
).
113.
T.
Jungk
,
Á.
Hoffmann
, and
E.
Soergel
,
J. Microsc.
227
,
72
(
2007
).
114.
T.
Jungk
,
Á.
Hoffmann
, and
E.
Soergel
,
Appl. Phys. A
86
,
353
(
2007
).
115.
Á.
Hoffmann
,
T.
Jungk
, and
E.
Soergel
,
Rev. Sci. Instrum.
78
,
016101
(
2007
).
116.
A.
Labuda
and
R.
Proksch
,
Appl. Phys. Lett.
106
,
253103
(
2015
).
117.
T.
Tybell
,
P.
Paruch
,
T.
Giamarchi
, and
J.-M.
Triscone
,
Phys. Rev. Lett.
89
,
097601
(
2002
).
118.
P.
Paruch
,
T.
Giamarchi
, and
J.-M.
Triscone
, “
Nanoscale studies of domain walls in epitaxial ferroelectric thin films
,” in
Physics of Ferroelectrics
, Volume 105 of Topics in Applied Physics, edited by
K.
Rabe
,
C. H.
Ahn
, and
J.-M.
Triscone
(
Springer-Verlag
,
Berlin
,
2007
), p.
339
.
119.
P.
Sutter
and
E.
Sutter
,
Adv. Funct. Mater.
23
,
2617
(
2013
).
120.
J. E.
Rault
,
T. O.
Menteş
,
A.
Locatelli
, and
N.
Barrett
,
Sci. Rep.
4
,
6792
(
2014
).
121.
N.
Barrett
,
J. E.
Rault
,
J. L.
Wang
,
C.
Mathieu
,
A.
Locatelli
,
T. O.
Mentes
,
M. A.
Niño
,
S.
Fusil
,
M.
Bibes
,
A.
Barthélémy
,
D.
Sando
,
W.
Ren
,
S.
Prosandeev
,
L.
Bellaiche
,
B.
Vilquin
,
A.
Petraru
,
I. P.
Krug
, and
C. M.
Schneider
,
J. Appl. Phys.
113
,
187217
(
2013
).
122.
C. A. F.
Vaz
,
A.
Kleibert
, and
M.
El Kazzi
, “
Nanoscale xpeem spectromicroscopy
,” in
21st Century Nanoscience—A Handbook
, edited by
K. D.
Sattler
(
Taylor & Francis/CRC Press
,
2020
), Chap.
17
.
123.
S.
Cherifi
,
R.
Hertel
,
S.
Fusil
,
H.
Béa
,
K.
Bouzehouane
,
J.
Allibe
,
M.
Bibes
, and
A.
Barthélémy
,
Phys. Status Solidi RRL
4
,
22
(
2010
).
124.
R.
Moubah
,
M.
Elzo
,
S. E.
Moussaoui
,
D.
Colson
,
N.
Jaouen
,
R.
Belkhou
, and
M.
Viret
,
Appl. Phys. Lett.
100
,
042406
(
2012
).
125.
J. E.
Rault
,
W.
Ren
,
S.
Prosandeev
,
S.
Lisenkov
,
D.
Sando
,
S.
Fusil
,
M.
Bibes
,
A.
Barthélémy
,
L.
Bellaiche
, and
N.
Barrett
,
Phys. Rev. Lett.
109
,
267601
(
2012
).
126.
A.
Sander
,
M.
Christl
,
C.-T.
Chiang
,
M.
Alexe
, and
W.
Widdra
,
J. Appl. Phys.
118
,
224102
(
2015
).
127.
D. D.
Fong
,
G. B.
Stephenson
,
S. K.
Streiffer
,
J. A.
Eastman
,
O.
Auciello
,
P. H.
Fuoss
, and
C.
Thompson
,
Science
304
,
1650
(
2004
).
128.
D. D.
Fong
and
C.
Thompson
,
Annu. Rev. Matter. Res.
36
,
431
(
2006
).
129.
G.
Catalan
,
A. H. G.
Vlooswijk
,
A.
Janssens
,
G.
Rispens
,
S.
Redfern
,
G.
Rijnders
,
D. H. A.
Blank
, and
B.
Noheda
,
Integr. Ferroelectr.
92
,
18
(
2007
).
130.
E. A.
Kraut
,
R. W.
Grant
,
J. R.
Waldrop
, and
S. P.
Eowalczyk
,
Phys. Rev. Lett.
44
,
1620
(
1980
).
131.
A. D.
Katnani
and
G.
Margaritondo
,
Phys. Rev. B
28
,
1944
(
1983
).
132.
J. R.
Waldrop
,
R. W.
Grant
,
S. P.
Kowalczyk
, and
E. A.
Kraut
,
J. Vac. Sci. Technol. A
3
,
835
(
1985
).
133.
J.
Zegenhagen
,
Eur. Phys. J. Appl. Phys.
70
,
20701
(
2015
).
134.
M.-A.
Husanu
and
C. A. F.
Vaz
, “
Spectroscopic characterisation of multiferroic interfaces
,” in
Spectroscopy of Complex Oxide Interfaces
, Volume 266 of Springer Series in Materials Science, edited by
C.
Cancellieri
and
V. N.
Strocov
(
Springer
,
Cham
,
2018
), p.
245
.
135.
R. V.
Vedrinskii
,
V. L.
Kraizman
,
A. A.
Novakovich
,
P. V.
Demekhin
,
S. V.
Urazhdin
,
B.
Ravel
, and
E. A.
Stern
,
J. Phys. IV France
7
,
C2
107
(
1997
).
136.
B.
Ravel
,
E. A.
Stern
,
R. I.
Vedrinskii
, and
V.
Kraizman
,
Ferroelectrics
206
,
407
(
1998
).
137.
F. M. F.
de Groot
,
Chem. Rev.
101
,
1779
(
2001
).
138.
D.
Cao
,
I.-K.
Jeong
,
R. H.
Heffner
,
T.
Darling
,
J.-K.
Lee
,
F.
Bridges
,
J.-S.
Park
, and
K.-S.
Hong
,
Phys. Rev. B
70
,
224102
(
2004
).
139.
A.
Mesquita
,
A.
Michalowicz
, and
V. R.
Mastelaro
,
J. Phys.: Conf. Ser.
190
,
012081
(
2009
).
140.
A.
Al-Zein
,
G.
Fraysse
,
J.
Rouquette
,
P.
Papet
,
J.
Haines
,
B.
Hehlen
,
C.
Levelut
,
G.
Aquilanti
, and
Y.
Joly
,
Phys. Rev. B
81
,
174110
(
2010
).
141.
E.
Arenholz
,
G.
van der Laan
,
A.
Fraile-Rodríguez
,
P.
Yu
,
Q.
He
, and
R.
Ramesh
,
Phys. Rev. B
82
,
140103
14010R
(
2010
).
142.
H. N.
Chapman
,
A.
Barty
,
M. J.
Bogan
,
S.
Boutet
,
M.
Frank
,
S. P.
Hau-Riege
,
S.
Marchesini
,
B. W.
Woods
,
S.
Bajt
,
W. H.
Benner
,
R. A.
London
,
E.
Plönjes
,
M.
Kuhlmann
,
R.
Treusch
,
S.
Düsterer
,
T.
Tschentscher
,
J. R.
Schneider
,
E.
Spiller
,
T.
Möller
,
C.
Bostedt
,
M.
Hoener
,
D. A.
Shapiro
,
K. O.
Hodgson
,
D.
van der Spoel
,
F.
Burmeister
,
M.
Bergh
,
C.
Caleman
,
G.
Huldt
,
M. M.
Seibert
,
F. R. N. C.
Maia
,
R. W.
Lee
,
A.
Szöke
,
N.
Timneanu
, and
J.
Hajdu
,
Nat. Phys.
2
,
839
(
2006
).
143.
T. F.
Nova
,
A. S.
Disa
,
M.
Fechner
, and
A.
Cavalleri
,
Science
364
,
1075
(
2019
).
144.
S. R.
Spurgeon
, “
Scanning transmission electron microscopy of oxide interfaces and heterostructures
,” arXiv:2001.00947 [cond-mat.mtrl-sci] (
2020
).
145.
D.
Su
,
Q.
Meng
,
C. A. F.
Vaz
,
M.-G.
Han
,
Y.
Segal
,
F. J.
Walker
,
M.
Sawicki
,
C.
Broadbridge
, and
C. H.
Ahn
,
Appl. Phys. Lett.
99
,
102902
(
2011
).
146.
P.
Gao
,
C. T.
Nelson
,
J. R.
Jokisaari
,
Y.
Zhang
,
S.-H.
Baek
,
C. W.
Bark
,
E.
Wang
,
Y.
Liu
,
C.-B.
Eom
, and
X.
Pan
,
Adv. Mater.
24
,
1106
(
2012
).
147.
S. R.
Spurgeon
,
P. V.
Balachandran
,
D. M.
Kepaptsoglou
,
A. R.
Damodaran
,
J.
Karthik
,
S.
Nejati
,
L.
Jones
,
H.
Ambaye
,
V.
Lauter
,
Q. M.
Ramasse
,
K. K. S.
Lau
,
L. W.
Martin
,
J. M.
Rondinelli
, and
M. L.
Taheri
,
Nat. Commun.
6
,
6735
(
2015
).
148.
W.
Zhong
,
R. D.
King-Smith
, and
D.
Vanderbilt
,
Phys. Rev. Lett.
72
,
3618
(
1994
).
149.
P.
Chen
,
Z.
Huang
,
C.
Li
,
B.
Zhang
,
N.
Bao
,
P.
Yang
,
X.
Yu
,
S.
Zeng
,
C.
Tang
,
X.
Wu
,
J.
Chen
,
J.
Ding
,
S. J.
Pennycook
,
A.
Ariando
,
T. V.
Venkatesan
, and
G. M.
Chow
,
Adv. Funct. Mater.
28
,
1801766
(
2018
).
150.
C. T.
Nelson
,
P.
Gao
,
J. R.
Jokisaari
,
C.
Heikes
,
C.
Adamo
,
A.
Melville
,
S.-H.
Baek
,
C. M.
Folkman
,
B.
Winchester
,
Y.
Guv
,
Y.
Liu
,
K.
Zhang
,
E.
Wang
,
J.
Li
,
L.-Q.
Chen
,
C.-B.
Eom
,
D. G.
Schlom
, and
X.
Pan
,
Science
334
,
968
(
2011
).
151.
P.
Gao
,
C. T.
Nelson
,
J. R.
Jokisaari
,
S.-H.
Baek
,
C. W.
Bark
,
Y.
Zhang
,
E.
Wang
,
D. G.
Schlom
,
C.-B.
Eom
, and
X.
Pan
,
Nat. Commun.
2
,
591
(
2011
).
152.
P.
Gao
,
J.
Britson
,
C. T.
Nelson
,
J. R.
Jokisaari
,
C.
Duan
,
M.
Trassin
,
S.-H.
Baek
,
H.
Guo
,
L.
Li
,
Y.
Wang
,
Y.-H.
Chu
,
A. M.
Minor
,
C.-B.
Eom
,
R.
Ramesh
,
L.-Q.
Chen
, and
X.
Pan
,
Nat. Commun.
5
,
3801
(
2014
).
153.
H. J.
Chang
,
S. V.
Kalinin
,
A. N.
Morozovska
,
M.
Huijben
,
Y.-H.
Chu
,
P.
Yu
,
R.
Ramesh
,
E. A.
Eliseev
,
G. S.
Svechnikov
,
S. J.
Pennycook
, and
A. Y.
Borisevich
,
Adv. Mater.
23
,
2474
(
2011
).
154.
Y.-M.
Kim
,
A.
Kumar
,
A.
Hatt
,
A. N.
Morozovska
,
A.
Tselev
,
M. D.
Biegalski
,
I.
Ivanov
,
E. A.
Eliseev
,
S. J.
Pennycook
,
J. M.
Rondinelli
,
S. V.
Kalinin
, and
A. Y.
Borisevich
,
Adv. Mater.
25
,
2497
(
2013
).
155.
D.
Zheng
,
J.
Zhang
,
Y.
Wang
,
Z.
Wang
,
W.
Zheng
,
D.
Li
,
L.
Zhang
,
C.
Jin
,
P.
Li
,
X.
Zhang
, and
H.
Bai
,
ACS Appl. Mater. Interfaces
10
,
40328
(
2018
).
156.
S. R.
Spurgeon
,
J. D.
Sloppy
,
D. M.
Kepaptsoglou
,
P. V.
Balachandran
,
S.
Nejati
,
J.
Karthik
,
A. R.
Damodaran
,
C. L.
Johnson
,
H.
Ambaye
,
R.
Goyette
,
V.
Lauter
,
Q. M.
Ramasse
,
J. C.
Idrobo
,
K. K. S.
Lau
,
S. E.
Lofland
, Jr.
,
J. M.
Rondinelli
,
L. W.
Martin
, and
M. L.
Taheri
,
ACS Nano
8
,
894
(
2014
).
157.
R.
Soni
,
A.
Petraru
,
H. S.
Nair
,
O.
Vavra
,
M.
Ziegler
,
S. K.
Kim
,
D. S.
Jeong
, and
H.
Kohlstedt
,
Nanoscale
8
,
10799
(
2016
).
158.
Q.
Meng
,
G.
Xu
,
H.
Xin
,
E. A.
Stach
,
Y.
Zhu
, and
D.
Su
,
J. Phys. Chem. A
123
,
4632
(
2019
).
159.
H.
Guo
,
Z.
Wang
,
S.
Dong
,
S.
Ghosh
,
M.
Saghayezhian
,
L.
Chen
,
Y.
Weng
,
A.
Herklotz
,
T. Z.
Ward
,
R.
Jin
,
S. T.
Pantelides
,
Y.
Zhub
,
J.
Zhang
, and
E. W.
Plummer
,
Proc. Natl. Acad. Sci.
114
,
E5062
(
2017
).
160.
H.
Wang
,
X.
Chi
,
Z.
Liu
,
H.
Yoong
,
L.
Tao
,
J.
Xiao
,
R.
Guo
,
J.
Wang
,
Z.
Dong
,
P.
Yang
,
C.-J.
Sun
,
C.
Li
,
X.
Yan
,
J.
Wang
,
G. M.
Chow
,
E. Y.
Tsymbal
,
H.
Tian
, and
J.
Chen
,
Nano Lett.
19
,
3057
(
2019
).
161.
M.
Marinova
,
J. E.
Rault
,
A.
Gloter
,
S.
Nemsak
,
G. K.
Palsson
,
J.-P.
Rueff
,
C. S.
Fadley
,
C.
Carrétéro
,
H.
Yamada
,
K.
March
,
V.
Garcia
,
S.
Fusil
,
A.
Barthélémy
,
O.
Stéphan
,
C.
Colliex
, and
M.
Bibes
,
Nano Lett.
15
,
2533
(
2015
).
162.
Z.
Yuan
,
J.
Ruan
,
L.
Xie
,
X.
Pan
,
D.
Wu
, and
P.
Wang
,
Appl. Phys. Lett.
110
,
171602
(
2017
).
163.
D. G.
Schlom
,
J. H.
Haeni
,
J.
Lettieri
,
C. D.
Theis
,
W.
Tian
,
J. C.
Jiang
, and
X. Q.
Pan
,
Mater. Sci. Eng. B
87
,
282
(
2001
).
164.
A.-B.
Posadas
,
M.
Lippmaa
,
F. J.
Walker
,
M.
Dawber
,
C. H.
Ahn
, and
J.-M.
Triscone
, “
Growth and novel applications of epitaxial oxide thin films
,” in
Physics of Ferroelectrics
, Volume 105 of Topics in Applied Physics, edited by
K.
Rabe
,
C. H.
Ahn
, and
J.-M.
Triscone
(
Springer-Verlag
,
Berlin
,
2007
), p.
219
.
165.
N.
Izyumskaya
,
Y.-I.
Alivov
,
S.-J.
Cho
,
H.
Morkoç
,
H.
Lee
, and
Y.-S.
Kang
,
Crit. Rev. Solid State Mater. Sci.
32
,
111
(
2007
).
166.
N.
Izyumskaya
,
Y.
Alivov
, and
H.
Morkoç
,
Crit. Rev. Solid State Mater. Sci.
34
,
89
(
2009
).
167.
M.
Coll
,
J.
Fontcuberta
,
M.
Althammer
,
M.
Bibes
,
H.
Boschker
,
A.
Calleja
,
G.
Cheng
,
M.
Cuoco
,
R.
Dittmann
,
B.
Dkhil
,
I. E.
Baggari
,
M.
Fanciulli
,
I.
Fina
,
E.
Fortunato
,
C.
Frontera
,
S.
Fujita
,
V.
Garcia
,
S. T. B.
Goennenwein
,
C.-G.
Granqvist
,
J.
Grollier
,
R.
Gross
,
A.
Hagfeldt
,
G.
Herranz
,
K.
Hono
,
E.
Houwman
,
M.
Huijben
,
A.
Kalaboukhov
,
D. J.
Keeble
,
G.
Koster
,
L. F.
Kourkoutis
,
J.
Levy
,
M.
Lira-Cantu
,
J. L.
MacManus-Driscoll
,
J.
Mannhart
,
R.
Martins
,
S.
Menzel
,
T.
Mikolajick
,
M.
Napari
,
M. D.
Nguyen
,
G.
Niklasson
,
C.
Paillard
,
S.
Panigrahi
,
G.
Rijnders
,
F.
Sánchez
,
P.
Sanchis
,
S.
Sanna
,
D. G.
Schlom
,
U.
Schroeder
,
K. M.
Shen
,
A.
Siemon
,
M.
Spreitzer
,
H.
Sukegawa
,
R.
Tamayo
,
J.
van den Brink
,
N.
Pryds
, and
F. M.
Granozio
,
Appl. Surf. Sci.
482
,
1
(
2019
).
168.
L.
Shi
and
J.
Ni
,
Phys. Rev. Lett.
97
,
126105
(
2006
).
169.
M.
Adachi
,
Y.
Akishige
,
K.
Deguchi
,
T.
Ikeda
,
M.
Okuyama
,
A.
Sakai
,
E.
Sawaguchi
,
T.
Takenaka
,
T.
Tsukamoto
, and
T.
Yagi
, “
Perovskite-type oxides
,” in
Ferroelectrics and Related Substances
, Volume
36A1
of Landolt-Börnstein - Group III: Crystal and Solid State Physics, edited by
Y.
Shiozaki
,
E.
Nakamura
, and
T.
Mitsui
(
Springer-Verlag
,
Berlin
,
2001
), p.
67
.
170.
J.
Robertson
,
Eur. Phys. J. Appl. Phys.
28
,
265
(
2004
).
171.
S.
Li
,
F.
Chen
,
R.
Schafranek
,
T. J. M.
Bayer
,
K.
Rachut
,
A.
Fuchs
,
S.
Siol
,
M.
Weidner
,
M.
Hohmann
,
V.
Pfeifer
,
J.
Morasch
,
C.
Ghinea
,
E.
Arveux
,
R.
Günzler
,
J.
Gassmann
,
C.
Körber
,
Y.
Gassenbauer
,
F.
Säuberlich
,
G. V.
Rao
,
S.
Payan
,
M.
Maglione
,
C.
Chirila
,
L.
Pintilie
,
L.
Jia
,
K.
Ellmer
,
M.
Naderer
,
K.
Reichmann
,
U.
Böttger
,
S.
Schmelzer
,
R. C.
Frunza
,
H.
Uršič
,
B.
Malič
,
W.-B.
Wu
,
P.
Erhart
, and
A.
Klein
,
Phys. Status Solidi RRL
8
,
571
(
2014
).
172.
Z.
Li
,
M.
Grimsditch
,
X.
Xu
, and
S.-K.
Chan
,
Ferroelectrics
141
,
313
(
1993
).
173.
M. H.
Park
,
Y. H.
Lee
,
H. J.
Kim
,
Y. J.
Kim
,
T.
Moon
,
K. D.
Kim
,
J.
Müller
,
A.
Kersch
,
U.
Schroeder
,
T.
Mikolajick
, and
C. S.
Hwang
,
Adv. Mater.
27
,
1811
(
2015
).
174.
M. J.
Haun
,
E.
Furman
,
H. A.
McKinstry
, and
L. E.
Cross
,
Ferroelectrics
99
,
27
(
1989
).
175.
S. E.
Reyes-Lillo
,
K. F.
Garrity
, and
K. M.
Rabe
,
Phys. Rev. B
90
,
140103(R)
(
2014
).
176.
D.
Lebeugle
,
D.
Colson
,
A.
Forget
, and
M.
Viret
,
Appl. Phys. Lett.
91
,
022907
(
2007
).
177.
J. F.
Scott
,
K.
Watanabe
,
A. J.
Hartmann
, and
R. N.
Lamb
,
Ferroelectrics
225
,
83
(
1999
).
178.
F.
Charra
and
G. G.
Gurzadyan
, in
High Frequency Properties of Dielectric Crystals
, Volume
30B
of Landolt-Börnstein - Group III: Condensed Matter, edited by
D. F.
Nelson
(
Springer-Verlag
,
Berlin
,
2000
).
179.
A.
Kumar
,
R. C.
Rai
,
N. J.
Podraza
,
S.
Denev
,
M.
Ramirez
,
Y.-H.
Chu
,
L. W.
Martin
,
J.
Ihlefeld
,
T.
Heeg
,
J.
Schubert
,
D. G.
Schlom
,
J.
Orenstein
,
R.
Ramesh
,
R. W.
Collins
,
J. L.
Musfeldt
, and
V.
Gopalan
,
Appl. Phys. Lett.
92
,
121915
(
2008
).
180.
C. H.
Peng
,
J.-F.
Chang
, and
S. B.
Desu
, in
Ferroelectric Thin Films II, edited by A. I. Kingon, E. R. Myers, and B. Tuttle
(
Mater. Res. Soc. Symp. Proc.
1992
), Vol.
243
, p.
21
.
181.
V. A.
Shvets
,
V. N.
Kruchinin
, and
V. A.
Gritsenko
,
Opt. Spectrosc.
123
,
728
(
2017
).
182.
T. V.
Perevalov
,
A. K.
Gutakovskii
,
V. N.
Kruchinin
,
V. A.
Gritsenko
, and
I. P.
Prosvirin
,
Mater. Res. Express
6
,
036403
(
2019
).
183.
J.
Fujisawa
,
T.
Eda
, and
M.
Hanaya
,
Chem. Phys. Lett.
685
,
23
(
2017
).
184.
R.
Schafranek
,
S.
Li
,
F.
Chen
,
W.
Wu
, and
A.
Klein
,
Phys. Rev. B
84
,
045317
(
2011
).
185.
M. J.
Haun
,
E.
Furman
,
S. J.
Jang
, and
L. E.
Cross
,
Ferroelectrics
99
,
63
(
1989
).
186.
I.
Vrejoiu
,
G.
Le Rhun
,
L.
Pintilie
,
D.
Hesse
,
M.
Alexe
, and
U.
Gösele
,
Adv. Mater.
18
,
1657
(
2006
).
187.
J.
Chakrabartty
,
R.
Nechache
,
C.
Harnagea
,
S.
Li
, and
F.
Rosei
,
Nanotechnology
27
,
215402
(
2016
).
188.
Y.
Peng
,
G.
Han
,
W.
Xiao
,
Y.
Liu
,
Q.
Li
,
C.
Zhang
,
J.
Zhang
, and
Y.
Hao
,
Superlattices Microstruct.
130
,
519
(
2019
).
189.
S. M.
Sze
,
Physics of Semiconductor Devices
, 2nd ed. (
John Wiley & Sons
,
New York
,
1981
).
190.
B. W.
Wessels
,
Annu. Rev. Mater. Res.
37
,
659
(
2007
).
191.
A.
Rosa
,
D.
Tulli
,
P.
Castera
,
A. M.
Gutierrez
,
A.
Griol
,
M.
Baquero
,
B.
Vilquin
,
F.
Eltes
,
S.
Abel
,
J.
Fompeyrine
, and
P.
Sanchis
,
Opt. Mater. Express
7
,
4328
(
2017
).
192.
M.
Acosta
,
N.
Novak
,
V.
Rojas
,
S.
Patel
,
R.
Vaish
,
J.
Koruza
, Jr.
,
G. A.
Rossetti
, and
J.
Rödel
,
Appl. Phys. Rev.
4
,
041305
(
2017
).
193.
V.
Buscaglia
,
M. T.
Buscaglia
, and
G.
Canu
, “
BaTiO3-based ceramics: Fundamentals, properties and applications
,” in
Encyclopedia of Materials: Technical Ceramics and Glasses
(
Elsevier
,
2020
), p.
1
.
194.
L. A.
Shebanov
,
Phys. Status Solidi (a)
65
,
321
(
1981
).
195.
T.
Terashima
,
Y.
Bando
,
K.
Ijima
,
K.
Yamamoto
,
K.
Hirata
,
K.
Hayashi
,
K.
Kamigaki
, and
H.
Terauchi
,
Phys. Rev, Lett.
65
,
2684
(
1990
).
196.
W. J.
Lin
,
T. Y.
Tseng
,
H. B.
Lu
,
S. L.
Tu
,
S. J.
Yang
, and
I. N.
Lin
,
J. Appl. Phys.
77
,
6466
(
1995
).
197.
Y.
Ishibashi
,
T.
Tsurumi
,
N.
Ohashi
, and
O.
Fukunaga
,
Solid State Ionics
108
,
91
(
1998
).
198.
W.
Tian
,
J. H.
Haeni
,
D. G.
Schlom
, and
X. Q.
Pan
, in
Strain-Induced Elevation of the Spontaneous Polarization in BaTiO3 Thin Films
(Mater. Res. Soc. Symp. Proc.
2001
), Vol.
655
, p.
CC7.8.1
.
199.
Y. S.
Kim
,
D. H.
Kim
,
J. D.
Kim
,
Y. J.
Chang
,
T. W.
Noh
,
J. H.
Kong
,
K.
Char
,
Y. D.
Park
,
S. D.
Bu
,
J.-G.
Yoon
, and
J.-S.
Chung
,
Appl. Phys. Lett.
86
,
102907
(
2005
).
200.
X.
Chen
,
S.
Yang
,
J.-H.
Kim
,
H.-D.
Kim
,
J.-S.
Kim
,
G.
Rojas
,
R.
Skomski
,
H.
Lu
,
A.
Bhattacharya
,
T.
Santos
,
N.
Guisinger
,
M.
Bode
,
A.
Gruverman
, and
A.
Enders
,
New J. Phys.
13
,
083037
(
2011
).
201.
G.
Niu
,
B.
Gautier
,
S.
Yin
,
G.
Saint-Girons
,
P.
Lecoeur
,
V.
Pillard
,
G.
Hollinger
, and
B.
Vilquin
,
Thin Solid Films
520
,
4595
(
2012
).
202.
H. S.
Wang
,
K.
Ma
,
D. F.
Cui
,
Z. Q.
Peng
,
Y. L.
Zhou
,
H. B.
Lu
,
Z. H.
Chen
,
L.
Li
, and
G. Z.
Yang
,
J. Cryst. Growth
177
,
67
(
1997
).
203.
A.
Zhu
,
C.
Li
,
K.
Zhou
,
Y.
Zheng
,
H.
Lu
,
H.
Wang
,
D.
Cui
,
Z.
Chen
, and
Z.
Mai
,
J. Mater. Sci. Lett.
16
,
1898
(
1997
).
204.
G. H.
Lee
,
M.
Yoshimoto
,
T.
Ohnishi
,
K.
Sasaki
, and
H.
Koinuma
,
Mater. Sci. Eng. B
56
,
213
(
1998
).
205.
T.
Suzuki
,
Y.
Nishi
, and
M.
Fujimoto
,
Philos. Mag. A
79
,
2461
(
1999
).
206.
C. L.
Li
,
Z. H.
Chen
,
Y. L.
Zhou
, and
D. F.
Cui
,
J. Phys.: Condens. Matter
13
,
5261
(
2001
).
207.
G. H.
Lee
,
B. C.
Shin
, and
I. S.
Kim
,
Mater. Lett.
50
,
134
(
2001
).
208.
A.
Visinoiu
,
M.
Alexe
,
H. N.
Lee
,
D. N.
Zakharov
,
A.
Pignolet
,
D.
Hesse
, and
U.
Gösele
, in
Initial growth stages of epitaxial BaTiO3 films on vicinal SrTiO3:Nb (001) substrates
(Mater. Res. Soc. Symp. Proc.
2002
), Vol.
688
, p.
C8.4.1
.
209.
J. Y.
Jo
,
Y. S.
Kim
,
D. H.
Kim
,
J. D.
Kim
,
Y. J.
Chang
,
J. H.
Kong
,
Y. D.
Park
,
T. K.
Song
,
J.-G.
Yoon
,
J. S.
Jung
, and
T. W.
Noh
,
Thin Solid Films
486
,
149
(
2005
).
210.
A.
Petraru
,
N. A.
Pertsev
,
H.
Kohlstedt
,
U.
Poppe
,
R.
Waser
,
A.
Solbach
, and
U.
Klemradt
,
J. Appl. Phys.
101
,
114106
(
2007
).
211.
P. S. S. R.
Krishnan
and
P. R.
Munroe
,
J. Asian Ceram. Soc.
1
,
149
(
2013
).
212.
R.
Takahashi
and
M.
Lippmaa
,
ACS Appl. Mater. Interfaces
9
,
21314
(
2017
).
213.
F.
Ambriz-Vargas
,
I.
Velasco-Davalos
,
R.
Thomas
, and
A.
Ruediger
,
J. Vac. Soc. Technol. B
34
,
02M101
(
2016
).
214.
Y. J.
Shin
,
J.
Jiang
,
Y.
Jia
,
F. J.
Walker
, and
C. H.
Ahn
,
APL Mater.
9
,
041104
(
2021
).
215.
Y.
Yano
,
K.
Iijima
,
Y.
Daitoh
,
T.
Terashima
,
Y.
Bando
,
Y.
Watanabe
,
H.
Kasatani
, and
H.
Terauchi
,
J. Appl. Phys.
76
,
7833
(
1994
).
216.
J. H.
Kim
and
S.
Hishita
,
J. Mater. Sci.
30
,
4645
(
1995
).
217.
Y. A.
Boikov
and
T.
Claeson
,
J. Appl. Phys.
89
,
5053
(
2001
).
218.
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
, and
C. B.
Eom
,
Science
306
,
1005
(
2004
).
219.
W.
Zhang
,
L.
Kang
,
M.
Yuan
,
Q.
Yang
, and
J.
Ouyang
,
J. Alloys Compd.
580
,
363
(
2013
).
220.
J.
Schubert
,
O.
Trithaveesak
,
A.
Petraru
,
C. L.
Jia
,
R.
Uecker
,
P.
Reiche
, and
D. G.
Schlom
,
Appl. Phys. Lett.
82
,
3460
(
2003
).
221.
C. H.
Lei
,
Thin Solid Films
515
,
1701
(
2006
).
222.
D. M.
Hwang
,
R.
Ramesh
,
C. Y.
Chen
,
X. D.
Wu
,
A.
Inam
,
M. S.
Hegde
,
B.
Wilkens
,
C. C.
Chang
,
L.
Nazar
,
T.
Venkatesan
,
S.
Miura
,
S.
Matsubara
,
Y.
Miyasaka
, and
N.
Shohata
,
J. Appl. Phys.
68
,
1772
(
1990
).
223.
L. H.
Chang
and
W. A.
Anderson
,
Appl. Surf. Sci.
92
,
52
(
1996
).
224.
J.
Hallmark
,
Z.
Yu
,
R.
Droopad
,
J.
Ramdani
,
J.
Curless
,
C.
Overgaard
,
J.
Finder
,
D.
Marshall
,
J.
Wang
, and
B.
Ooms
,
Integr. Ferroelectr.
27
,
41
(
1999
).
225.
V.
Vaithyanathan
,
J.
Lettieri
,
W.
Tian
,
A.
Sharan
,
A.
Vasudevarao
,
Y. L.
Li
,
A.
Kochhar
,
H.
Ma
,
J.
Levy
,
P.
Zschack
,
J. C.
Woicik
,
L. Q.
Chen
,
V.
Gopalan
, and
D. G.
Schlom
,
J. Appl. Phys.
100
,
024108
(
2006
).
226.
L.
Qiao
and
X.
Bi
,
Appl. Phys. Lett.
92
,
062912
(
2008
).
227.
G.
Niu
,
S.
Yin
,
G.
Saint-Girons
,
B.
Gautier
,
P.
Lecoeur
,
V.
Pillard
,
G.
Hollinger
, and
B.
Vilquin
,
Microelectron. Eng.
88
,
1232
(
2011
).
228.
C.
Dubourdieu
,
J.
Bruley
,
T. M.
Arruda
,
A.
Posadas
,
J.
Jordan-Sweet
,
M. M.
Frank
,
E.
Cartier
,
D. J.
Frank
,
S. V.
Kalinin
,
A. A.
Demkov
, and
V.
Narayanan
,
Nat. Nanotechnol.
8
,
748
(
2013
).
229.
S. R.
Singamaneni
,
W.
Fan
,
J. T.
Prater
, and
J.
Narayan
,
J. Appl. Phys.
116
,
224104
(
2014
).
230.
J. H.
Ngai
,
D. P.
Kumah
,
C. H.
Ahn
, and
F. J.
Walker
,
Appl. Phys. Lett.
104
,
062905
(
2014
).
231.
L.
Kornblum
,
M. D.
Morales-Acosta
,
E. N.
Jin
,
C. H.
Ahn
, and
F. J.
Walker
,
Adv. Mater. Interfaces
2
,
1500193
(
2015
).
232.
E. L.
Lin
,
A. B.
Posadas
,
H. W.
Wu
,
D. J.
Smith
,
A. A.
Demkov
, and
J. G.
Ekerdt
,
J. Cryst. Growth
476
,
6
(
2017
).
233.
Z.
Wang
,
T.
Yasuda
,
S.
Hatatani
, and
S.
Oda
,
Jpn. J. Appl. Phys.
38
,
Part 1
,
6817
(
1999
).
234.
S.
Tinte
,
K. M.
Rabe
, and
D.
Vanderbilt
,
Phys. Rev. B
68
,
144105
(
2003
).
235.
H.
Morioka
,
S.
Yokoyama
,
T.
Oikawa
,
H.
Funakubo
, and
K.
Saito
,
Appl. Phys. Lett.
85
,
3516
(
2004
).
236.
C.
Ederer
and
N. A.
Spaldin
,
Phys. Rev. Lett.
95
,
257601
(
2005
).
237.
A. A.
Belik
,
S.
Iikubo
,
K.
Kodama
,
N.
Igawa
,
S.
Shamoto
,
S.
Niitaka
,
M.
Azuma
,
Y.
Shimakawa
,
M.
Takano
,
F.
Izumi
, and
E.
Takayama-Muromachi
,
Chem. Mater.
18
,
798
(
2006
).
238.
R. J.
Zeches
,
M. D.
Rossell
,
J. X.
Zhang
,
A. J.
Hatt
,
Q.
He
,
C.-H.
Yang
,
A.
Kumar
,
C. H.
Wang
,
A.
Melville
,
C.
Adamo
,
G.
Sheng
,
Y.-H.
Chu
,
J. F.
Ihlefeld
,
R.
Erni
,
C.
Ederer
,
V.
Gopalan
,
L. Q.
Chen
,
D. G.
Schlom
,
N. A.
Spaldin
,
L. W.
Martin
, and
R.
Ramesh
,
Science
326
,
977
(
2009
).
239.
H.
Béa
,
B.
Dupé
,
S.
Fusil
,
R.
Mattana
,
E.
Jacquet
,
B.
Warot-Fonrose
,
F.
Wilhelm
,
A.
Rogalev
,
S.
Petit
,
V.
Cros
,
A.
Anane
,
F.
Petroff
,
K.
Bouzehouane
,
G.
Geneste
,
B.
Dkhil
,
S.
Lisenkov
,
I.
Ponomareva
,
L.
Bellaiche
,
M.
Bibes
, and
A.
Barthélémy
,
Phys. Rev. Lett.
102
,
217603
(
2009
).
240.
L.
Zhang
,
J.
Chen
,
L.
Fan
,
O.
Diéguez
,
J.
Cao
,
Z.
Pan
,
Y.
Wang
,
J.
Wang
,
M.
Kim
,
S.
Deng
,
J.
Wang
,
H.
Wang
,
J.
Deng
,
R.
Yu
,
J. F.
Scott
, and
X.
Xing
,
Science
361
,
494
(
2018
).
241.
A.
Visinoiu
,
M.
Alexe
,
H. N.
Lee
,
D. N.
Zakharov
,
A.
Pignolet
,
D.
Hesse
, and
U.
Gösele
,
J. Appl. Phys.
91
,
10157
(
2002
).
242.
J.
Shin
,
S. V.
Kalinin
,
A. Y.
Borisevich
,
E. W.
Plummer
, and
A. P.
Baddorf
,
Appl. Phys. Lett.
91
,
202901
(
2007
).
243.
M.
Kawai
,
D.
Kan
,
S.
Isojima
,
H.
Kurata
,
S.
Isoda
,
Y.
Shimakawa
,
S.
Kimura
, and
O.
Sakata
,
J. Appl. Phys.
102
,
114311
(
2007
).
244.
T.
Zhao
,
F.
Chen
,
H.
Lu
,
G.
Yang
, and
Z.
Chen
,
J. Appl. Phys.
87
,
7442
(
2000
).
245.
M.
Arredondo
,
M.
Saunders
,
A.
Petraru
,
H.
Kohlstedt
,
I.
Vrejoiu
,
M.
Alexe
,
D.
Hesse
,
N. D.
Browning
,
P.
Munroe
, and
V.
Nagarajan
,
J. Mater. Sci.
44
,
5297
(
2009
).
247.
W.
Pies
and
A.
Weiss
, in
Crystal Structure Data of Inorganic Compounds
, Volume 7b1 of Landolt-Börnstein—Group III: Crystal and Solid State Physics (
Springer-Verlag
,
Berlin
,
1975
).
248.
R.
Allmann
,
W.
Pies
, and
A.
Weiss
, in
Crystal Structure Data of Inorganic Compounds
, Volume 7d2 of Landolt-Börnstein - Group III: Crystal and Solid State Physics (
Springer-Verlag
,
Berlin
,
1980
).
249.
W.
Pies
and
A.
Weiss
, in
Crystal Structure Data of Inorganic Compounds
, Volume 7e of Landolt-Börnstein—Group III: Crystal and Solid State Physics (
Springer-Verlag
,
Berlin
,
1976
).
251.
J. W.
Reiner
,
K. F.
Garrity
,
F. J.
Walker
,
S.
Ismail-Beigi
, and
C. H.
Ahn
,
Phys. Rev. Lett.
101
,
105503
(
2008
).
252.
J. W.
Reiner
,
A. M.
Kolpak
,
Y.
Segal
,
K. F.
Garrity
,
S.
Ismail-Beigi
,
C. H.
Ahn
, and
F. J.
Walker
,
Adv. Mater.
22
,
2919
(
2010
).
253.
S.-H.
Baek
and
C.-B.
Eom
,
Acta Mater.
61
,
2734
(
2013
).
254.
L.
Mazet
,
S. M.
Yang
,
S. V.
Kalinin
,
S.
Schamm-Chardon
, and
C.
Dubourdieu
,
Sci. Technol. Adv. Mater.
16
,
036005
(
2015
).
255.
Y.
Kim
,
C.
Pham
, and
J. P.
Chang
,
J. Phys. D: Appl. Phys.
48
,
063001
(
2015
).
256.
J. H.
Ngai
,
K.
Ahmadi-Majlan
,
J.
Moghadam
,
M.
Chrysler
,
D.
Kumah
,
F. J.
Walker
,
C. H.
Ahn
,
T.
Droubay
,
Y.
Du
,
S. A.
Chambers
,
M.
Bowden
,
X.
Shen
, and
D.
Su
,
J. Mater. Res.
32
,
249
(
2017
).
257.
K. J.
Kormondy
,
Y.
Popoff
,
M.
Sousa
,
F.
Eltes
,
D.
Caimi
,
M. D.
Rossell
,
M.
Fiebig
,
P.
Hoffmann
,
C.
Marchiori
,
M.
Reinke
,
M.
Trassin
,
A. A.
Demkov
,
J.
Fompeyrine
, and
S.
Abel
,
Nanotechnology
28
,
075706
(
2017
).
258.
S. R.
Singameneni
,
J. T.
Prater
, and
J.
Narayan
,
Appl. Phys. Rev.
3
,
031301
(
2016
).
259.
D. P.
Kumah
,
J. H.
Ngai
, and
L.
Kornblum
,
Adv. Funct. Mater.
30
,
1901597
(
2020
).
260.
M.
Spreitzer
,
D.
Klement
,
T. P.
Potǒcnik
,
U.
Trstenjak
,
Z.
Jovanović
,
M. D.
Nguyen
,
H.
Yuan
,
J. E.
ten Elshof
,
E.
Houwman
,
G.
Koster
,
G.
Rijnders
,
J.
Fompeyrine
,
L.
Kornblum
,
D. P.
Fenning
,
Y.
Liang
,
W.-Y.
Tong
, and
P.
Ghosez
,
APL Mater.
9
,
040701
(
2021
).
261.
M. B.
Nardelli
,
F. J.
Walker
, and
R. A.
McKee
,
Phys. Status Solidi (b)
241
,
2279
(
2004
).
262.
F. J.
Walker
and
R. A.
McKee
, “
High-k crystalline gate dielectrics: A research perspective
,” in
High Dielectric Constant Materials
, edited by
H. R.
Huff
and
D. C.
Gilmer
(
Springer-Verlag
,
Berlin
,
2005
), p.
607
.
263.
J.
Robertson
, “
New high-k materials for CMOS applications
,” in
Comprehensive Semiconductor Science and Technology
, edited by
R.
Fornari
(
Elsevier B.V
.,
2011
), Vol.
4
, p.
132
.
264.
R. A.
McKee
,
F. J.
Walker
, and
M. F.
Chisholm
,
Phys. Rev. Lett.
81
,
3014
(
1998
).
265.
R. A.
McKee
,
F. J.
Walker
, and
M. F.
Chisholm
,
Science
293
,
468
(
2001
).
266.
C.
Marchiori
,
M.
Sousa
,
A.
Guiller
,
H.
Siegwart
,
J.-P.
Locquet
,
J.
Fompeyrine
,
G. J.
Norga
, and
J. W.
Seo
,
Appl. Phys. Lett.
88
,
072913
(
2006
).
267.
G.
Delhaye
,
C.
Merckling
,
M.
El-Kazzi
,
G.
Saint-Girons
,
M.
Gendry
,
Y.
Robach
,
G.
Hollinger
,
L.
Largeau
, and
G.
Patriarche
,
J. Appl. Phys.
100
,
124109
(
2006
).
268.
Y.
Segal
,
J. W.
Reiner
,
A. M.
Kolpak
,
Z.
Zhang
,
S.
Ismail-Beigi
,
C. H.
Ahn
, and
F. J.
Walker
,
Phys. Rev. Lett.
102
,
116101
(
2009
).
269.
P.
Ponath
,
K.
Fredrickson
,
A. B.
Posadas
,
Y.
Ren
,
X.
Wu
,
R. K.
Vasudevan
,
M. B.
Okatan
,
S.
Jesse
,
T.
Aoki
,
M. R.
McCartney
,
D. J.
Smith
,
S. V.
Kalinin
,
K.
Lai
, and
A. A.
Demkov
,
Nat. Commun.
6
,
6067
(
2015
).
270.
D. P.
Kumah
,
M.
Dogan
,
J. H.
Ngai
,
D.
Qiu
,
Z.
Zhang
,
D.
Su
,
E. D.
Specht
,
S.
Ismail-Beigi
,
C. H.
Ahn
, and
F. J.
Walker
,
Phys. Rev. Lett.
116
,
106101
(
2016
).
271.
A. S.
Disa
,
F. J.
Walker
, and
C. H.
Ahn
,
Adv. Mat. Interfaces
7
,
1901772
(
2020
).
272.
K.
Nashimoto
,
D. K.
Fork
, and
T. H.
Geballe
,
Appl. Phys. Lett.
60
,
1199
(
1992
).
273.
B.
Meunier
,
R.
Bachelet
,
G.
Grenet
,
C.
Botella
,
P.
Regreny
,
L.
Largeau
,
J.
Penuelas
, and
G.
Saint-Girons
,
J. Cryst. Growth
433
,
139
(
2016
).
274.
S. v
Náray-Szabó
,
Naturwissenschaften
31
,
466
(
1943
).
275.
H. D.
Megaw
,
Proc. Phys. Soc.
58
,
133
(
1946
).
276.
G.
Shirane
,
S.
Hoshino
, and
K.
Suzuki
,
J. Phys. Soc. Jpn.
5
,
453
(
1950
).
277.
G.
Shirane
and
S.
Hoshino
,
J. Phys. Soc. Jpn.
6
,
265
(
1951
).
278.
N.
Sicron
,
B.
Ravel
,
Y.
Yacoby
,
E. A.
Stern
,
F.
Dogan
, and
J. J.
Rehr
,
Phys. Rev. B
50
,
13168
(
1994
).
279.
J. F.
Scott
and
C. A.
Paz de Araujo
,
Science
246
,
1400
(
1989
).
280.
S. A.
Mabud
and
A. M.
Glazer
,
J. Appl. Crystallogr.
12
,
49
(
1979
).
281.
A. M.
Glazer
and
S. A.
Mabud
,
Acta Crystallogr., Sect. B
34
,
1065
(
1978
).
282.
283.
W. L.
Warren
,
J.
Robertson
,
D.
Dimos
,
B. A.
Tuttle
,
G. E.
Pike
, and
D. A.
Payne
,
Phys. Rev. B
53
,
3080
(
1996
).
284.
Y.
Kuroiwa
,
S.
Aoyagi
,
A.
Sawada
,
J.
Harada
,
E.
Nishibori
,
M.
Takata
, and
M.
Sakata
,
Phys. Rev. Lett.
87
,
217601
(
2001
).
285.
X.
Gu
,
N.
Izyumskaya
,
V.
Avrutin
,
H.
Morkoç
,
T. D.
Kang
, and
H.
Lee
,
Appl. Phys. Lett.
89
,
122912
(
2006
).
286.
Y.-M.
Wu
and
J.-T.
Lo
,
Jpn. J. Appl. Phys.
37
,
Part 1
,
5645
(
1998
).
287.
V. R.
Palkar
,
S. C.
Purandare
, and
R.
Pinto
,
J. Phys. D: Appl. Phys.
32
,
R1
(
1999
).
288.
M.
Okuyama
and
Y.
Hamakawa
,
Int. J. Eng. Sci.
29
,
391
(
1991
).
289.
L.
Sun
,
Y.-F.
Chen
,
P.
Li
,
T.
Yu
,
J.-X.
Chen
, and
N.-B.
Ming
,
J. Phys.: Condens. Matter
8
,
10185
(
1996
).
290.
G.
Rispens
and
B.
Noheda
,
Integr. Ferroelectr.
92
,
30
(
2007
).
291.
K.
Wasa
,
Y.
Haneda
,
T.
Sato
,
H.
Adachi
,
I.
Kanno
,
D. G.
Schlom
,
S.
Trolier-McKinstry
,
Q.
Gang
, and
C.-B.
Eom
, “
Step-flow growth of perovskite PbTiO3 thin films epitaxially grown on a miscut SrTiO3 substrate
,” in
Superconducting and Related Oxides: Physics and Nanoengineering III
, edited by
D.
Pavuna
and
I.
Bozovic
(
SPIE
,
1998
), Vol.
3481
, p.
182
.
292.
S.
Venkatesan
,
A.
Vlooswijk
,
B. J.
Kooi
,
A.
Morelli
,
G.
Palasantzas
,
J. T. M.
De Hosson
, and
B.
Noheda
,
Phys. Rev. B
78
,
104112
(
2008
).
293.
Y.-F.
Chen
,
T.
Yu
,
J.-X.
Chen
,
L.
Shun
,
P.
Li
, and
N.-B.
Ming
,
Appl. Phys. Lett.
66
,
148
(
1995
).
294.
Y.
Gao
,
G.
Bai
,
K. L.
Merkle
,
Y.
Shi
,
H. L. M.
Chang
,
Z.
Shen
, and
D. J.
Lam
,
J. Mater. Res.
8
,
145
(
1993
).
295.
H.
Fujisawa
,
Y.
Seioh
,
M.
Kume
, and
M.
Shimizu
,
Jpn. J. Appl. Phys.
47
,
Part 1
,
7505
(
2008
).
296.
B.
Jaffe
, Jr.
,
W. R.
Cook
, and
H.
Jaffe
,
Piezoelectric Ceramics
(
Academic Press
,
London
,
1971
).
297.
D. I.
Woodward
,
J.
Knudsen
, and
I. M.
Reaney
,
Phys. Rev. B
72
,
104110
(
2005
).
298.
B.
Noheda
,
D. E.
Cox
,
G.
Shirane
,
J. A.
Gonzalo
,
L. E.
Cross
, and
S.-E.
Park
,
Appl. Phys. Lett.
74
,
2059
(
1999
).
299.
B.
Noheda
,
J. A.
Gonzalo
,
L. E.
Cross
,
R.
Guo
,
S.-E.
Park
,
D. E.
Cox
, and
G.
Shirane
,
Phys. Rev. B
61
,
8687
(
2000
).
300.
V. R.
Cooper
,
I.
Grinberg
,
N. R.
Martin
, and
A. M.
Rappe
,
AIP Conf. Proc.
626
,
26
(
2002
).
301.
N. A.
Pertsev
,
V. G.
Kukhar
,
H.
Kohlstedt
, and
R.
Waser
,
Phys. Rev. B
67
,
054107
(
2003
).
302.
B.
Noheda
and
D. E.
Cox
,
Phase Transitions
79
,
5
(
2006
).
303.
N.
Zhang
,
H.
Yokota
,
A. M.
Glazer
,
Z.
Ren
,
D. A.
Keen
,
D. S.
Keeble
,
P. A.
Thomas
, and
Z.-G.
Ye
,
Nat. Commun.
5
,
5231
(
2014
).
304.
T. R.
Shrout
and
S. J.
Zhang
,
J. Electroceram.
19
,
111
(
2007
).
305.
R. S.
Solanki
,
S. K.
Mishra
,
A.
Senyshyn
,
I.
Ishii
,
C.
Moriyoshi
,
T.
Suzuki
,
Y.
Kuroiwa
, and
D.
Pandey
,
Phys. Rev. B
86
,
174117
(
2012
).
306.
F.
Cordero
,
F.
Trequattrini
,
F.
Craciun
, and
C.
Galassi
,
Phys. Rev. B
87
,
094108
(
2013
).
307.
M. J.
Haun
,
E.
Furman
,
S. J.
Jang
,
H. A.
McKinstry
, and
L. E.
Cross
,
J. Appl. Phys.
62
,
3331
(
1987
).
308.
C. M.
Foster
,
G.-R.
Bai
,
R.
Csencsits
,
J.
Vetrone
,
R.
Jammy
,
L. A.
Wills
,
E.
Carr
, and
J.
Amano
,
J. Appl. Phys.
81
,
2349
(
1997
).
309.
T.
Oikawa
,
M.
Aratani
,
K.
Saito
, and
H.
Funakubo
,
J. Cryst. Growth
237–239
,
455
(
2002
).
310.
S.
Yokoyama
,
H.
Funakubo
,
H.
Morioka
,
K.
Saito
,
T.
Yamada
, and
M.
Ishikawa
,
Ferroelectrics
389
,
10
(
2009
).
311.
N.
Izyumskaya
,
V.
Avrutin
,
X.
Gu
,
B.
Xiao
,
S.
Chevtchenko
,
J.-G.
Yoon
,
H.
Morkoç
,
L.
Zhou
, and
D. J.
Smith
,
Appl. Phys. Lett.
91
,
182906
(
2007
).
312.
I.
Vrejoiu
,
D.
Hesse
, and
M.
Alexe
, “
Single crystalline PZT films and the impact of extended structural defects on the ferroelectric properties
,” in
Handbook of Advanced Dielectric, Piezoelectric and Ferroelectric Materials
, edited by
Z.-G.
Ye
(
Woodhead Publishing Limited
,
Cambridge
,
2008
), p.
695
.
313.
S.
Gariglio
,
N.
Stucki
,
J.-M.
Triscone
, and
G.
Triscone
,
Appl. Phys. Lett.
90
,
202905
(
2007
).
314.
P.-E.
Janolin
,
J. Mater. Sci.
44
,
5025
(
2009
).
315.
C. A. F.
Vaz
,
J.
Hoffman
,
Y.
Segal
,
J. W.
Reiner
,
R. D.
Grober
,
Z.
Zhang
,
C. H.
Ahn
, and
F. J.
Walker
,
Phys. Rev. Lett.
104
,
127202
(
2010
).
316.
C. A. F.
Vaz
,
Y.
Segal
,
J.
Hoffman
,
F. J.
Walker
, and
C. H.
Ahn
,
J. Vac. Sci. Technol. B
28
,
C5A6
(
2010
).
317.
H. J. A.
Molegraaf
,
J.
Hoffman
,
C. A. F.
Vaz
,
S.
Gariglio
,
D.
van der Marel
,
C. H.
Ahn
, and
J.-M.
Triscone
,
Adv. Mater.
21
,
3470
(
2009
).
318.
S.
Zhang
,
F.
Li
,
J.
Luo
,
R.
Sahul
, and
T. R.
Shrout
,
IEEE Trans. Ultrason., Ferroelectr., Freq. Control
60
,
1572
(
2013
).
319.
S.
Zhang
and
F.
Li
,
J. Appl. Phys.
111
,
031301
(
2012
).
320.
S.
Zhang
,
F.
Li
,
X.
Jiang
,
J.
Kim
,
J.
Luo
, and
X.
Geng
,
Prog. Mater. Sci.
68
,
1
(
2015
).
321.
P.
Miao
,
Y.
Zhao
,
N.
Luo
,
D.
Zhao
,
A.
Chen
,
Z.
Sun
,
M.
Guo
,
M.
Zhu
,
H.
Zhang
, and
Q.
Li
,
Sci. Rep.
6
,
19965
(
2016
).
322.
U.
Gabor
,
D.
Vengust
,
Z.
Samardžija
,
A.
Matavž
,
V.
Bobnar
,
D.
Suvorov
, and
M.
Spreitzer
,
Appl. Surf. Sci.
513
,
145787
(
2020
).
323.
M.
Tyunina
,
J.
Levoska
, and
S.
Leppävuori
,
J. Mater. Sci.: Mater. Electron.
14
,
369
(
2003
).
324.
V.
Nagarajan
,
C. S.
Ganpule
,
B.
Nagaraj
,
S.
Aggarwal
,
S. P.
Alpay
,
A. L.
Roytburd
,
E. D.
Williams
, and
R.
Ramesh
,
Appl. Phys. Lett.
75
,
4183
(
1999
).
325.
M.
Tyunina
,
J.
Levoska
,
P.-E.
Janolin
, and
A.
Dejneka
,
Phys. Rev. B
87
,
224107
(
2013
).
326.
V.
Nagarajan
,
S. P.
Alpay
,
C. S.
Ganpule
,
B. K.
Nagaraj
,
S.
Aggarwal
,
E. D.
Williams
,
A. L.
Roytburd
, and
R.
Ramesh
,
Appl. Phys. Lett.
77
,
438
(
2000
).
327.
M. D.
Nguyen
,
C. T. Q.
Nguyen
,
H. N.
Vu
, and
G.
Rijnders
,
J. Eur. Ceram. Soc.
38
,
95
(
2018
).
328.
X.
Wang
,
L.
Zhang
,
X.
Hao
, and
S.
An
,
Mater. Res. Bull.
65
,
73
(
2015
).
329.
M. D.
Biegalski
,
D. H.
Kim
,
S.
Choudhury
,
L. Q.
Chen
,
H. M.
Christen
, and
K.
Dorr
,
Appl. Phys. Lett.
98
(
14
),
142902
(
2011
).
330.
M. D.
Biegalski
,
K.
Dorr
,
D. H.
Kim
, and
H. M.
Christen
,
Appl. Phys. Lett.
96
,
151905
(
2010
).
331.
C.
Thiele
,
K.
Dörr
,
O.
Bilani
,
J.
Rödel
, and
L.
Schultz
,
Phys. Rev. B
75
,
054408
(
2007
).
332.
A. D.
Rata
,
A.
Herklotz
,
K.
Nenkov
,
L.
Schultz
, and
K.
Dörr
,
Phys. Rev. Lett.
100
,
076401
(
2008
).
333.
M.
Buzzi
,
R. V.
Chopdekar
,
J. L.
Hockel
,
A.
Bur
,
T.
Wu
,
N.
Pilet
,
P.
Warnicke
,
G. P.
Carman
,
L. J.
Heyderman
, and
F.
Nolting
,
Phys. Rev. Lett.
111
,
027204
(
2013
).
334.
S.
Finizio
,
M.
Foerster
,
M.
Buzzi
,
B.
Krüger
,
M.
Jourdan
,
C. A. F.
Vaz
,
J.
Hockel
,
T.
Miyawaki
,
A.
Tkach
,
S.
Valencia
,
F.
Kronast
,
G. P.
Carman
,
F.
Nolting
, and
M.
Kläui
,
Phys. Rev. Appl.
1
,
021001
(
2014
).
335.
K. Y.
Yun
,
D.
Ricinschi
,
T.
Kanashima
,
M.
Noda
, and
M.
Okuyama
,
Jpn. J. Appl. Phys., Part 2
43
,
L647
(
2004
).
336.
K. Y.
Yun
,
M.
Noda
,
M.
Okuyama
,
H.
Saeki
,
H.
Tabata
, and
K.
Saito
,
J. Appl. Phys.
96
,
3399
(
2004
).
337.
V. V.
Shvartsman
,
W.
Kleemann
,
R.
Haumont
, and
J.
Kreisel
,
Appl. Phys. Lett.
90
,
172115
(
2007
).
338.
D.
Lebeugle
,
D.
Colson
,
A.
Forget
,
M.
Viret
,
P.
Bonville
,
J. F.
Marucco
, and
S.
Fusil
,
Phys. Rev. B
76
,
024116
(
2007
).
339.
J. B.
Neaton
,
C.
Ederer
,
U. V.
Waghmare
,
N. A.
Spaldin
, and
K. M.
Rabe
,
Phys. Rev. B
71
,
014113
(
2005
).
340.
P.
Ravindran
,
R.
Vidya
,
A.
Kjekshus
,
H.
Fjellvåg
, and
O.
Eriksson
,
Phys. Rev. B
74
,
224412
(
2006
).
341.
H.
Zheng
,
F.
Straub
,
Q.
Zhan
,
P.-L.
Yang
,
W.-K.
Hsieh
,
F.
Zavaliche
,
Y.-H.
Chu
,
U.
Dahmen
, and
R.
Ramesh
,
Adv. Mater.
18
,
2747
(
2006
).
342.
D.
Lebeugle
,
D.
Colson
,
A.
Forget
,
M.
Viret
,
A. M.
Bataille
, and
A.
Gusakov
,
Phys. Rev. Lett.
100
,
227602
(
2008
).
343.
H.
Béa
,
M.
Bibes
,
F.
Ott
,
B.
Dupeé
,
X.-H.
Zhu
,
S.
Petit
,
S.
Fusil
,
C.
Deranlot
,
K.
Bouzehouane
, and
A.
Barthélémy
,
Phys. Rev. Lett.
100
,
017204
(
2008
).
344.
M.
Bibes
and
A.
Barthélémy
,
Nat. Mater.
7
,
425
(
2008
).
345.
S. M.
Wu
,
S. A.
Cybart
,
P.
Yu
,
M. D.
Rossell
,
J. X.
Zhang
,
R.
Ramesh
, and
R. C.
Dynes
,
Nat. Mater.
9
,
756
(
2010
).
346.
J. T.
Heron
,
M.
Trassin
,
K.
Ashraf
,
M.
Gajek
,
Q.
He
,
S. Y.
Yang
,
D. E.
Nikonov
,
Y.-H.
Chu
,
S.
Salahuddin
, and
R.
Ramesh
,
Phys. Rev. Lett.
107
,
217202
(
2011
).
347.
S. M.
Wu
,
S. A.
Cybart
,
D.
Yi
,
J. M.
Parker
,
R.
Ramesh
, and
R. C.
Dynes
,
Phys. Rev. Lett.
110
,
067202
(
2013
).
348.
J. T.
Heron
,
D. G.
Schlom
, and
R.
Ramesh
,
Appl. Phys. Rev.
1
,
021303
(
2014
).
349.
W.
Saenrang
,
B. A.
Davidson
,
F.
Maccherozzi
,
J. P.
Podkaminer
,
J.
Irwin
,
R. D.
Johnson
,
J. W.
Freeland
,
J.
Íñiguez
,
J. L.
Schad
,
K.
Reierson
,
J. C.
Frederick
,
C. A. F.
Vaz
,
L.
Howald
,
T. H.
Kim
,
S.
Ryu
,
M. v
Veenendaal
,
P. G.
Radaelli
,
S. S.
Dhesi
,
M. S.
Rzchowski
, and
C. B.
Eom
,
Nat. Commun.
8
,
1583
(
2017
).
350.
B.
Xu
,
B.
Dupé
,
C.
Xu
,
H.
Xiang
, and
L.
Bellaiche
,
Phys. Rev. B
98
,
184420
(
2018
).
351.
B.
Xu
,
S.
Meyer
,
M. J.
Verstraete
,
L.
Bellaiche
, and
B.
Dupé
,
Phys. Rev. B
103
,
214423
(
2021
).
352.
I.
Sosnowska
,
T.
Peterlin-Neumaier
, and
E.
Steichele
,
J. Phys. C: Sol. State Phys.
15
,
4835
(
1982
).
353.
S.
Picozzi
and
C.
Ederer
,
J. Phys.: Condens. Matter
21
,
303201
(
2009
).
354.
I.
Gross
,
W.
Akhtar
,
V.
Garcia
,
L. J.
Martínez
,
S.
Chouaieb
,
K.
Garcia
,
C.
Carrétéro
,
A.
Barthélémy
,
P.
Appel
,
P.
Maletinsky
,
J.-V.
Kim
,
J. Y.
Chauleau
,
N.
Jaouen
,
M.
Viret
,
M.
Bibes
,
S.
Fusil
, and
V.
Jacques
,
Nature
549
,
252
(
2017
).
355.
C.
Ederer
and
N. A.
Spaldin
,
Phys. Rev. B
71
,
060401
(
2005
).
356.
M.
Ramazanoglu
,
M.
Laver
,
W.
Ratcliff
 II
,
S. M.
Watson
,
W. C.
Chen
,
A.
Jackson
,
K.
Kothapalli
,
S.
Lee
,
S.-W.
Cheong
, and
V.
Kiryukhin
,
Phys. Rev. Lett.
107
,
207206
(
2011
).
357.
J.
Jeong
,
M.
Duc Le
,
P.
Bourges
,
S.
Petit
,
S.
Furukawa
,
S.-A.
Kim
,
S.
Lee
,
S.-W.
Cheong
, and
J.-G.
Park
,
Phys. Rev. Lett.
113
,
107202
(
2014
).
358.
J.-G.
Park
,
M. D.
Le
,
J.
Jeong
, and
S.
Lee
,
J. Phys.: Condens. Matter
26
,
433202
(
2014
).
359.
F.
Bai
,
J.
Wang
,
M.
Wuttig
,
J.
Li
,
N.
Wang
,
A. P.
Pyatakov
,
A. K.
Zvezdin
,
L. E.
Cross
, and
D.
Viehland
,
Appl. Phys. Lett.
86
,
032511
(
2005
).
360.
H.
Béa
,
M.
Bibes
,
S.
Petit
,
J.
Kreisel
, and
A.
Barthélémy
,
Philos. Mag. Lett.
87
,
165
(
2007
).
361.
D.
Sando
,
A.
Agbelele
,
D.
Rahmedov
,
J.
Liu
,
P.
Rovillain
,
C.
Toulouse
,
I. C.
Infante
,
A. P.
Pyatakov
,
S.
Fusil
,
E.
Jacquet
,
C.
Carrétéro
,
C.
Deranlot
,
S.
Lisenkov
,
D.
Wang
,
J. M.
Le Breton
,
M.
Cazayous
,
A.
Sacuto
,
J.
Juraszek
,
A. K.
Zvezdin
,
L.
Bellaiche
,
B.
Dkhil
,
A.
Barthélémy
, and
M.
Bibes
,
Nat. Mater.
12
(
7
),
641
(
2013
).
362.
D.
Sando
,
F.
Appert
,
B.
Xu
,
O.
Paull
,
S. R.
Burns
,
C.
Carrétéro
,
B.
Dupé
,
Y.
Gallais
,
A.
Sacuto
,
M.
Cazayous
,
B.
Dkhil
,
J. M.
Le Breton
,
A.
Barthélémy
,
M.
Bibes
,
L.
Bellaiche
,
V.
Nagarajan
, and
J.
Juraszek
,
Appl. Phys. Rev.
6
,
041404
(
2019
).
363.
A.
Haykal
,
J.
Fischer
,
W.
Akhtar
,
J.-Y.
Chauleau
,
D.
Sando
,
A.
Finco
,
F.
Godel
,
Y. A.
Birkhölzer
,
C.
Carrétéro
,
N.
Jaouen
,
M.
Bibes
,
M.
Viret
,
S.
Fusil
,
V.
Jacques
, and
V.
Garcia
,
Nat. Commun.
11
,
1704
(
2020
).
364.
J.
Wang
,
J. B.
Neaton
,
H.
Zheng
,
V.
Nagarajan
,
S. B.
Ogale
,
B.
Liu
,
D.
Viehland
,
V.
Vaithyanathan
,
D. G.
Schlom
,
U. V.
Waghmare
,
N. A.
Spaldin
,
K. M.
Rabe
,
M.
Wuttig
, and
R.
Ramesh
,
Science
299
,
1719
(
2003
).
365.
F.
Zavaliche
,
S. Y.
Yang
,
T.
Zhao
,
Y. H.
Chu
,
M. P.
Cruz
,
C. B.
Eom
, and
R.
Ramesh
,
Phase Transitions
79
,
991
(
2006
).
366.
A. M.
Kadomtseva
,
Y. F.
Popov
,
A. P.
Pyatakov
,
G. P.
Vorob'ev
, and
D.
Viehland
,
Phase Transitions
79
,
1019
(
2006
).
367.
R.
Ramesh
,
F.
Zavaliche
,
Y. H.
Chu
,
L. W.
Martin
,
S. Y.
Yang
,
M. P.
Cruz
,
M.
Barry
,
K.
Lee
,
P.
Yang
, and
Q.
Zhan
,
Philos. Mag. Lett.
87
,
155
(
2007
).
368.
S.
Geprägs
,
M.
Opel
,
S. T. B.
Goennenwein
, and
R.
Gross
,
Philos. Mag. Lett.
87
,
141
(
2007
).
369.
G.
Catalan
and
J. F.
Scott
,
Adv. Mater.
21
,
2463
(
2009
).
370.
D.
Sando
,
A.
Barthélémy
, and
M.
Bibes
,
J. Phys.: Condens. Matter
26
,
473201
(
2014
).
371.
D.
Sando
,
B.
Xu
,
L.
Bellaiche
, and
V.
Nagarajan
,
Appl. Phys. Rev.
3
,
011106
(
2016
).
372.
D.
Ricinschi
,
K.-Y.
Yun
, and
M.
Okuyama
,
J. Phys.: Condens. Matter
18
,
L97
(
2006
).
373.
H.
Yamada
,
V.
Garcia
,
S.
Fusil
,
S.
Boyn
,
M.
Marinova
,
A.
Gloter
,
S.
Xavier
,
J.
Grollier
,
E.
Jacquet
,
C.
Carrétéro
,
C.
Deranlot
,
M.
Bibes
, and
A.
Barthélémy
,
ACS Nano
7
,
5385
(
2013
).
374.
J.
Zhou
,
M.
Trassin
,
Q.
He
,
N.
Tamura
,
M.
Kunz
,
C.
Cheng
,
J.
Zhang
,
W.-I.
Liang
,
J.
Seidel
,
C.-L.
Hsin
, and
J.
Wu
,
J. Appl. Phys.
112
,
064102
(
2012
).
375.
K.
Chu
,
B. K.
Jang
,
J. H.
Sung
,
Y. A.
Shin
,
E. S.
Lee
,
K.
Song
,
J. H.
Lee
,
C. S.
Woo
,
S. J.
Kim
,
S. Y.
Choi
,
T. Y.
Koo
,
Y. H.
Kim
,
S. H.
Oh
,
M. H.
Jo
, and
C. H.
Yang
,
Nat. Nanotechnol.
10
,
972
(
2015
).
376.
R. C.
Haislmaier
,
N. J.
Podraza
,
S.
Denev
,
A.
Melville
,
D. G.
Schlom
, and
V.
Gopalan
,
Appl. Phys. Lett.
103
,
031906
(
2013
).
377.
T. S.
Böscke
,
J.
Müller
,
D.
Bräuhaus
,
U.
Schröder
, and
U.
Böttger
,
Appl. Phys. Lett.
99
,
102903
(
2011
).
378.
M. H.
Park
,
T.
Schenk
,
C. M.
Fancher
,
E. D.
Grimley
,
C.
Zhou
,
C.
Richter
,
J. M.
LeBeau
,
J. L.
Jones
,
T.
Mikolajick
, and
U.
Schroeder
,
J. Mater. Chem. C
5
,
4677
(
2017
).
379.
Z.
Fan
,
J.
Chen
, and
J.
Wang
,
J. Adv. Dielectr.
6
,
1630003
(
2016
).
380.
X.
Tian
,
S.
Shibayama
,
T.
Nishimura
,
T.
Yajima
,
S.
Migita
, and
A.
Toriumi
,
Appl. Phys. Lett.
112
,
102902
(
2018
).
381.
R.
Materlik
,
C.
Künneth
, and
A.
Kersch
,
J. Appl. Phys.
117
,
134109
(
2015
).
382.
M. H.
Park
,
Y. H.
Lee
,
H. J.
Kim
,
T.
Schenk
,
W.
Lee
,
K. D.
Kim
,
F. P. G.
Fengler
,
T.
Mikolajick
,
U.
Schroeder
, and
C. S.
Hwang
,
Nanoscale
9
,
9973
(
2017
).
383.
S.
Liu
and
B. M.
Hanrahan
,
Phys. Rev. Mater.
3
,
054404
(
2019
).
384.
M.
Dogan
,
N.
Gong
,
T.-P.
Ma
, and
S.
Ismail-Beigi
,
Phys. Chem. Chem. Phys.
21
,
12150
(
2019
).
385.
S. J.
Kim
,
J.
Mohan
,
S. R.
Summerfelt
, and
J.
Kim
,
JOM
71
,
246
(
2019
).
386.
M. H.
Park
,
Y. H.
Lee
,
T.
Mikolajick
,
U.
Schroeder
, and
C. S.
Hwang
,
Adv. Electron. Mater.
5
,
1800522
(
2019
).
387.
T.
Shimizu
,
K.
Katayama
,
T.
Kiguchi
,
A.
Akama
,
T. J.
Konno
, and
H.
Funakubo
,
Appl. Phys. Lett.
107
,
032910
(
2015
).
388.
K.
Katayama
,
T.
Shimizu
,
O.
Sakata
,
T.
Shiraishi
,
S.
Nakamura
,
T.
Kiguchi
,
A.
Akama
,
T. J.
Konno
,
H.
Uchida
, and
H.
Funakubo
,
Appl. Phys. Lett.
109
,
112901
(
2016
).
389.
T.
Shimizu
,
K.
Katayama
,
T.
Kiguchi
,
A.
Akama
,
T. J.
Konno
,
O.
Sakata
, and
H.
Funakubo
,
Sci. Rep.
6
,
32931
(
2016
).
390.
T.
Mimura
,
T.
Shimizu
,
H.
Uchida
,
O.
Sakata
, and
H.
Funakubo
,
Appl. Phys. Lett.
113
,
102901
(
2018
).
391.
T.
Shimizu
,
J. Ceram. Soc. Jpn.
126
,
667
(
2018
).
392.
K.
Lee
,
T. Y.
Lee
,
S. M.
Yang
,
D. H.
Lee
,
J.
Park
, and
S. C.
Chae
,
Appl. Phys. Lett.
112
,
202901
(
2018
).
393.
T.
Li
,
M.
Ye
,
Z.
Sun
,
N.
Zhang
,
W.
Zhang
,
S.
Inguva
,
C.
Xie
,
L.
Chen
,
Y.
Wang
,
S.
Ke
, and
H.
Huang
,
ACS Appl. Mater. Interfaces
11
,
4139
(
2019
).
394.
J.
Lyu
,
I.
Fina
,
J.
Fontcuberta
, and
F.
Sánchez
,
ACS Appl. Mater. Interfaces
11
,
6224
(
2019
).
395.
I.
Fina
and
F.
Sánchez
,
ACS Appl. Electron. Mater.
3
,
1530
(
2021
).
396.
X.
Li
,
C.
Li
,
Z.
Xu
,
Y.
Li
,
Y.
Yang
,
H.
Hu
,
Z.
Jiang
,
J.
Wang
,
J.
Ren
,
C.
Zheng
,
C.
Lu
, and
Z.
Wen
,
Phys. Status Solidi RRL
15
,
2000481
(
2021
).
397.
T.
Song
,
R.
Bachelet
,
G.
Saint-Girons
,
N.
Dix
,
I.
Fina
, and
F.
Sánchez
,
J. Mater. Chem. C
9
,
12224
(
2021
).
398.
J.
Cao
,
S.
Shi
,
Y.
Zhu
, and
J.
Chen
,
Phys. Status Solidi RRL
15
,
2100025
(
2021
).
399.
Y.
Qi
,
S.
Singh
,
C.
Lau
,
F.-T.
Huang
,
X.
Xu
,
F. J.
Walker
,
C. H.
Ahn
,
S.-W.
Cheong
, and
K. M.
Rabe
,
Phys. Rev. Lett.
125
,
257603
(
2020
).
400.
T.
Mikolajick
,
S.
Slesazeck
,
M. H.
Park
, and
U.
Schroeder
,
MRS Bull.
43
,
340
(
2018
).
401.
M. H.
Park
,
Y. H.
Lee
,
T.
Mikolajick
,
U.
Schroeder
, and
C. S.
Hwang
,
MRS Commun.
8
,
795
(
2018
).
402.
M.
Dragoman
,
M.
Aldrigo
,
D.
Dragoman
,
S.
Iordanescu
,
A.
Dinescu
, and
M.
Modreanu
,
Phys. Status Solidi RRL
15
,
2000521
(
2021
).
403.
M.
Dogan
and
S.
Ismail-Beigi
,
J. Phys. Chem. C
123
,
14350
(
2019
).
404.
M.
Dogan
,
S.
Fernandez-Peña
,
L.
Kornblum
,
Y.
Jia
,
D. P.
Kumah
,
J. W.
Reiner
,
Z.
Krivokapic
,
A. M.
Kolpak
,
S.
Ismail-Beigi
,
C. H.
Ahn
, and
F. J.
Walker
,
Nano Lett.
18
,
241
(
2018
).
405.
K. M.
Rabe
,
Curr. Opin. Solid State Mater. Sci.
9
,
122
(
2005
).
406.
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
, and
D. G.
Schlom
,
Nature
430
,
758
(
2004
).
407.
A.
Antons
,
J. B.
Neaton
,
K. M.
Rabe
, and
D.
Vanderbilt
,
Phys. Rev. B
71
,
024102
(
2005
).
408.
Y. S.
Kim
,
D. J.
Kim
,
T. H.
Kim
,
T. W.
Noh
,
J. S.
Choi
,
B. H.
Park
, and
J.-G.
Yoon
,
Appl. Phys. Lett.
91
,
042908
(
2007
).
409.
S.
Aggarwal
and
R.
Ramesh
,
Annu. Rev. Mater. Sci.
28
,
463
(
1998
).
410.
J. F.
Scott
,
Jpn. J. Appl. Phys.
38
,
Part 1
,
2272
(
1999
).
411.
J. F.
Scott
,
Ferroelectric Memories
(
Springer-Verlag
,
Berlin
,
2000
).
412.
L.
Pintilie
,
C.
Ghica
,
C. M.
Teodorescu
,
I.
Pintilie
,
C.
Chirila
,
I.
Pasuk
,
L.
Trupina
,
L.
Hrib
,
A. G.
Boni
,
N. G.
Apostol
,
L. E.
Abramiuc
,
R.
Negrea
,
M.
Stefan
, and
D.
Ghica
,
Sci. Rep.
5
,
14974
(
2015
).
413.
S. V.
Kalinin
,
Y.
Kim
,
D. D.
Fong
, and
A. N.
Morozovska
,
Rep. Prog. Phys.
81
,
036502
(
2018
).
414.
L. C.
Tănase
,
L. E.
Abramiuc
,
D. G.
Popescu
,
A.-M.
Trandafir
,
N. G.
Apostol
,
I. C.
Bucur
,
L.
Hrib
,
L.
Pintilie
,
I.
Pasuk
,
L.
Trupină
, and
C. M.
Teodorescu
,
Phys. Rev. Appl.
10
,
034020
(
2018
).
415.
P.
Yu
,
W.
Luo
,
D.
Yi
,
J. X.
Zhang
,
M. D.
Rossell
,
L.
You
,
C.-H.
Yang
,
G.
Singh-Bhalla
,
S. Y.
Yang
,
Q.
He
,
Q. M.
Ramasse
,
R.
Erni
,
L. W.
Martin
,
Y. H.
Chu
,
S. T.
Pantelides
,
S. J.
Pennycook
, and
R.
Ramesh
,
Proc. Natl. Acad. Sci.
109
,
9710
(
2012
).
416.
J.
Chen
,
H.
Lu
,
H.-J.
Liu
,
Y.-H.
Chu
,
S.
Dunn
,
K.
Ostrikov
,
A.
Gruverman
, and
N.
Valanoor
,
Appl. Phys. Lett.
102
,
182904
(
2013
).
417.
F. C.
Frank
and
J. H.
van der Merwe
,
Proc. R. Soc. A
198
,
205
(
1949
).
418.
J. H.
van der Merwe
,
J. Appl. Phys.
34
,
117
(
1963
).
419.
J. W.
Matthews
and
A. E.
Blakeslee
,
J. Cryst. Growth
27
,
118
(
1974
).
421.
J. Y.
Tsao
,
Materials Fundamentals of Molecular Beam Epitaxy
(
Academic Press, Inc
.,
1993
).
422.
R.
Beanland
,
D. J.
Dunstan
, and
P. J.
Goodhew
,
Adv. Phys.
45
,
87
(
1996
).
423.
F. K.
LeGoues
,
M.
Copel
, and
R.
Tromp
,
Phys. Rev. Lett.
63
,
1826
(
1989
).
424.
425.
B. J.
Spencer
,
P. W.
Voorhees
, and
S. H.
Davis
,
Phys. Rev. Lett.
67
,
3696
(
1991
).
426.
S. S.
Iyer
and
F. K.
LeGoues
,
J. Appl. Phys.
65
,
4693
(
1989
).
427.
P. S. S. R.
Krishnan
,
A. N.
Morozovska
,
E. A.
Eliseev
,
Q. M.
Ramasse
,
D.
Kepaptsoglou
,
W.-I.
Liang
,
Y.-H.
Chu
,
P.
Munroe
, and
V.
Nagarajan
,
J. Appl. Phys.
115
,
054103
(
2014
).
428.
D. H.
Kim
,
H. N.
Lee
,
M. D.
Biegalski
, and
H. M.
Christen
,
Appl. Phys. Lett.
92
,
012911
(
2008
).
429.
A.
Gruverman
,
B. J.
Rodriguez
,
A. I.
Kingon
,
R. J.
Nemanich
,
A. K.
Tagantsev
,
J. S.
Cross
, and
M.
Tsukada
,
Appl. Phys. Lett.
83
,
728
(
2003
).
430.
G.
Catalan
,
L. J.
Sinnamon
, and
J. M.
Gregg
,
J. Phys.: Condens. Matter
16
,
2253
(
2004
).
431.
G.
Catalan
,
B.
Noheda
,
J.
McAneney
,
L. J.
Sinnamon
, and
J. M.
Gregg
,
Phys. Rev. B
72
,
020102(R)
(
2005
).
432.
P. V.
Yudin
and
A. K.
Tagantsev
,
Nanotechnology
24
,
432001
(
2013
).
433.
B. C.
Jeon
,
D.
Lee
,
M. H.
Lee
,
S. M.
Yang
,
S. C.
Chae
,
T. K.
Song
,
S. D.
Bu
,
J.-S.
Chung
,
J.-G.
Yoon
, and
T. W.
Noh
,
Adv. Mater.
25
,
5643
(
2013
).
434.
R.
Guo
,
L.
Shen
,
H.
Wang
,
Z.
Lim
,
W.
Lu
,
P.
Yang
,
Ariando
,
A.
Gruverman
,
T.
Venkatesan
,
Y. P.
Feng
, and
J.
Chen
,
Adv. Mater. Interfaces
3
,
1600737
(
2016
).
435.
J. L.
Blok
,
D. H. A.
Blank
,
G.
Rijnders
,
K. M.
Rabe
, and
D.
Vanderbilt
,
Phys. Rev. B
84
,
205413
(
2011
).
436.
C.
Daumont
,
W.
Ren
,
I. C.
Infante
,
S.
Lisenkov
,
J.
Allibe
,
C.
Carrétéro
,
S.
Fusil
,
E.
Jacquet
,
T.
Bouvet
,
F.
Bouamrane
,
S.
Prosandeev
,
G.
Geneste
,
B.
Dkhil
,
L.
Bellaiche
,
A.
Barthélémy
, and
M.
Bibes
,
J. Phys.: Condens. Matter
24
,
162202
(
2012
).
437.
C.
Ederer
and
N. A.
Spaldin
,
Phys. Rev. B
71
,
224103
(
2005
).
438.
H. N.
Lee
,
S. M.
Nakhmanson
,
M. F.
Chisholm
,
H. M.
Christen
,
K. M.
Rabe
, and
D.
Vanderbilt
,
Phys. Rev. Lett.
98
,
217602
(
2007
).
439.
J. X.
Zhang
,
Q.
He
,
M.
Trassin
,
W.
Luo
,
D.
Yi
,
M. D.
Rossell
,
P.
Yu
,
L.
You
,
C. H.
Wang
,
C. Y.
Kuo
,
J. T.
Heron
,
Z.
Hu
,
R. J.
Zeches
,
H. J.
Lin
,
A.
Tanaka
,
C. T.
Chen
,
L. H.
Tjeng
,
Y.-H.
Chu
, and
R.
Ramesh
,
Phys. Rev. Lett.
107
,
147602
(
2011
).
440.
A. J.
Hatt
,
N. A.
Spaldin
, and
C.
Ederer
,
Phys. Rev. B
81
,
054109
(
2010
).
441.
N.
Cohen
and
O.
Diéguez
,
Israel J. Chem.
60
,
833
841
(
2020
).
442.
R. W.
Whatmore
,
R.
Clarke
, and
A. M.
Glazer
,
J. Phys. C: Solid State Phys.
11
,
3089
(
1978
).
443.
N. A.
Pertsev
,
A. G.
Zembilgotov
, and
A. K.
Tagantsev
,
Phys. Rev. Lett.
80
,
1988
(
1998
).
444.
M. M.
Saad
,
P.
Baxter
,
R. M.
Bowman
,
J. M.
Gregg
,
F. D.
Morrison
, and
J. F.
Scott
,
J. Phys.: Condens. Matter
16
,
L451
(
2004
).
445.
Y. L.
Li
and
L. Q.
Chen
,
Appl. Phys. Lett.
88
,
072905
(
2006
).
446.
I. C.
Infante
,
S.
Lisenkov
,
B.
Dupé
,
M.
Bibes
,
S.
Fusil
,
E.
Jacquet
,
G.
Geneste
,
S.
Petit
,
A.
Courtial
,
J.
Juraszek
,
L.
Bellaiche
,
A.
Barthélémy
, and
B.
Dkhil
,
Phys. Rev. Lett.
105
,
057601
(
2010
).
447.
S. P.
Alpay
,
I. B.
Misirlioglu
,
V.
Nagarajan
, and
R.
Ramesh
,
Appl. Phys. Lett.
85
,
2044
(
2004
).
448.
I. B.
Misirlioglu
,
A. L.
Vasiliev
,
M.
Aindow
, and
S. P.
Alpay
,
Integr. Ferroelectr.
71
,
67
(
2005
).
449.
I. B.
Misirlioglu
,
G.
Akcay
, and
S. P.
Alpay
,
Integr. Ferroelectr.
83
,
67
(
2006
).
450.
C. L.
Jia
,
S. B.
Mi
,
K.
Urban
,
I.
Vrejoiu
,
M.
Alexe
, and
D.
Hesse
,
Phys. Rev. Lett.
102
,
117601
(
2009
).
451.
A. Yu.
Emelyanov
and
N. A.
Pertsev
,
Phys. Rev. B
68
,
214103
(
2003
).
452.
M.-W.
Chu
,
I.
Szafraniak
,
R.
Scholz
,
C.
Harnagea
,
D.
Hesse
,
M.
Alexe
, and
U.
Gösele
,
Nat. Mater.
3
,
87
(
2004
).
453.
V.
Nagarajan
,
C. L.
Jia
,
H.
Kohlstedt
,
R.
Waser
,
I. B.
Misirlioglu
,
S. P.
Alpay
, and
R.
Ramesh
,
Appl. Phys. Lett.
86
,
192910
(
2005
).
454.
I. B.
Misirlioglu
,
A. L.
Vasiliev
,
S. P.
Alpay
,
M.
Aindow
, and
R.
Ramesh
,
J. Mater. Sci.
41
,
697
(
2006
).
455.
I. B.
Misirlioglu
,
S. P.
Alpay
,
M.
Aindow
, and
V.
Nagarajan
,
Appl. Phys. Lett.
88
,
102906
(
2006
).
456.
I.
Vrejoiu
,
G.
Le Rhun
,
N. D.
Zakharov
,
D.
Hesse
,
L.
Pintilie
, and
M.
Alexe
,
Philos. Mag.
86
,
4477
(
2006
).
457.
W.
Pompe
,
X.
Gong
,
Z.
Suo
, and
J. S.
Speck
,
J. Appl. Phys.
74
,
6012
(
1993
).
458.
J. S.
Speck
and
W.
Pompe
,
J. Appl. Phys.
76
,
466
(
1994
).
459.
J. S.
Speck
,
A. C.
Daykin
,
A.
Seifert
,
A. E.
Romanov
, and
W.
Pompe
,
J. Appl. Phys.
78
,
1696
(
1995
).
460.
J. S.
Speck
,
A.
Seifert
,
W.
Pompe
, and
R.
Ramesh
,
J. Appl. Phys.
76
,
477
(
1994
).
461.
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
, and
R.
Ramesh
,
Nat. Mater.
5
,
823
(
2006
).
462.
D. D.
Fong
,
A. M.
Kolpak
,
J. A.
Eastman
,
S. K.
Streiffer
,
P. H.
Fuoss
,
G. B.
Stephenson
,
C.
Thompson
,
D. M.
Kim
,
K. J.
Choi
,
C. B.
Eom
,
I.
Grinberg
, and
A. M.
Rappe
,
Phys. Rev. Lett.
96
,
127601
(
2006
).
463.
J. E.
Spanier
,
A. M.
Kolpak
,
J. J.
Urban
,
I.
Grinberg
,
L.
Ouyang
,
W. S.
Yun
,
A. M.
Rappe
, and
H.
Park
,
Nano Lett.
6
,
735
(
2006
).
464.
P.
Aguado-Puente
and
J.
Junquera
,
Phys. Rev. Lett.
100
,
177601
(
2008
).
465.
A. M.
Bratkovsky
and
A. P.
Levanyuk
,
J. Comput. Theor. Nanosci.
6
,
465
(
2009
).
466.
J.
Junquera
and
P.
Ghosez
,
Nature
422
,
506
(
2003
).
467.
N.
Sai
,
A. M.
Kolpak
, and
A. M.
Rappe
,
Phys. Rev. B
72
,
020101
(
2005
).
468.
S. P.
Alpay
,
I. B.
Misirlioglu
, and
V.
Nagarajan
,
Appl. Phys. Lett.
90
,
236101
(
2007
).
469.
N. D.
Mermim
and
H.
Wagner
,
Phys. Rev. Lett.
17
,
1133
(
1966
).
470.
Y. S.
Kim
,
J. Y.
Jo
,
D. J.
Kim
,
Y. J.
Chang
,
J. H.
Lee
,
T. W.
Noh
,
T. K.
Song
,
J.-G.
Yoon
,
J.-S.
Chung
,
S. I.
Baik
,
Y.-W.
Kim
, and
C. U.
Jung
,
Appl. Phys. Lett.
88
,
072909
(
2006
).
471.
Y. J.
Shin
,
Y.
Kim
,
S. J.
Kang
,
H. H.
Nahm
,
P.
Murugavel
,
J. R.
Kim
,
M. R.
Cho
,
L.
Wang
,
S. M.
Yang
,
J. G.
Yoon
,
J. S.
Chung
,
M.
Kim
,
H.
Zhou
,
S. H.
Chang
, and
T. W.
Noh
,
Adv. Mater.
29
,
1602795
(
2017
).
472.
S. R.
Lee
,
L.
Baasandorj
,
J. W.
Chang
,
I. W.
Hwang
,
J. R.
Kim
,
J.-G.
Kim
,
K.-T.
Ko
,
S. B.
Shim
,
M. W.
Choi
,
M.
You
,
C.-H.
Yang
,
J.
Kim
, and
J.
Song
,
Nano Lett.
19
,
2243
(
2019
).
473.
T.
Tybell
,
C. H.
Ahn
, and
J.-M.
Triscone
,
Appl. Phys. Lett.
75
,
856
(
1999
).
474.
P.
Maksymovych
,
M.
Huijben
,
M.
Pan
,
S.
Jesse
,
N.
Balke
,
Y.-H.
Chu
,
H. J.
Chang
,
A. Y.
Borisevich
,
A. P.
Baddorf
,
G.
Rijnders
,
D. H. A.
Blank
,
R.
Ramesh
, and
S. V.
Kalinin
,
Phys. Rev. B
85
,
014119
(
2012
).
475.
M.
Tyunina
,
L. D.
Yao
,
D.
Chvostova
,
T.
Kocourek
,
M.
Jelinek
,
A.
Dejneka
, and
S.
van Dijken
,
New J. Phys.
17
,
043048
(
2015
).
476.
L.
Pintilie
and
M.
Alexe
,
J. Appl. Phys.
98
,
124103
(
2005
).
477.
L.
Pintilie
,
I.
Boerasu
,
M. J. M.
Gomes
,
T.
Zhao
,
R.
Ramesh
, and
M.
Alexe
,
J. Appl. Phys.
98
,
124104
(
2005
).
478.
A.
Klein
,
J. Am. Ceram. Soc.
99
,
369
(
2016
).
479.
M.
Stengel
,
P.
Aguado-Puente
,
N. A.
Spaldin
, and
J.
Junquera
,
Phys. Rev. B
83
,
235112
(
2011
).
480.
F.
Chen
,
R.
Schafranek
,
W.
Wu
, and
A.
Klein
,
J. Phys. D: Appl. Phys.
44
,
255301
(
2011
).
481.
F.
Chen
and
A.
Klein
,
Phys. Rev. B
86
,
094105
(
2012
).
482.
D.
Goren
,
N.
Amir
, and
Y.
Nemirovsky
,
J. Appl. Phys.
71
,
318
(
1992
).
483.
L.
Pintilie
,
V.
Stancu
,
L.
Trupina
, and
I.
Pintilie
,
Phys. Rev. B
82
,
085319
(
2010
).
484.
H. T.
Yi
,
T.
Choi
,
S. G.
Choi
,
Y. S.
Oh
, and
S.-W.
Cheong
,
Adv. Mater.
23
,
3403
(
2011
).
485.
X.
Liu
,
Y.
Wang
,
J. D.
Burton
, and
E. Y.
Tsymbal
,
Phys. Rev. B
88
,
165139
(
2013
).
486.
J. E.
Rault
,
G.
Agnus
,
T.
Maroutian
,
V.
Pillard
,
P.
Lecoeur
,
G.
Niu
,
B.
Vilquin
,
M. G.
Silly
,
A.
Bendounan
,
F.
Sirotti
, and
N.
Barrett
,
Phys. Rev. B
87
,
155146
(
2013
).
487.
L.
Pintilie
, “
Charge transport in ferroelectric thin films
,” in
Ferroelectrics—Physical Effects
, edited by
M.
Lallart
(
InTech
,
Rijeka
,
2011
), p.
101
.
488.
G.
Catalan
,
J.
Seidel
,
R.
Ramesh
, and
J. F.
Scott
,
Rev. Mod. Phys.
84
,
119
(
2012
).
489.
N. A.
Spaldin
, “
Novel functionalities at oxide (multi)ferroic domain walls
,”
J. Phys. D: Appl. Phys.
49
,
32
(
2016
).
490.
J.
Seidel
,
L. W.
Martin
,
Q.
He
,
Q.
Zhan
,
Y.-H.
Chu
,
A.
Rother
,
M. E.
Hawkridge
,
P.
Maksymovych
,
P.
Yu
,
M.
Gajek
,
N.
Balke
,
S. V.
Kalinin
,
S.
Gemming
,
F.
Wang
,
G.
Catalan
,
J. F.
Scott
,
N. A.
Spaldin
,
J.
Orenstein
, and
R.
Ramesh
,
Nat. Mater.
8
,
229
(
2009
).
491.
J.
Seidel
,
P.
Maksymovych
,
Y.
Batra
,
A.
Katan
,
S.-Y.
Yang
,
Q.
He
,
A. P.
Baddorf
,
S. V.
Kalinin
,
C.-H.
Yang
,
J.-C.
Yang
,
Y.-H.
Chu
,
E. K. H.
Salje
,
H.
Wormeester
,
M.
Salmeron
, and
R.
Ramesh
,
Phys. Rev. Lett.
105
,
197603
(
2010
).
492.
S.
Farokhipoor
and
B.
Noheda
,
Phys. Rev. Lett.
107
,
127601
(
2011
).
493.
J.
Guyonnet
,
I.
Gaponenko
,
S.
Gariglio
, and
P.
Paruch
,
Adv. Mater.
23
,
5377
(
2011
).
494.
D.
Meier
,
J.
Seidel
,
A.
Cano
,
K.
Delaney
,
Y.
Kumagai
,
M.
Mostovoy
,
N. A.
Spaldin
,
R.
Ramesh
, and
M.
Fiebig
,
Nat. Mater.
11
,
284
(
2012
).
495.
T.
Sluka
,
A. K.
Tagantsev
,
P.
Bednyakov
, and
N.
Setter
,
Nat. Commun.
4
,
1808
(
2013
).
496.
Y.-M.
Kim
,
A.
Morozovska
,
E.
Eliseev
,
M. P.
Oxley
,
R.
Mishra
,
S. M.
Selbach
,
T.
Grande
,
S. T.
Pantelides
,
S. V.
Kalinin
, and
A. Y.
Borisevich
,
Nat. Mater.
13
,
1019
(
2014
).
497.
A.
Zenkevich
,
M.
Minnekaev
,
Y.
Matveyev
,
Y.
Lebedinskii
,
K.
Bulakh
,
A.
Chouprik
,
A.
Baturin
,
K.
Maksimova
,
S.
Thiess
, and
W.
Drube
,
Appl. Phys. Lett.
102
,
062907
(
2013
).
498.
C.-L.
Wu
,
P.-W.
Lee
,
Y.-C.
Chen
,
L.-Y.
Chang
,
C.-H.
Chen
,
C.-W.
Liang
,
P.
Yu
,
Q.
He
,
R.
Ramesh
, and
Y.-H.
Chu
,
Phys. Rev. B
83
,
020103(R)
(
2011
).
499.
B. C.
Huang
,
Y. T.
Chen
,
Y. P.
Chiu
,
Y. C.
Huang
,
J. C.
Yang
, and
Y. C.
Chen
,
Appl. Phys. Lett.
100
,
122903
(
2012
).
500.
G.
Gerra
,
A. K.
Tagantsev
,
N.
Setter
, and
K.
Parlinski
,
Phys. Rev. Lett.
96
(
10
),
107603
(
2006
).
501.
M. F.
Chisholm
,
W.
Luo
,
M. P.
Oxley
,
S. T.
Pantelides
, and
H. N.
Lee
,
Phys. Rev. Lett.
105
,
197602
(
2010
).
502.
I. I.
Naumov
,
L.
Bellaiche
, and
H.
Fu
,
Nature
432
,
737
(
2004
).
503.
J. F.
Scott
,
J. Phys.: Condens. Matter
18
,
R361
(
2006
).
504.
505.
M. J.
Polking
,
A. P.
Alivisatos
, and
R.
Ramesh
,
MRS Commun.
5
,
27
(
2015
).
506.
C. G.
Duan
,
R. F.
Sabirianov
,
W. N.
Mei
,
S. S.
Jaswal
, and
E. Y.
Tsymbal
,
Nano Lett.
6
(
3
),
483
(
2006
).
507.
L.
Bégon-Lours
,
V.
Rouco
,
Q.
Qiao
,
A.
Sander
,
M. A.
Roldán
,
R.
Bernard
,
J.
Trastoy
,
A.
Crassous
,
E.
Jacquet
,
K.
Bouzehouane
,
M.
Bibes
,
J.
Santamaría
,
A.
Barthélémy
,
M.
Varela
, and
J. E.
Villegas
,
Phys. Rev. Mater.
2
,
084405
(
2018
).
508.
C.
Noguera
,
J. Phys.: Condens. Matter
12
,
R367
(
2000
).
509.
N.
Nakagawa
,
H. Y.
Hwang
, and
D. A.
Muller
,
Nat. Mater.
5
,
204
(
2006
).
510.
J.
Goniakowski
,
F.
Finocchi
, and
C.
Noguera
,
Rep. Prog. Phys.
71
,
016501
(
2008
).
511.
S.
Hong
,
S. M.
Nakhmanson
, and
D. D.
Fong
,
Rep. Prog. Phys.
79
,
076501
(
2016
).
512.
X. H.
Liu
,
Y.
Wang
,
P. V.
Lukashev
,
J. D.
Burton
, and
E. Y.
Tsymbal
,
Phys. Rev. B
85
,
125407
(
2012
).
513.
G.
Gerra
,
A. K.
Tagantsev
, and
N.
Setter
,
Phys. Rev. Lett.
98
(
20
),
207601
(
2007
).
514.
H.
Lu
,
X.
Liu
,
J. D.
Burton
,
C. W.
Bark
,
Y.
Wang
,
Y.
Zhang
,
D. J.
Kim
,
A.
Stamm
,
P.
Lukashev
,
D. A.
Felker
,
C. M.
Folkman
,
P.
Gao
,
M. S.
Rzchowski
,
X. Q.
Pan
,
C. B.
Eom
,
E. Y.
Tsymbal
, and
A.
Gruverman
,
Adv. Mater.
24
,
1209
(
2012
).
515.
R. D.
Shannon
and
C. T.
Prewitt
,
Acta Cryst. B
25
,
925
(
1969
).
516.
R. D.
Shannon
and
C. T.
Prewitt
,
Acta Cryst. B
26
,
1046
(
1970
).
517.
A. K.
Tagantsev
,
I.
Stolichnov
,
E. L.
Colla
, and
N.
Setter
,
J. Appl. Phys.
90
,
1387
(
2001
).
518.
X. J.
Lou
,
J. Appl. Phys.
105
,
024101
(
2009
).
519.
P.-E.
Janolin
,
F. L.
Marrec
,
J.
Chevreul
, and
B.
Dkhil
,
Appl. Phys. Lett.
90
,
192910
(
2007
).
520.
P.-E.
Janolin
,
B.
Fraisse
,
F. L.
Marrec
, and
B.
Dkhil
,
Appl. Phys. Lett.
90
,
212904
(
2007
).
521.
S.
Venkatesan
,
B. J.
Kooi
,
J. T. M.
De Hosson
,
A. H. G.
Vlooswijk
, and
B.
Noheda
,
J. Appl. Phys.
102
,
104105
(
2007
).
522.
S.
Stemmer
,
S. K.
Streiffer
,
F.
Ernst
, and
M.
Rühle
,
Phys. Status Solidi (a)
147
,
135
(
1995
).
523.
B. J.
Spencer
,
S. H.
Davis
, and
P. W.
Voorhees
,
Phys. Rev. B
47
,
9760
(
1993
).
524.
B. J.
Spencer
,
P. W.
Voorhees
, and
S. H.
Davis
,
J. Appl. Phys.
73
,
4955
(
1993
).
525.
A. L.
Roitburd
,
Phys. Status Solidi (a)
37
,
329
(
1976
).
526.
N. A.
Pertsev
and
A. Y.
Emel'yanov
,
Phys. Solid State
39
,
109
(
1997
).
527.
Y.
Kamimura
,
K.
Edagawa
, and
S.
Takeuchi
,
Acta Mater.
61
,
294
(
2013
).
528.
C. A. F.
Vaz
,
J.
Hoffman
,
A.-B.
Posadas
, and
C. H.
Ahn
,
Appl. Phys. Lett.
94
,
022504
(
2009
).
529.
D.
Ferré
,
P.
Carrez
, and
P.
Cordier
,
Phys. Rev. B
77
,
014106
(
2008
).
530.
D.
Ferré
,
P.
Carrez
, and
P.
Cordier
,
Phys. Chem. Miner.
36
,
233
(
2009
).
531.
A. M.
Walker
,
P.
Carrez
, and
P.
Cordier
,
Miner. Mag.
74
,
381
(
2010
).
532.
H. P.
Sun
,
W.
Tian
,
X. Q.
Pan
,
J. H.
Haeni
, and
D. G.
Schlom
,
Appl. Phys. Lett.
84
,
3298
(
2004
).
533.
W.
Wunderlich
,
M.
Fujimoto
, and
H.
Ohsato
,
Mater. Sci. Eng. A
309–310
,
148
(
2001
).
534.
M.
Kawai
,
D.
Kan
,
S.
Isojima
,
H.
Kurata
,
S.
Isoda
,
S.
Kimura
,
O.
Sakata
, and
Y.
Shimakawa
,
Deposition Rate Effect on Critical Thickness of BaTiO3 Epitaxial Thin Film Grown on SrTiO3 (001)
(
Mater. Res. Soc. Symp. Proc.
,
2008
), Vol.
1034
, p.
1034-K10
.
535.
C. L.
Canedy
,
H.
Li
,
S. P.
Alpay
,
L.
Salamanca-Riba
,
A. L.
Roytburd
, and
R.
Ramesh
,
Appl. Phys. Lett.
77
,
1695
(
2000
).
536.
V.
Metlenko
,
A. H. H.
Ramadan
,
F.
Gunkel
,
H.
Du
,
H.
Schraknepper
,
S.
Hoffmann-Eifert
,
R.
Dittmann
,
R.
Waser
, and
R. A.
De Souza
,
Nanoscale
6
,
12864
(
2014
).
537.
S. P.
Waldow
and
R. A.
De Souza
,
ACS Appl. Mater. Interfaces
8
,
12246
(
2016
).
538.
E.
Navickas
,
Y.
Chen
,
Q.
Lu
,
W.
Wallisch
,
T. M.
Huber
,
J.
Bernardi
,
M.
Stöger-Pollach
,
G.
Friedbacher
,
H.
Hutter
,
B.
Yildiz
, and
J.
Fleig
,
ACS Nano
11
,
11475
(
2017
).
539.
M.
Arredondo
,
Q. M.
Ramasse
,
M.
Weyland
,
R.
Mahjoub
,
I.
Vrejoiu
,
D.
Hesse
,
N. D.
Browning
,
M.
Alexe
,
P.
Munroe
, and
V.
Nagarajan
,
Adv. Mater.
22
,
2430
(
2010
).
540.
F.
Sandiumenge
,
Front. Mater.
6
,
13
(
2019
).
541.
D.
Lee
and
T. W.
Noh
,
Philos. Trans. R. Soc. A
370
,
4944
(
2012
).
542.
X.
Yu
,
L.
Wu
,
B.
Zhang
,
H.
Zhou
,
Y.
Dong
,
X.
Wu
,
R.
Kou
,
P.
Yang
,
J.
Chen
,
C.-J.
Sun
,
Y.
Zhu
, and
G. M.
Chow
,
Phys. Rev. B
25
,
104405
(
2019
).
543.
C. M.
Foster
,
W.
Pompe
,
A. C.
Daykin
, and
J. S.
Speck
,
J. Appl. Phys.
79
,
1405
(
1996
).
544.
L.
Feigl
,
I. B.
Misirlioglu
,
I.
Vrejoiu
,
M.
Alexe
, and
D.
Hesse
,
J. Appl. Phys.
105
,
061607
(
2009
).
545.
P. E.
Hopkins
,
C.
Adamo
,
L.
Ye
,
B. D.
Huey
,
S. R.
Lee
,
D. G.
Schlom
, and
J. F.
Ihlefeld
,
Appl. Phys. Lett.
102
,
121903
(
2013
).
546.
547.
F.
Riesz
,
Czech. J. Phys.
44
,
131
(
1994
).
548.
H. W.
Jang
,
D.
Ortiz
,
S.-H.
Baek
,
C. M.
Folkman
,
R. R.
Das
,
P.
Shafer
,
Y.
Chen
,
C. T.
Nelson
,
X.
Pan
,
R.
Ramesh
, and
C.-B.
Eom
,
Adv. Mater.
21
,
817
(
2009
).
549.
J. E.
Giencke
,
C. M.
Folkman
,
S.-H.
Baek
, and
C.-B.
Eom
,
Curr. Opin. Solid State Mater. Sci.
18
,
39
(
2014
).
550.
M.
Mietschke
,
S.
Oswald
,
S.
Fähler
,
L.
Schultz
, and
R.
Hühne
,
Thin Solid Films
589
,
792
(
2015
).
551.
P.
Chekhonina
,
M.
Mietschke
,
D.
Pohl
,
F.
Schmidt
,
S.
Fähler
,
W.
Skrotzki
,
K.
Nielsch
, and
R.
Hühne
,
Mater. Charact.
129
,
234
(
2017
).
552.
H.
Fujisawa
,
H.
Yane
,
Y.
Hiki
,
S.
Nakashima
, and
M.
Shimizu
,
J. Korean Phys. Soc.
59
,
2560
(
2011
).
553.
Y.
Lin
and
C. L.
Chen
,
J. Mater. Sci.
44
,
5274
(
2009
).
554.
T. H.
Kim
,
S. H.
Baek
,
S. Y.
Jang
,
S. M.
Yang
,
S. H.
Chang
,
T. K.
Song
,
J.-G.
Yoon
,
C. B.
Eom
,
J.-S.
Chung
, and
T. W.
Noh
,
Appl. Phys. Lett.
98
,
022904
(
2011
).
555.
M. J.
Zhuo
,
Y. L.
Zhu
, and
X. L.
Ma
,
Philos. Mag. Lett.
86
,
469
(
2006
).
556.
D. A.
McQuarrie
,
Statistical Mechanics
(
University Science Books
,
Sausalito
,
2000
).
557.
N.-H.
Chan
,
R. K.
Sharma
, and
D. M.
Smyth
,
J. Am. Ceram. Soc.
64
,
556
(
1981
).
558.
H.
Yamada
and
G. R.
Miller
,
J. Solid State Chem.
6
,
169
(
1973
).
559.
M.
Martin
, “
Diffusion in oxides
,” in
Diffusion in Condensed Matter
(
Springer-Verlag
,
Berlin
,
2005
), p.
209
.
560.
R. A.
De Souza
,
Adv. Funct. Mater.
25
,
6326
(
2015
).
561.
F. A.
Kröger
and
H. J.
Vink
,
Solid State Phys.
3
,
307
(
1956
).
562.
D. M.
Smyth
,
Curr. Opin. Solid State Mater. Sci.
1
,
692
(
1996
).
563.
M. V.
Raymond
and
D. M.
Smyth
,
J. Phys. Chem. Solids
57
,
1507
(
1996
).
564.
D. M.
Smyth
,
E. K.
Chang
, and
D. H.
Liu
,
Phase Transitions
58
,
57
(
1996
).
565.
N.-H.
Chan
and
D. M.
Smyth
,
J. Electrochem. Soc.
123
,
1584
(
1976
).
566.
G. A.
Boni
,
C. F.
Chirila
,
V.
Stancu
,
L.
Amarande
,
I.
Pasuk
,
L.
Trupina
,
C. M.
Istrate
,
C.
Radu
,
A.
Tomulescu
,
S.
Neatu
,
I.
Pintilie
, and
L.
Pintilie
,
Nanomaterials
11
,
1177
(
2021
).
567.
M.
Siebenhofer
,
T.
Huber
,
W.
Artner
,
J.
Fleig
, and
M.
Kubicek
,
Acta Mater.
203
,
116461
(
2021
).
568.
D. M.
Smyth
,
Ferroelectrics
151
,
115
(
1994
).
569.
M.
Morgenbesser
,
A.
Viernstein
,
A.
Schmid
,
C.
Herzig
,
N.
Bodenmueller
,
M.
Kubicek
,
S.
Taibl
,
G.
Bismashofer
,
J.
Stahn
,
C. A. F.
Vaz
,
M.
Doebeli
,
F.
Baiutti
,
J.
de Dios Sirvent
,
M. O.
Liedke
,
M.
Butterling
,
M.
Kaminski
,
M.
Tolkiehn
,
V.
Vonk
,
A.
Stierle
,
A.
Wagner
,
A.
Tarancon
,
A.
Limbeck
, and
J.
Fleig
, “
Unravelling the origin of ultra-low conductivity in doped SrTiO3 thin films: Sr vacancies as the key driver for fermi level pinning
,” (unpublished).
570.
A.
Viernstein
,
G.
Bimashofer
,
M.
Kubicek
,
M.
Morgenbesser
,
T.
Huber
,
E.
Ellmeyer
,
M.
Siebenhofer
,
C. A. F.
Vaz
,
J.
Stahn
, and
J.
Fleig
, “
Mechanism of photo-ionic stoichiometry changes in SrTiO3
,” (unpublished).
571.
R.-A.
Eichel
,
H.
Kungl
, and
P.
Jakes
,
Mater. Technol.
28
,
241
(
2013
).
572.
K.
Shimoyama
,
M.
Kiyohara
,
K.
Kubo
,
A.
Uedono
, and
K.
Yamabe
,
J. Appl. Phys.
92
,
4625
(
2002
).
573.
T.
Tanaka
,
K.
Matsunaga
,
Y.
Ikuhara
, and
T.
Yamamoto
,
Phys, Rev. B
68
,
205213
(
2003
).
574.
S. R.
Basu
,
L. W.
Martin
,
Y. H.
Chu
,
M.
Gajek
,
R.
Ramesh
,
R. C.
Rai
,
X.
Xu
, and
J. L.
Musfeldt
,
Appl. Phys. Lett.
92
,
091905
(
2008
).
575.
Y.
Feng
,
J.
Wu
,
Q.
Chi
,
W.
Li
,
Y.
Yu
, and
W.
Fei
,
Chem. Rev.
120
,
1710
(
2020
).
576.
S.
Pöykkö
and
D. J.
Chadi
,
Appl. Phys. Lett.
76
,
499
(
2000
).
577.
M.
Kotiuga
 et al.,
Proc. Natl. Acad. Sci.
116
,
21992
(
2019
).
578.
T. H.
Etsell
and
S. N.
Flengas
,
Chem. Rev.
70
,
339
(
1970
).
579.
C.
López-Gándara
,
F. M.
Ramos
, and
A.
Cirera
,
J. Sens.
2009
,
258489
(
2009
).
580.
X.-D.
Zhang
,
J.-J.
Li
, and
X.
Guo
,
Rev. Sci. Instrum.
86
,
115103
(
2015
).
581.
C.
Ahamer
,
A. K.
Opitz
,
G. M.
Rupp
, and
J.
Fleig
,
J. Electrochem. Soc.
164
,
F790
(
2017
).
582.
D.
Ricci
,
G.
Bano
,
G.
Pacchioni
, and
F.
Illas
,
Phys, Rev. B
68
,
224105
(
2003
).
583.
W. L.
Warren
,
K.
Vanheusden
,
D.
Dimos
,
G. E.
Pike
, and
B. A.
Tuttle
,
J. Am. Ceram. Soc.
79
,
536
(
1996
).
584.
F.
Cordero
,
Phys, Rev. B
76
,
172106
(
2007
).
585.
D. D.
Cuong
,
B.
Lee
,
K. M.
Choi
,
H.-S.
Ahn
,
S.
Han
, and
J.
Lee
,
Phys. Rev. Lett.
98
,
115503
(
2007
).
586.
D. J.
Wouters
,
G. J.
Willems
, and
H. E.
Maes
,
Microelectron. Eng.
29
,
249
(
1995
).
587.
W. L.
Warren
,
G. E.
Pike
,
B. A.
Tuttle
, and
D.
Dimos
,
Appl. Phys. Lett.
70
,
2010
(
1997
).
588.
C.
Ahn
,
A.
Cavalleri
,
A.
Georges
,
S.
Ismail-Beigi
,
A. J.
Millis
, and
J.-M.
Triscone
,
Nat. Mater.
(published online 2021).
589.
H.
Vo
,
S.
Zhang
,
W.
Wang
, and
G.
Galli
,
J. Chem. Phys.
154
,
174704
(
2021
).
590.
M.
Imada
,
A.
Fujimori
, and
Y.
Tokura
,
Rev. Mod. Phys.
70
,
1039
(
1998
).
591.
A.
Moreo
,
S.
Yunoki
, and
E.
Dagotto
,
Science
283
,
2034
(
1999
).
592.
Y.
Tokura
and
Y.
Tomioka
,
J. Magn. Magn. Mater.
200
,
1
(
1999
).
593.
E.
Dagotto
,
T.
Hotta
, and
A.
Moreo
,
Phys. Rep.
344
,
1
(
2001
).
594.
Y.
Tokura
,
Rep. Prog. Phys.
69
,
797
(
2006
).
595.
A.
Zunger
and
O. I.
Malyi
,
Chem. Rev.
121
,
3031
(
2021
).
596.
X.
Hong
,
J. B.
Yau
,
J. D.
Hoffman
,
C. H.
Ahn
,
Y.
Bason
, and
L.
Klein
,
Phys. Rev. B
74
,
174406
(
2006
).
597.
S.
Dasgupta
,
B.
Das
,
M.
Knapp
,
R. A.
Brand
,
H.
Ehrenberg
,
R.
Kruk
, and
H.
Hahn
,
Adv. Mater.
26
,
4639
(
2014
).
598.
U.
Bauer
,
L.
Yao
,
A. J.
Tan
,
P.
Agrawal
,
S.
Emori
,
H. L.
Tuller
,
S.
van Dijken
, and
G. S. D.
Beach
,
Nat. Mater.
14
,
174
(
2015
).
599.
D. A.
Gilbert
,
A. J.
Grutter
,
E.
Arenholz
,
K.
Liu
,
B. J.
Kirby
,
J. A.
Borchers
, and
B. B.
Maranville
,
Nat. Commun.
7
,
12264
(
2016
).
600.
J. A.
Van Orman
and
K. L.
Crispin
,
Rev. Miner. Geochem.
72
,
757
(
2010
).
601.
A.
van der Ven
,
J.
Bhattacharya
, and
A. A.
Belak
,
Acc. Chem. Res.
46
,
1216
(
2013
).
602.
D.
Cooper
,
C.
Baeumer
,
N.
Bernier
,
A.
Marchewka
,
C.
La Torre
,
R. E.
Dunin-Borkowski
,
S.
Menzel
,
R.
Waser
, and
R.
Dittmann
,
Adv. Mater.
29
,
1700212
(
2017
).
603.
S.
Mathews
,
R.
Ramesh
,
T.
Venkatesan
, and
J.
Benedetto
,
Science
276
,
238
(
1997
).
604.
T.
Venkatesan
,
D. C.
Kundaliya
,
T.
Wu
, and
S. B.
Ogale
,
Philos. Mag. Lett.
87
,
279
(
2007
).
605.
M.
Weisheit
,
S.
Fähler
,
A.
Marty
,
Y.
Souche
,
C.
Poinsignon
, and
D.
Givord
,
Science
315
,
349
(
2007
).
606.
A.
Molinari
,
H.
Hahn
, and
R.
Kruk
,
Adv. Mater.
31
,
1806662
(
2019
).
607.
S. L.
Miller
and
P. J.
McWhorter
,
J. Appl. Phys.
72
,
5999
(
1992
).
608.
Y.-G.
Park
,
T.
Kanki
,
H.-Y.
Lee
,
H.
Tanaka
, and
T.
Kawai
,
Solid-State Electron.
47
,
2221
(
2003
).
609.
H.
Ishiwara
, “
FeFET and ferroelectric random access memories
,” in
Multifunctional Oxide Heterostructures
, edited by
E. Y.
Tsymbal
,
E. R. A.
Dagotto
,
C.-B.
Eom
, and
R.
Ramesh
(
Oxford Scholarship Online
,
2012
), Chap.
12
, p.
340
.
610.
C.
Visani
,
A.
Crassous
, and
J. E.
Villegas
,
Eur. Phys. J. Spec. Top.
222
,
1241
(
2013
).
611.
J. H.
Ngai
,
F. J.
Walker
, and
C. H.
Ahn
,
Annu. Rev. Mater. Res.
44
,
1
(
2014
).
612.
J.
Bardeen
,
L. N.
Cooper
, and
J. R.
Schrieffer
,
Phys. Rev.
106
,
162
(
1957
).
613.
J.
Bardeen
,
L. N.
Cooper
, and
J. R.
Schrieffer
,
Phys. Rev.
108
,
1175
(
1957
).
615.
B.
Keimer
,
S. A.
Kivelson
,
M. R.
Norman
,
S.
Uchida
, and
J.
Zaanen
,
Nature
518
,
179
(
2015
).
616.
I.
Božović
,
J.
Wua
,
X.
He
, and
A. T.
Bollinger
,
Physica C
558
,
30
(
2019
).
617.
R. E.
Glover
 III
and
M. D.
Sherrill
,
Phys. Rev. Lett.
5
,
248
(
1960
).
618.
H. L.
Stadler
,
Phys. Rev. Lett.
14
,
979
(
1965
).
619.
T. G.
Berlincourt
,
Phys. Lett.
29A
,
308
(
1969
).
620.
A.
Cassinese
,
G. M.
De Luca
,
A.
Prigiobbo
,
M.
Salluzzo
, and
R.
Vaglio
,
Appl. Phys. Lett.
84
,
3933
(
2004
).
621.
W. J.
Gallagher
,
IEEE Trans. Magn.
MAG-21
,
709
(
1985
).
622.
T.
Nishino
,
M.
Hatano
,
H.
Hasegawa
,
F.
Murai
,
T.
Kure
,
A.
Hiraiwa
,
K.
Yagi
, and
U.
Kawabe
,
IEEE Electron Device Lett.
10
,
61
(
1989
).
623.
A. W.
Kleinsasser
, “
Superconducting field-effect devices
,” in
The New Superconducting Electronics
, edited by
H.
Weinstock
and
R. W.
Ralston
(
Kluwer Academic Publishers
,
1993
), p.
249
.
624.
J.
Mannhart
,
Supercond. Sci. Technol.
9
,
49
(
1996
).
625.
F.
Paolucci
,
G.
De Simoni
,
P.
Solinas
,
E.
Strambini
,
C.
Puglia
,
N.
Ligato
, and
F.
Giazotto
,
AVS Quantum Sci.
1
,
016501
(
2019
).
626.
N. W.
Ashcroft
and
N. D.
Mermin
,
Solid State Physics
(
Saunder College
,
Philadelphia
,
1976
).
627.
O.
Madelung
,
Introduction to Solid-State Theory
(
Springer-Verlag
,
Berlin
,
1978
).
628.
J.
Mannhart
,
J. G.
Bednorz
,
K. A.
Müller
,
D. G.
Schlom
, and
J.
Ströbel
,
J. Alloys Compd.
195
,
519
(
1993
).
629.
K. A.
Parendo
,
K. H. S. B.
Tan
,
A.
Bhattacharya
,
M.
Eblen-Zayas
,
N. E.
Staley
, and
A. M.
Goldman
,
Phys. Rev. Lett.
94
,
197004
(
2005
).
630.
G.
De Simoni
,
F.
Paolucci
,
P.
Solinas
,
E.
Strambini
, and
F.
Giazotto
,
Nat. Nanotechnol.
13
,
802
(
2018
).
631.
J.
Mannhart
,
J.
Ströbel
,
J. G.
Bednorz
, and
C.
Gerber
,
Appl. Phys. Lett.
62
,
630
(
1993
).
632.
X. X.
Xi
,
J. Supercond.
7
,
137
(
1994
).
633.
J. G.
Bednorz
and
K. A.
Müller
,
Z. Phys. B
64
,
189
(
1986
).
634.
T.
Timusk
and
B.
Statt
,
Rep. Prog. Phys.
62
,
61
(
1999
).
635.
A.
Mourachkine
, in
High-Temperature Superconductivity in Cuprates
, in
Fundamental Theories of Physics
Vol. 123 (
Kluwer Academic Publishers
,
New York
,
2002
).
636.
R. L.
Meng
,
L.
Beauvais
,
X. N.
Zhang
,
Z. J.
Huang
,
Y. Y.
Sun
,
Y. Y.
Xue
, and
C. W.
Chu
,
Physica C
216
,
21
(
1993
).
637.
A. T.
Fjory
,
A. F.
Hebard
,
R. H.
Eick
,
P. M.
Mankiewich
,
R. E.
Howard
, and
M. L.
O'Malley
,
Phys. Rev. Lett.
65
,
3441
(
1990
).
638.
J.
Mannhart
,
D. G.
Schlom
,
J. G.
Bednorz
, and
K. A.
Müller
,
Phys. Rev. Lett.
67
,
2099
(
1991
).
639.
A.
Walkenhorst
,
C.
Doughty
,
X. X.
Xi
,
Q.
Li
,
C. J.
Lobb
,
S. N.
Mao
, and
T.
Venkatesan
,
Phys. Rev. Lett.
69
,
2709
(
1996
).
640.
X. X.
Xi
,
C.
Doughty
,
A.
Walkenhorst
,
C.
Kwon
,
Q.
Li
, and
T.
Venkatesan
,
Phys. Rev. Lett.
68
,
1240
(
1992
).
641.
V. C.
Matijasevic
,
S.
Bogers
,
N. Y.
Chen
,
H. M.
Appelboom
,
P.
Hadley
, and
J. E.
Mooij
,
Physica C
235–240
,
2097
(
1994
).
642.
T.
Frey
,
J.
Mannhart
,
J. G.
Bednorz
, and
E. J.
Williams
,
Phys. Rev. B
51
,
3257
(
1995
).
643.
V.
Talyansky
,
S. B.
Ogale
,
I.
Takeuchi
,
C.
Doughty
, and
T.
Venkatesan
,
Phys. Rev. B
53
,
14575
(
1996
).
644.
D.
Matthey
,
S.
Gariglio
, and
J.-M.
Triscone
,
Appl. Phys. Lett.
83
,
3758
(
2003
).
645.
M.
Salluzzo
,
A.
Cassinese
,
G. M.
De Luca
,
A.
Gambardella
,
A.
Prigiobbo
, and
R.
Vaglio
,
Phys. Rev. B
70
,
214528
(
2004
).
646.
V. V.
Lemanov
,
A. L.
Kholkin
, and
A. B.
Sherman
,
Supercond. Sci. Technol.
6
,
814
(
1993
).
647.
V. V.
Lemanov
,
A. L.
Kholkin
, and
A. B.
Sherman
,
Ferroelectrics
151
,
109
(
1994
).
648.
A.
Ignatiev
,
N.-J.
Wu
,
H.
Lin
,
T.-Q.
Huang
,
S.
Endicter
,
X.-Y.
Li
, and
D.
Liu
,
Integr. Ferroelectr.
10
,
327
(
1995
).
649.
C. H.
Ahn
,
S.
Gariglio
,
P.
Paruch
,
T.
Tybell
,
L.
Antognazza
, and
J.-M.
Triscone
,
Science
284
,
1152
(
1999
).
650.
R.
Aidam
,
D.
Fuchs
, and
R.
Schneider
,
Physica C
328
,
21
(
1999
).
651.
S.
Gariglio
,
C. H.
Ahn
,
D.
Matthey
, and
J.-M.
Triscone
,
Phys. Rev. Lett.
88
,
067002
(
2002
).
652.
D.
Matthey
,
S.
Gariglio
,
C. H.
Ahn
, and
J.-M.
Triscone
,
Physica C
372–376
,
583
(
2002
).
653.
A.
Crassous
,
R.
Bernard
,
S.
Fusil
,
K.
Bouzehouane
,
D. L.
Bourdais
,
S.
Enouz-Vedrenne
,
J.
Briatico
,
M.
Bibes
,
A.
Barthélémy
, and
J. E.
Villegas
,
Phys. Rev. Lett.
107
,
247002
(
2011
).
654.
A.
Crassous
,
R.
Bernard
,
S.
Fusil
,
K.
Bouzehouane
,
J.
Briatico
,
M.
Bibes
,
A.
Barthélémy
, and
J. E.
Villegas
,
J. Appl. Phys.
113
,
024910
(
2013
).
655.
J. T.
Ye
,
S.
Inoue
,
K.
Kobayashi
,
Y.
Kasahara
,
H. T.
Yuan
,
H.
Shimotani
, and
Y.
Iwasa
,
Nat. Mater.
9
,
125
(
2010
).
656.
G.
Dubuis
,
A. T.
Bollinger
,
D.
Pavuna
, and
I.
Božović
,
J. Appl. Phys.
111
,
112632
(
2012
).
657.
E.
Sterpetti
,
J.
Biscaras
,
A.
Erb
, and
A.
Shukla
,
Nat. Commun.
8
,
2060
(
2017
).
658.
Y.
Zhang
and
X.
Xu
,
Comput. Mater. Sci.
179
,
109583
(
2020
).
659.
J. L.
Tallon
and
N. E.
Flower
,
Physica C
204
,
237
(
1993
).
660.
L.
Bégon-Lours
,
V.
Rouco
,
A.
Sander
,
J.
Trastoy
,
R.
Bernard
,
E.
Jacquet
,
K.
Bouzehouane
,
S.
Fusil
,
V.
Garcia
,
A.
Barthélémy
,
M.
Bibes
,
J.
Santamaría
, and
J. E.
Villegas
,
Phys. Rev. Appl.
7
,
064015
(
2017
).
661.
K. S.
Takahashi
,
D.
Matthey
,
D.
Jaccard
,
J.-M.
Triscone
,
K.
Shibuya
,
T.
Ohnishi
, and
M.
Lippmaa
,
Appl. Phys. Lett.
84
,
1722
(
2004
).
662.
K. S.
Takahashi
,
M.
Gabay
,
D.
Jaccard
,
K.
Shibuya
,
T.
Ohnishi
,
M.
Lippmaa
, and
J.-M.
Triscone
,
Nature
441
,
195
(
2006
).
663.
T. M.
Mishonov
,
Phys. Rev. Lett.
67
,
3195
(
1991
).
664.
A. T.
Fiory
,
A. F.
Hebard
,
R. H.
Eick
,
P. M.
Mankiewich
,
R. E.
Howard
, and
M. L.
O'Malley
,
Phys. Rev. Lett.
67
,
3196
(
1991
).
665.
S.-W.
Cheong
,
S. E.
Brown
,
Z.
Fisk
,
R. S.
Kwok
,
J. D.
Thompson
,
E.
Zirngiebl
,
G.
Gruner
,
D. E.
Peterson
,
G. L.
Wells
,
R. B.
Schwarz
, and
J. R.
Cooper
,
Phys. Rev. B
36
,
3913
(
1987
).
666.
J. D.
Jorgensen
,
B. W.
Veal
,
A. P.
Paulikas
,
L. J.
Nowicki
,
G. W.
Crabtree
,
H.
Claus
, and
W. K.
Kwok
,
Phys. Rev. B
41
,
1863
(
1990
).
667.
A. D.
Caviglia
,
S.
Gariglio
,
N.
Reyren
,
D.
Jaccard
,
T.
Schneider
,
M.
Gabay
,
S.
Thiel
,
G.
Hammerl
,
J.
Mannhart
, and
J.-M.
Triscone
,
Nature
456
,
624
(
2008
).
668.
K.
Ueno
,
S.
Nakamura
,
H.
Shimotani
,
A.
Ohtomo
,
N.
Kimura
,
T.
Nojima
,
H.
Aoki
,
Y.
Iwasa
, and
M.
Kawasaki
,
Nat. Mater.
7
,
855
(
2008
).
669.
M.
Ben Shalom
,
M.
Sachs
,
D.
Rakhmilevitch
,
A.
Palevski
, and
Y.
Dagan
,
Phys. Rev. Lett.
104
,
126802
(
2010
).
670.
A. D.
Caviglia
,
M.
Gabay
,
S.
Gariglio
,
N.
Reyren
,
C.
Cancellieri
, and
J.-M.
Triscone
,
Phys. Rev. Lett.
104
,
126803
(
2010
).
671.
G.
Herranz
,
G.
Singh
,
N.
Bergeal
,
A.
Jouan
,
J.
Lesueur
,
J.
Gázquez
,
M.
Varela
,
M.
Scigaj
,
N.
Dix
,
F.
Sánchez
, and
J.
Fontcuberta
,
Nat. Commun.
6
,
6028
(
2015
).
672.
S.
Hurand
,
A.
Jouan
,
C.
Feuillet-Palma
,
G.
Singh
,
J.
Biscaras
,
E.
Lesne
,
N.
Reyren
,
A.
Barthélémy
,
M.
Bibes
,
J. E.
Villegas
,
C.
Ulysse
,
X.
Lafosse
,
M.
Pannetier-Lecoeur
,
S.
Caprara
,
M.
Grilli
,
J.
Lesueur
, and
N.
Bergeal
,
Sci. Rep.
5
,
12751
(
2015
).
673.
E.
Fillis-Tsirakis
,
C.
Richter
,
J.
Mannhart
, and
H.
Boschker
,
New J. Phys.
18
,
013046
(
2016
).
674.
C.
Bell
,
S.
Harashima
,
Y.
Kozuka
,
M.
Kim
,
B. G.
Kim
,
Y.
Hikita
, and
H. Y.
Hwang
,
Phys. Rev. Lett.
103
,
226802
(
2009
).
675.
A. G.
Swartz
,
H.
Inoue
,
T. A.
Merz
,
Y.
Hikita
,
S.
Raghu
,
T. P.
Devereaux
,
S.
Johnston
, and
H. Y.
Hwang
,
Proc. Natl. Acad. Sci.
115
,
1475
(
2018
).
676.
T.
Terashima
,
K.
Shimura
,
Y.
Bando
,
Y.
Matsuda
,
A.
Fujiyama
, and
S.
Komiyama
,
Phys. Rev. Lett.
67
,
1362
(
1991
).
677.
J.-M.
Triscone
and
Ø.
Fischer
,
Rep. Prog. Phys.
60
,
1673
(
1997
).
678.
T.
Tybell
,
C. H.
Ahn
, and
J.-M.
Triscone
,
Appl. Phys. Lett.
72
,
1454
(
1998
).
679.
C. H.
Ahn
,
K. M.
Rabe
, and
J.-M.
Triscone
,
Science
303
,
488
(
2004
).
680.
P. W.
Anderson
,
Phys. Rev.
109
,
1492
(
1958
).
682.
N. F.
Mott
,
Rev. Mod. Phys.
40
,
677
(
1968
).
683.
N. F.
Mott
and
Z.
Zinamon
,
Rep. Prog. Phys.
33
,
881
(
1970
).
684.
N. F.
Mott
,
Metal-Insulator Transitions
(
Taylor & Francies Ltd
.,
London
,
1974
).
685.
N. F.
Mott
,
Phys. Today
31
(
11
),
42
(
1978
).
686.
F.
Gebhard
,
The Mott Metal-Insulator Transition
(
Springer-Verlag
,
Berlin/Heidelberg
,
2000
), p. STMP
137
.
687.
Y.
Zhou
and
S.
Ramanathan
,
Crit. Rev. Solid State Mater. Sci.
38
,
286
(
2013
).
688.
S.
Salahuddin
and
S.
Datta
,
Nano Lett.
8
,
405
(
2008
).
689.
F. A.
McGuire
,
Y.-C.
Lin
,
K.
Price
,
G. B.
Rayner
,
S.
Khandelwal
,
S.
Salahuddin
, and
A. D.
Franklin
,
Nano Lett.
17
,
4801
(
2017
).
690.
J.
Íñiguez
,
P.
Zubko
,
I.
Luk'yanchuk
, and
A.
Cano
,
Nat. Rev. Mater.
4
,
243
(
2019
).
691.
A. H.
Wilson
,
Proc. R. Soc. London, Ser. A
133
,
458
(
1931
).
692.
A. S.
Balchan
and
H. G.
Drickamer
,
J. Chem. Phys.
34
,
1948
(
1961
).
693.
B. M.
Riggleman
and
H. G.
Drickamer
,
J. Chem. Phys.
37
,
446
(
1962
).
694.
B. M.
Riggleman
and
H. G.
Drickamer
,
J. Chem. Phys.
38
,
2721
(
1963
).
695.
N.
Sakai
,
K.
Takemura
, and
K.
Tsuji
,
J. Phys. Soc. Jpn.
51
,
1811
(
1982
).
696.
S.
Desgreniers
,
Y. K.
Vohra
, and
A. L.
Ruoff
,
J. Phys. Chem.
94
,
1117
(
1990
).
697.
Y.
Ma
,
A. R.
Oganov
, and
C. W.
Glass
,
Phys. Rev. B
76
,
064101
(
2007
).
698.
G.
Weck
,
S.
Desgreniers
,
B.
Loubeyre
, and
M.
Mezouar
,
Phys. Rev. Lett.
102
,
255503
(
2009
).
699.
S. F.
Elatresh
and
S. A.
Bonev
,
Phys. Chem. Chem. Phys.
22
,
12577
(
2020
).
700.
T.
Ando
,
A. B.
Fowler
, and
F.
Stern
,
Rev. Mod. Phys.
54
,
437
(
1982
).
701.
A.
Ohtomo
and
H. Y.
Hwang
,
Nature
427
,
423
(
2004
).
702.
M.
Cuoco
and
J.
van den Brink
, “
What theoretical approaches can provide: A perspective on oxide electronics
,”
Appl. Surf. Sci.
482
,
6
(
2019
).
703.
K.
Ueno
,
I. H.
Inoue
,
H.
Akoh
,
M.
Kawasaki
,
Y.
Tokura
, and
H.
Takagi
,
Appl. Phys. Lett.
83
,
1755
(
2003
).
704.
K.
Shibuya
,
T.
Ohnishi
,
M.
Lippmaa
,
M.
Kawasaki
, and
H.
Koinuma
,
Appl. Phys. Lett.
85
,
425
(
2004
).
705.
K.
Shibuya
,
T.
Ohnishi
,
T.
Uozumi
,
T.
Sato
,
M.
Lippmaa
,
M.
Kawasaki
,
K.
Nakajima
,
T.
Chikyow
, and
H.
Koinuma
,
Appl. Phys. Lett.
88
,
212116
(
2006
).
706.
K.
Shibuya
,
T.
Ohnishi
,
T.
Sato
, and
M.
Lippmaa
,
J. Appl. Phys.
102
,
083713
(
2007
).
707.
H.
Ohta
,
Y.
Masuoka
,
R.
Asahi
,
T.
Kato
,
Y.
Ikuhara
,
K.
Nomura
, and
H.
Hosono
,
Appl. Phys. Lett.
95
,
113505
(
2009
).
708.
K.
Ueno
,
I. H.
Inoue
,
T.
Yamada
,
H.
Akoh
,
Y.
Tokura
, and
H.
Takagic
,
Appl. Phys. Lett.
84
,
3726
(
2004
).
709.
A.
Yoshikawa
,
K.
Uchida
,
K.
Koumoto
,
T.
Kato
,
Y.
Ikuhara
, and
H.
Ohta
,
Appl. Phys. Express
2
,
121103
(
2009
).
710.
M. S. J.
Marshall
,
A.
Malashevich
,
A. S.
Disa
,
M.-G.
Han
,
H.
Chen
,
Y.
Zhu
,
S.
Ismail-Beigi
,
F. J.
Walker
, and
C. H.
Ahn
,
Phys. Rev. Appl.
2
,
051001
(
2014
).
711.
N. F.
Mott
and
M.
Kaveh
,
Adv. Phys.
34
,
329
(
1985
).
712.
M. H.
Cohen
,
H.
Fritzsche
, and
S. R.
Ovshinsky
,
Phys. Rev. Lett.
22
,
1065
(
1969
).
713.
T. S.
Santos
,
S. J.
May
,
J. L.
Robertson
, and
A.
Bhattacharya
,
Phys. Rev. B
80
,
155114
(
2009
).
714.
R. E.
Peierls
,
Quantum Theory of Solids
(
Clarendon Press
,
Oxford
,
1955
).
715.
G.
Grüner
,
Rev. Mod. Phys.
60
,
1129
(
1988
).
716.
J.
Varignon
,
M.
Bibes
, and
A.
Zunger
,
Nat. Commun.
10
,
1658
(
2019
).
717.
J.
Varignon
,
M.
Bibes
, and
A.
Zunger
,
Phys. Rev. B
100
,
035119
(
2019
).
718.
Y.
Zhang
,
J.
Furness
,
R.
Zhang
,
Z.
Wang
,
A.
Zunger
, and
J.
Sun
,
Phys. Rev. B
102
,
045112
(
2020
).
719.
I.
Malyi
and
A.
Zunger
,
Appl. Phys. Rev.
7
,
041310
(
2020
).
720.
J.
Zaanen
and
G. A.
Sawatzky
,
J. Solid State Chem.
88
,
8
(
1990
).
721.
K.
van Benthem
,
C.
Elsässer
, and
R. H.
French
,
J. Appl. Phys.
90
,
6156
(
2001
).
722.
Y.
Okimoto
,
T.
Katsufuji
,
Y.
Okada
,
T.
Arima
, and
Y.
Tokura
,
Phys. Rev. B
51
,
9581
(
1995
).
723.
T.
Mizokawa
and
A.
Fujimori
,
Phys. Rev. B
54
,
5368
(
1996
).
724.
S.
Miyasaka
,
Y.
Okimoto
,
M.
Iwama
, and
Y.
Tokura
,
Phys. Rev. B
68
,
100406(R)
(
2003
).
725.
T.
Arima
,
Y.
Tokura
, and
J. B.
Torrance
,
Phys. Rev. B
48
,
17006
(
1993
).
726.
A.
Kudo
,
H.
Yanagi
,
H.
Hosono
, and
H.
Kawazoe
,
Appl. Phys. Lett.
73
,
220
(
1998
).
727.
K.
Tezuka
,
Y.
Hinatsu
,
A.
Nakamura
,
T.
Inami
,
Y.
Shimojo
, and
Y.
Morii
,
J. Solid State Chem.
141
,
404
(
1998
).
728.
M.
Eibschütz
,
S.
Shtrikman
, and
D.
Treves
,
Phys. Rev.
156
,
562
(
1967
).
729.
K.
Mikhalev
,
S.
Verkhovskii
,
A.
Gerashenko
,
A.
Mirmelstein
,
V.
Bobrovskii
,
K.
Kumagai
,
Y.
Furukawa
,
T.
D'yachkova
, and
Y.
Zainulin
,
Phys. Rev. B
69
,
132415
(
2004
).
730.
J. M.
Ginder
,
M. G.
Roe
,
Y.
Song
,
R. P.
McCall
,
J. R.
Gaines
,
E.
Ehrenfreund
, and
A. J.
Epstein
,
Phys. Rev. B
37
,
7506
(
1988
).
731.
A. E.
Bocquet
,
T.
Mizokawa
,
K.
Morikawa
,
A.
Fujimori
,
S. R.
Barman
,
K.
Maiti
,
D. D.
Sarma
,
Y.
Tokura
, and
M.
Onoda
,
Phys. Rev. B
53
,
1161
(
1996
).
732.
T.
Saitoh
,
A. E.
Bocquet
,
T.
Mizokawa
, and
A.
Fujimori
,
Phys. Rev. B
52
,
7934
(
1995
).
733.
A. E.
Bocquet
,
T.
Mizokawa
,
T.
Saitoh
,
H.
Namatame
, and
A.
Fujimori
,
Phys. Rev. B
46
,
3771
(
1992
).
734.
T.
Saitoh
,
T.
Mizokawa
,
A.
Fujimori
,
M.
Abbate
,
Y.
Takeda
, and
M.
Takano
,
Phys. Rev. B
55
,
4257
(
1997
).
735.
M.
Abbate
,
G.
Zampieri
,
F.
Prado
,
A.
Caneiro
,
J. M.
Gonzalez-Calbet
, and
M.
Vallet-Regi
,
Phys. Rev. B
65
,
155101
(
2002
).
736.
B.
Keimer
,
A.
Aharony
,
A.
Auerbach
,
R. J.
Birgeneau
,
A.
Cassanho
,
Y.
Endoh
,
R. W.
Erwin
,
M. A.
Kastner
, and
G.
Shirane
,
Phys. Rev. B
45
,
7430
(
1992
).
737.
A.
Fujimori
,
E.
Takayama-Muromachi
,
Y.
Uchida
, and
B.
Okai
,
Phys. Rev. B
35
,
8814
(
1987
).
738.
Z.
Shen
,
J. W.
Allen
,
J. J.
Yeh
,
J. S.
Kang
,
W.
Ellis
,
W.
Spicer
,
I.
Lindau
,
M. B.
Maple
,
Y. D.
Dalichaouch
,
M. S.
Torikachvili
,
J. Z.
Sun
, and
T. H.
Geballe
,
Phys. Rev. B
36
,
8414
(
1987
).
739.
H.
Yasuoka
,
H.
Mazaki
,
T.
Terashima
, and
Y.
Bando
,
Physica C
175
,
192
(
1991
).
740.
A.
Zibold
,
L.
Widder
,
H. P.
Geserich
,
G.
Bräuchle
,
H.
Claus
,
H. V.
Löhneysen
,
N.
Nücker
,
A.
Erb
, and
G.
Müller-Vogt
,
Physica C
212
,
365
(
1993
).
742.
T.
Hatano
,
Y.
Ogimoto
,
N.
Ogawa
,
M.
Nakano
,
S.
Ono
,
Y.
Tomioka
,
K.
Miyano
,
Y.
Iwasa
, and
Y.
Tokura
,
Sci. Rep.
3
,
2904
(
2013
).
743.
R. M.
Kusters
,
J.
Singleton
,
D. A.
Keen
,
R.
McGreevy
, and
W.
Hayes
,
Physica B
155
,
362
(
1989
).
744.
R.
von Helmolt
,
J.
Wecker
,
B.
Holzapfel
,
L.
Schultz
, and
K.
Samwer
,
Phys. Rev. Lett.
71
,
2331
(
1993
).
745.
K.
Chahara
,
T.
Ohno
,
M.
Kasai
, and
Y.
Kozono
,
Appl. Phys. Lett
63
,
1990
(
1993
).
746.
S.
Jin
,
T. H.
Tiefel
,
M.
McCormack
,
R. A.
Fastnacht
,
R.
Ramesh
, and
L. H.
Chen
,
Science
264
,
413
(
1994
).
747.
Y.
Tokura
,
A.
Urushibara
,
Y.
Moritomo
,
T.
Arima
,
A.
Asamitsu
,
G.
Kido
, and
N.
Furukawa
,
J. Phys. Soc. Jpn.
63
,
3931
(
1994
).
748.
Y.
Konishi
,
Z.
Fang
,
M.
Izumi
,
T.
Manako
,
M.
Kasai
,
H.
Kuwahara
,
M.
Kawasaki
,
K.
Terakura
, and
Y.
Tokura
,
J. Phys. Soc. Jpn.
68
,
3790
(
1999
).
749.
C. A. F.
Vaz
,
J. A.
Moyer
,
D.
Arena
,
C. H.
Ahn
, and
V. E.
Henrich
,
Phys. Rev. B
90
,
024414
(
2014
).
750.
D. M.
Newns
,
T.
Doderer
,
C. C.
Tsuei
,
W. M.
Donath
,
J. A.
Misewich
,
A.
Gupta
,
B. M.
Grossman
,
A.
Schrott
,
B. A.
Scott
,
P. C.
Pattnaik
,
R. J.
von Gutfeld
, and
J. Z.
Sun
,
J. Electroceramics
4
,
339
(
2000
).
751.
J. A.
Misewich
and
A. G.
Schrott
,
Appl. Phys. Lett.
76
,
3632
(
2000
).
752.
Z.
Yang
,
C.
Ko
, and
S.
Ramanathan
,
Annu. Rev. Mater. Res.
41
,
337
(
2011
).
753.
Y.
Zhou
and
S.
Ramanathan
,
Proc. IEEE
103
,
1289
(
2015
).
754.
D. M.
Newns
,
J. A.
Misewich
,
C. C.
Tsuei
,
A.
Gupta
,
B. A.
Scott
, and
A.
Schrott
,
Appl. Phys. Lett.
73
,
780
(
1998
).
755.
M.
Nakamura
,
A.
Sawa
,
H.
Sato
,
H.
Akoh
,
M.
Kawasaki
, and
Y.
Tokura
,
Phys. Rev. B
75
,
155103
(
2007
).
756.
J.
Son
,
S.
Rajan
,
S.
Stemmer
, and
S. J.
Allen
,
J. Appl. Phys.
110
,
084503
(
2011
).
757.
T.
Hatano
,
Z.
Sheng
,
M.
Nakamura
,
M.
Nakano
,
M.
Kawasaki
,
Y.
Iwasa
, and
Y.
Tokura
,
Adv. Mater.
26
,
2874
(
2014
).
758.
C. H.
Ahn
,
J.-M.
Triscone
,
N.
Archibald
,
M.
Decroux
,
R. H.
Hammond
,
T. H.
Geballe
,
Œ.
Fischer
, and
M. R.
Beasley
,
Science
269
,
373
(
1995
).
759.
Y.
Watanabe
,
Appl. Phys. Lett.
66
,
1770
(
1995
).
760.
X.
Hong
,
A.
Posadas
,
A.
Lin
, and
C. H.
Ahn
,
Phys. Rev. B
68
,
134415
(
2003
).
761.
X.
Hong
,
A.
Posadas
, and
C. H.
Ahn
,
Appl. Phys. Lett.
86
,
142501
(
2005
).
762.
A. S.
Dhoot
,
C.
Israel
,
X.
Moya
,
N. D.
Mathur
, and
R. H.
Friend
,
Phys. Rev. Lett.
102
,
136402
(
2009
).
763.
J.
Hoffman
,
X.
Pan
,
J. W.
Reiner
,
F. J.
Walker
,
J. P.
Han
,
C. H.
Ahn
, and
T. P.
Ma
,
Adv. Mater.
22
,
2957
(
2010
).
764.
J.
Hoffman
,
X.
Hong
, and
C. H.
Ahn
,
Nanotechnology
22
,
254014
(
2011
).
765.
H.
Yamada
,
M.
Marinova
,
P.
Altuntas
,
A.
Crassous
,
L.
Bégon-Lours
,
S.
Fusil
,
E.
Jacquet
,
V.
Garcia
,
K.
Bouzehouane
,
A.
Gloter
,
J. E.
Villegas
,
A.
Barthélémy
, and
M.
Bibes
,
Sci. Rep.
3
,
2834
(
2013
).
766.
X.
Li
,
Q.
Zhu
,
L.
Vistoli
,
A.
Barthélémy
,
M.
Bibes
,
S.
Fusil
,
V.
Garcia
, and
A.
Gloter
,
Adv. Mater. Interfaces
7
,
2000601
(
2020
).
767.
L.
Zhang
,
X. G.
Chen
,
H. J.
Gardner
,
M. A.
Koten
,
J. E.
Shield
, and
X.
Hong
,
Appl. Phys. Lett.
107
,
152906
(
2015
).
768.
X.
Chen
,
X.
Zhang
,
M. A.
Koten
,
H.
Chen
,
Z.
Xiao
,
L.
Zhang
,
J. E.
Shield
,
P. A.
Dowben
, and
X.
Hong
,
Adv. Mater.
29
,
1701385
(
2017
).
769.
A.
Urushibara
,
Y.
Moritomo
,
T.
Arima
,
A.
Asamitsu
,
G.
Kido
, and
Y.
Tokura
,
Phys. Rev. B
51
,
14103
(
1995
).
770.
M.
Fäth
,
S.
Freisem
,
A. A.
Menovsky
,
Y.
Tomioka
,
J.
Aarts
, and
J. A.
Mydosh
,
Science
285
,
1540
(
1999
).
771.
A.
Biswas
,
M.
Rajeswari
,
R. C.
Srivastava
,
Y. H.
Li
,
T.
Venkatesan
,
R. L.
Greene
, and
A. J.
Millis
,
Phys. Rev. B
61
,
9665
(
2000
).
772.
J. C.
Loudon
,
N. D.
Mathur
, and
P. A.
Midgley
,
Nature
420
,
797
(
2002
).
773.
T.
Wu
,
S. B.
Ogale
,
J. E.
Garrison
,
B.
Nagaraj
,
A.
Biswas
,
Z.
Chen
,
R. L.
Greene
,
R.
Ramesh
,
T.
Venkatesan
, and
A. J.
Millis
,
Phys. Rev. Lett.
86
,
5998
(
2001
).
774.
M.
Eblen-Zayas
,
A.
Bhattacharya
,
N. E.
Staley
,
A. L.
Kobrinskii
, and
A. M.
Goldman
,
Phys. Rev. Lett.
94
,
037204
(
2005
).
775.
M. L.
Medarde
,
J. Phys.: Condens. Matter
9
,
1679
(
1997
).
776.
J. B.
Torrance
,
P.
Lacorre
,
A. I.
Nazzal
,
E. J.
Ansaldo
, and
C.
Niedermayer
,
Phys. Rev. B
45
,
8209
(
1992
).
777.
S. J.
May
,
J. W.
Kim
,
J. M.
Rondinelli
,
E.
Karapetrova
,
N. A.
Spaldin
,
A.
Bhattacharya
, and
P. J.
Ryan
,
Phys. Rev. B
82
,
014110
(
2010
).
778.
D. P.
Kumah
,
A. S.
Disa
,
J. H.
Ngai
,
H.
Chen
,
A.
Malashevich
,
J. W.
Reiner
,
S.
Ismail-Beigi
,
F. J.
Walker
, and
C. H.
Ahn
,
Adv. Mater.
26
,
1935
(
2014
).
779.
D. P.
Kumah
,
A.
Malashevich
,
A. S.
Disa
,
D. A.
Arena
,
F. J.
Walker
,
S.
Ismail-Beigi
, and
C. H.
Ahn
,
Phys. Rev. Appl.
2
,
054004
(
2014
).
780.
A.
Malashevich
,
M. S. J.
Marshall
,
C.
Visani
,
A. S.
Disa
,
H.
Xu
,
F. J.
Walker
,
C. H.
Ahn
, and
S.
Ismail-Beigi
,
Nano Lett.
18
,
573
(
2018
).
781.
K. I.
Kugel'
and
D. I.
Khomskii
,
Sov. Phys. Usp.
25
,
231
(
1982
).
782.
V. I.
Anisimov
,
J.
Zaanen
, and
O. K.
Andersen
,
Phys. Rev. B
44
,
943
(
1991
).
783.
A. I.
Liechtenstein
,
V. I.
Anisimov
, and
J.
Zaanen
,
Phys. Rev. B
52
,
R5467
(
1995
).
784.
Y.
Tokura
and
N.
Nagaosa
,
Science
288
,
462
(
2000
).
785.
D. I.
Khomskii
and
M. V.
Mostovoy
,
J. Phys. A: Math. Gen.
36
,
9197
(
2003
).
786.
J.
van den Brink
,
G.
Khaliullin
, and
D.
Khomskii
, “
Orbital effects in manganites
,” in
Colossal Magnetoresistive Manganites
, edited by
T.
Chatterji
(
Kluwer Academic Publishers
,
Dordrecht
,
2004
), p.
263
.
787.
H. A.
Jahn
and
E.
Teller
,
Proc. R. Soc. London, Ser. A
161
,
220
(
1937
).
788.
Y.
Murakami
,
S.
Ishihara
, and
S.
Maekawa
, “
Charge and orbital ordering of manganites observed by resonant x-ray scattering
,” in
Colossal Magnetoresistive Manganites
, edited by
T.
Chatterji
(
Kluwer Academic Publishers
,
Dordrecht
,
2004
), p.
383
.
789.
T.
Miller
,
M.
Gensch
, and
S.
Wall
,
Phys. Scr.
91
,
124002
(
2016
).
790.
M. D. N.
Regueiro
,
M.
Altarelli
, and
C. T.
Chen
,
Phys. Rev. B
51
,
629
(
1995
).
791.
S.
Konno
,
K.
Taniguchi
,
H.
Sagayama
, and
T.
Arima
,
Appl. Phys. Express
2
,
033004
(
2009
).
792.
J.
Varignon
,
N. C.
Bristowe
,
E.
Bousquet
, and
P.
Ghosez
,
Sci. Rep.
5
,
15364
(
2015
).
793.
L.
Chen
,
C.
Xu
,
H.
Tian
,
H.
Xiang
,
J.
Íñiguez
,
Y.
Yang
, and
L.
Bellaiche
,
Phys. Rev. Lett.
122
,
247701
(
2019
).
794.
C.
Aruta
,
G.
Ghiringhelli
,
A.
Tebano
,
N. G.
Boggio
,
N. B.
Brookes
,
P. G.
Medaglia
, and
G.
Balestrino
,
Phys. Rev. B
73
,
235121
(
2006
).
795.
C.
Aruta
,
C.
Adamo
,
A.
Galdi
,
P.
Orgiani
,
V.
Bisogni
,
N. B.
Brookes
,
J. C.
Cezar
,
P.
Thakur
,
C. A.
Perroni
,
G.
De Filippis
,
V.
Cataudella
,
D. G.
Schlom
,
L.
Maritato
, and
G.
Ghiringhelli
,
Phys. Rev. B
80
,
140405(R)
(
2009
).
796.
D.
Pesquera
,
G.
Herranz
,
A.
Barla
,
E.
Pellegrin
,
F.
Bondino
,
E.
Magnano
,
F.
Sánchez
, and
J.
Fontcuberta
,
Nat. Commun.
3
,
1189
(
2012
).
797.
P.
Yu
,
J.-S.
Lee
,
S.
Okamoto
,
M. D.
Rossell
,
M.
Huijben
,
C.-H.
Yang
,
Q.
He
,
J. X.
Zhang
,
S. Y.
Yang
,
M. J.
Lee
,
Q. M.
Ramasse
,
R.
Erni
,
Y.-H.
Chu
,
D. A.
Arena
,
C.-C.
Kao
,
L. W.
Martin
, and
R.
Ramesh
,
Phys. Rev. Lett.
105
,
027201
(
2010
).
798.
J.-C.
Yang
,
Q.
He
,
P.
Yu
, and
Y.-H.
Chu
,
Annu. Rev. Mater. Res.
45
,
249
(
2015
).
799.
H.
Chen
,
Q.
Qiao
,
M. S. J.
Marshall
,
A. B.
Georgescu
,
A.
Gulec
,
P. J.
Phillips
,
R. F.
Klie
,
F. J.
Walker
,
C. H.
Ahn
, and
S.
Ismail-Beigi
,
Nano Lett.
14
,
4965
(
2014
).
800.
D.
Preziosi
,
M.
Alexe
,
D.
Hesse
, and
M.
Salluzzo
,
Phys. Rev. Lett.
115
,
157401
(
2015
).
801.
B. T.
Thole
,
P.
Carra
,
F.
Sette
, and
G.
van der Laan
,
Phys. Rev. Lett.
68
,
1943
(
1992
).
802.
P.
Carra
,
B. T.
Thole
,
M.
Altarelli
, and
X.
Wang
,
Phys. Rev. Lett.
70
,
694
(
1993
).
803.
A.
Ankudinov
and
J. J.
Rehr
,
Phys. Rev. B
51
,
1282
(
1995
).
804.
D. J.
Huang
,
H.-T.
Jeng
,
C. F.
Chang
,
G. Y.
Guo
,
J.
Chen
,
W. P.
Wu
,
S. C.
Chung
,
S. G.
Shyu
,
C. C.
Wu
,
H.-J.
Lin
, and
C. T.
Chen
,
Phys. Rev. B
66
,
174440
(
2002
).
805.
A. L.
Ankudinov
,
J. J.
Rehr
,
H.
Wende
,
A.
Scherz
, and
K.
Baberschke
,
Europhys. Lett.
66
,
441
(
2004
).
806.
E. J.
Verwey
,
Nature
144
,
327
(
1939
).
807.
E. J. W.
Verwey
and
P. W.
Haayman
,
Physics
8
,
979
(
1941
).
808.
F.
de Boer
,
J. H.
van Santen
, and
E. J. W.
Verwey
,
J. Chem. Phys.
18
,
1032
(
1950
).
809.
P. W.
Anderson
,
Phys. Rev.
102
,
1008
(
1956
).
810.
K.
Yamauchi
and
S.
Picozzi
,
Phys. Rev. B
85
,
085131
(
2012
).
811.
M.
Angst
,
Phys. Status Solidi RRL
7
,
383
(
2013
).
812.
E. O.
Wollan
and
W. C.
Koehler
,
Phys. Rev.
100
,
545
(
1955
).
813.
J. B.
Goodenough
,
Phys. Rev.
100
,
564
(
1955
).
814.
C. H.
Chen
and
S.-W.
Cheong
,
Phys. Rev. Lett.
76
,
4042
(
1996
).
815.
P. G.
Radaelli
,
D. E.
Cox
,
M.
Marezio
, and
S.-W.
Cheong
,
Phys. Rev. Lett.
55
,
3015
(
1997
).
816.
Z.
Jirák
,
S.
Krupička
,
Z.
Šimša
,
M.
Dlouhá
, and
S.
Vratislav
,
J. Magn. Magn. Mater.
53
,
153
(
1985
).
817.
M. v
Zimmermann
,
J. P.
Hill
,
D.
Gibbs
,
M.
Blume
,
D.
Casa
,
B.
Keimer
,
Y.
Murakami
,
Y.
Tomioka
, and
Y.
Tokura
,
Phys. Rev. Lett.
83
,
4872
(
1999
).
818.
M. v
Zimmermann
,
C. S.
Nelson
,
J. P.
Hill
,
D.
Gibbs
,
M.
Blume
,
D.
Casa
,
B.
Keimer
,
Y.
Murakami
,
C.-C.
Kao
,
C.
Venkataraman
,
T.
Gog
,
Y.
Tomioka
, and
Y.
Tokura
,
Phys. Rev. B
64
,
195133
(
2001
).
819.
C.
Frontera
,
J. L.
García-Muñoz
,
M. A. G.
Aranda
,
C.
Ritter
,
A.
Llobet
,
L.
Ranno
,
M.
Respaud
,
J.
Vanacken
,
J. M.
Broto
,
A. E.
Carrillo
,
A.
Calleja
, and
J.
García
,
Appl. Phys. A
74
,
S1787
(
2002
).
820.
B. H.
Kim
,
J. S.
Kim
,
T. H.
Park
,
D. S.
Lee
,
K. H.
Kim
,
B. G.
Kim
, and
Y. W.
Park
,
J. Phys.: Condens. Matter
19
,
296205
(
2007
).
821.
I. E.
Baggari
,
B. H.
Savitzky
,
A. S.
Admasu
,
J.
Kim
,
S.-W.
Cheong
,
R.
Hovdend
, and
L. F.
Kourkoutis
,
Proc. Natl. Acad. Sci.
115
,
1445
(
2018
).
822.
R.
Comin
and
A.
Damascelli
,
Annu. Rev. Condens. Matter Phys.
7
,
369
(
2016
).
823.
C.
Schüßler-Langeheine
,
J.
Schlappa
,
A.
Tanaka
,
Z.
Hu
,
C. F.
Chang
,
E.
Schierle
,
M.
Benomar
,
H.
Ott
,
E.
Weschke
,
G.
Kaindl
,
O.
Friedt
,
G. A.
Sawatzky
,
H.-J.
Lin
,
C. T.
Chen
,
M.
Braden
, and
L. H.
Tjeng
,
Phys. Rev. Lett.
95
,
156402
(
2005
).
824.
J.
García
and
G.
Subías
,
Phys. Rev. B
68
,
127101
(
2003
).
825.
M. v
Zimmermann
,
S.
Grenier
,
C. S.
Nelson
,
J. P.
Hill
,
D.
Gibbs
,
M.
Blume
,
D.
Casa
,
B.
Keimer
,
Y.
Murakami
,
C.-C.
Kao
,
C.
Venkataraman
,
T.
Gog
,
Y.
Tomioka
, and
Y.
Tokura
,
Phys. Rev. B
68
,
127102
(
2003
).
826.
S.
Grenier
,
J. P.
Hill
,
D.
Gibbs
,
K. J.
Thomas
,
M. V.
Zimmermann
,
C. S.
Nelson
,
V.
Kiryukhin
,
Y.
Tokura
,
Y.
Tomioka
,
D.
Casa
,
T.
Gog
, and
C.
Venkataraman
,
Phys. Rev. B
69
,
134419
(
2004
).
827.
G.
Subías
,
J.
García
,
M. G.
Proietti
, and
J.
Blasco
,
Phys. Rev. B
56
,
8183
(
1997
).
828.
J.
García
,
M. C.
Sánchez
,
J.
Blasco
,
G.
Subías
, and
M. G.
Proietti
,
J. Phys. Condens. Matter
13
,
3243
(
2001
).
829.
F.
Bridges
,
C. H.
Booth
,
M.
Anderson
,
G. H.
Kwei
,
J. J.
Neumeier
,
J.
Snyder
,
J.
Mitchell
,
J. S.
Gardner
, and
E.
Brosha
,
Phys. Rev. B
63
,
214405
(
2001
).
830.
J.
Herrero-Martín
,
J.
García
,
G.
Subías
,
J.
Blasco
, and
M. C.
Sánchez
,
Phys. Rev. B
70
,
024408
(
2004
).
831.
K. J.
Thomas
,
J. P.
Hill
,
S.
Grenier
,
Y.-J.
Kim
,
P.
Abbamonte
,
L.
Venema
,
A.
Rusydi
,
Y.
Tomioka
,
Y.
Tokura
,
D. F.
McMorrow
,
G.
Sawatzky
, and
M.
van Veenendaal
,
Phys. Rev. Lett.
92
,
237204
(
2004
).
832.
R. J.
Goff
and
J. P.
Attfield
,
Phys. Rev. B
70
,
140404(R)
(
2004
).
833.
H.
Kawano
,
R.
Kajimoto
,
H.
Yoshizawa
,
Y.
Tomioka
,
H.
Kuwahara
, and
Y.
Tokura
,
Phys. Rev. Lett.
78
,
4253
(
1997
).
834.
J.
Herrero-Martín
,
J.
García
,
G.
Subías
,
J.
Blasco
,
M. C.
Sánchez
, and
S.
Stanescu
,
Phys. Rev. B
73
,
224407
(
2006
).
835.
M. S.
Senn
,
J. P.
Wright
, and
J. P.
Attfield
,
Nature
481
,
173
(
2012
).
836.
M. S.
Senn
,
I.
Loa
,
J. P.
Wright
, and
J. P.
Attfield
,
Phys. Rev. B
85
,
125119
(
2012
).
837.
G.
Subías
,
J.
García
,
J.
Blasco
,
J.
Herrero-Martín
,
M. C.
Sánchez
,
J.
Orna
, and
L.
Morellón
,
J. Synchrotron Radiat.
19
,
159
(
2012
).
838.
F.
Walz
,
J. Phys.: Condens. Mater
14
,
R285
(
2002
).
839.
J.
García
and
G.
Subías
,
J. Phys.: Condens. Matter
16
,
R145
(
2004
).
840.
C. A. F.
Vaz
,
C. H.
Ahn
, and
V. E.
Henrich
, “
Spin-polarisation effects in oxide materials
,” in
Epitaxial Ferromagnetic Films and Spintronic Applications
(
Research Signpost
,
Kerala
,
2009
), p.
145
.
841.
E.
Pachoud
,
J.
Cumby
,
G.
Perversi
,
J. P.
Wright
, and
J. P.
Attfield
,
Nat. Commun.
11
,
1671
(
2020
).
842.
P.
Beaud
,
S. L.
Johnson
,
E.
Vorobeva
,
U.
Staub
,
R. A.
De Souza
,
C. J.
Milne
,
Q. X.
Jia
, and
G.
Ingold
,
Phys. Rev. Lett.
103
,
155702
(
2009
).
843.
T. P.
Ma
,
Y.
Yang
,
Z.
Ding
,
Z. H.
Chen
,
H. B.
Zhao
,
P.
Werner
,
S. S. P.
Parkin
, and
Y. Z.
Wu
,
Phys. Rev. B
95
,
014103
(
2017
).
844.
B.
Carter-Stiglitz
,
B.
Moskowitz
,
P.
Solheid
,
T. S.
Berquó
,
M.
Jackson
, and
A.
Kosterov
,
J. Geophys. Res.
111
,
B12S05
, (
2006
).
845.
M.
Fiebig
,
J. Phys. D: Appl. Phys.
38
,
R123
(
2005
).
846.
N. A.
Spaldin
and
R.
Ramesh
,
MRS Bull.
33
,
1047
(
2006
).
847.
K. F.
Wang
,
J. M.
Liu
, and
Z. F.
Ren
,
Adv. Phys.
58
,
321
(
2009
).
848.
C. A. F.
Vaz
,
J.
Hoffman
,
C. H.
Ahn
, and
R.
Ramesh
,
Adv. Mater.
22
,
2900
(
2010
).
849.
L. W.
Martin
,
Y.-H.
Chu
, and
R.
Ramesh
,
Mater. Sci. Eng. R
68
,
89
(
2010
).
850.
C. A. F.
Vaz
,
J. Phys.: Condens. Matter
24
,
333201
(
2012
).
851.
C. A. F.
Vaz
and
U.
Staub
,
J. Mater. Chem. C
1
,
6731
(
2013
).
852.
S.
Dong
,
J.-M.
Liu
,
S.-W.
Cheong
, and
Z.
Ren
,
Adv. Phys.
64
,
519
(
2015
).
853.
C.
Lu
,
W.
Hu
,
Y.
Tian
, and
T.
Wu
,
Appl. Phys. Rev.
2
,
021304
(
2015
).
854.
V.
Garcia
,
M.
Bibes
, and
A.
Barthélémy
,
C. R. Phys.
16
,
168
(
2015
).
855.
Z.
Chu
,
M.
PourhosseiniAsl
, and
S.
Dong
,
J. Phys. D: Appl. Phys.
51
,
243001
(
2018
).
856.
J. F.
Scott
,
J. Mater. Chem.
22
,
4567
(
2012
).
857.
S.
Fusil
,
V.
Garcia
,
A.
Barthélémy
, and
M.
Bibes
,
Annu. Rev. Mater. Res.
44
,
91
(
2014
).
858.
M.
Trassin
,
J. Phys.: Condens. Matter
28
,
033001
(
2016
).
859.
S.
Picozzi
,
K.
Yamauchi
,
I. A.
Sergienko
,
C.
Sen
,
B.
Sanyal
, and
E.
Dagotto
,
J. Phys.: Condens. Matter
20
,
434208
(
2008
).
860.
T.
Kimura
,
Annu. Rev. Condens. Mater. Phys.
3
,
93
(
2012
).
861.
K.
Yamauchi
and
P.
Barone
,
J. Phys.: Condens. Matter
26
,
103201
(
2014
).
862.
P. S.
Halasyamani
and
K. R.
Poeppelmeier
,
Chem. Mater.
10
,
2753
(
1998
).
863.
R.
Ramesh
and
N. A.
Spaldin
,
Nat. Mater.
6
,
21
(
2007
).
864.
R.
Haumont
,
J.
Kreisel
, and
P.
Bouvier
,
Phase Transitions
79
,
1043
(
2006
).
865.
J. F.
Scott
,
J. Magn. Magn. Mater.
321
,
1689
(
2009
).
866.
A. N.
Kalinkin
and
V. M.
Skorikov
,
Russ. J. Inorg. Chem.
55
,
1794
(
2010
).
867.
F.
Matsukura
,
Y.
Tokura
, and
H.
Ohno
,
Nat. Nanotechnol.
10
,
209
(
2015
).
868.
C.
Song
,
B.
Cui
,
F.
Li
,
X.
Zhou
, and
F.
Pan
,
Prog. Mater. Sci.
87
,
33
(
2017
).
869.
M.
Zhu
,
T.
Nan
,
B.
Peng
,
Y.
Zhang
,
Z.
Zhou
,
X.
Yang
,
W.
Ren
,
N. X.
Sun
, and
M.
Liu
,
IEEE Trans. Magn.
53
,
0800116
(
2017
).
870.
C.-W.
Nan
,
M. I.
Bichurin
,
S.
Dong
,
D.
Viehland
, and
G.
Srinivasan
,
J. Appl. Phys.
103
,
031101
(
2008
).
871.
G.
Srinivasan
,
Annu. Rev. Mater. Res.
40
,
153
(
2010
).
872.
T.
Taniyama
,
J. Phys.: Condens. Matter
27
,
504001
(
2015
).
873.
A.
Chen
,
Q.
Su
,
H.
Han
,
E.
Enriquez
, and
Q.
Jia
,
Adv. Mater.
31
,
1803241
(
2019
).
874.
P.
Borisov
,
A.
Hochstrat
,
X.
Chen
,
W.
Kleemann
, and
C.
Binek
,
Phys. Rev. Lett.
94
,
117203
(
2005
).
875.
J.
Dho
and
M. G.
Blamire
,
Appl. Phys. Lett.
87
,
252504
(
2005
).
876.
L. W.
Martin
,
Y.-H.
Chu
,
M. B.
Holcomb
,
M.
Huijben
,
P.
Yu
,
S.-J.
Han
,
D.
Lee
,
S. X.
Wang
, and
R.
Ramesh
,
Nano Lett.
8
,
2050
(
2008
).
877.
Q.
Yang
,
Z.
Zhou
,
N. X.
Sun
, and
M.
Liu
,
Phys. Lett. A
381
,
1213
(
2017
).
878.
J.
Stöhr
and
H. C.
Siegmann
,
Magnetism
(
Springer-Verlag
,
Berlin
,
2006
).
879.
880.
C.-G.
Duan
,
J. P.
Velev
,
R. F.
Sabirianov
,
Z.
Zhu
,
J.
Chu
,
S. S.
Jaswal
, and
E. Y.
Tsymbal
,
Phys. Rev. Lett.
101
,
137201
(
2008
).
881.
J. M.
Rondinelli
,
M.
Stengel
, and
N. A.
Spaldin
,
Nat. Nanotechnol.
3
,
46
(
2008
).
882.
M. K.
Niranjan
,
C.-G.
Duan
,
S. S.
Jaswal
, and
E. Y.
Tsymbal
,
Appl. Phys. Lett.
96
,
222504
(
2010
).
883.
K.
Nakamura
,
R.
Shimabukuro
,
Y.
Fujiwara
,
T.
Akiyama
,
T.
Ito
, and
A. J.
Freeman
,
Phys. Rev. Lett.
102
,
187201
(
2009
).
884.
T.
Maruyama
,
Y.
Shiota
,
T.
Nozaki
,
K.
Ohta
,
N.
Toda
,
M.
Mizuguchi
,
A. A.
Tulapurkar
,
T.
Shinjo
,
M.
Shiraishi
,
S.
Mizukami
,
Y.
Ando
, and
Y.
Suzuki
,
Nat. Nanotechnol.
4
,
158
(
2009
).
885.
M.
Tsujikawa
and
T.
Oda
,
Phys. Rev. Lett.
102
,
247203
(
2009
).
886.
K.
Nakamura
,
T.
Akiyama
,
T.
Ito
,
M.
Weinert
, and
A. J.
Freeman
,
Phys. Rev. B
81
,
220409
(
2010
).
887.
R.
Shimabukuro
,
K.
Nakamura
,
T.
Akiyama
, and
T.
Ito
,
Physica E
42
,
1014
(
2010
).
888.
S.
Haraguchi
,
M.
Tsujikawa
,
J.
Gotou
, and
T.
Oda
,
J. Phys. D: Appl. Phys.
44
,
064005
(
2011
).
889.
B.
Heinrich
and
J. F.
Cochran
,
Adv. Phys.
42
,
523
(
1993
).
890.
M. T.
Johnson
,
P. J. H.
Bloemen
,
F. J. A.
der Broeder
, and
J. J.
de Vries
,
Rep. Prog. Phys.
59
,
1409
(
1996
).
891.
D.
Sander
,
J. Phys.: Condens. Matter
16
,
R603
(
2004
).
892.
C. A. F.
Vaz
,
J. A. C.
Bland
, and
G.
Lauhoff
,
Rep. Prog. Phys.
71
,
056501
(
2008
).
893.
T.
Seki
,
M.
Kohda
,
J.
Nitta
, and
K.
Takanashi
,
Appl. Phys. Lett.
98
,
212505
(
2011
).
894.
M.
Endo
,
S.
Kanai
,
S.
Ikeda
,
F.
Matsukura
, and
H.
Ohno
,
Appl. Phys. Lett.
96
,
212503
(
2010
).
895.
C.
Fowley
,
N.
Decorde
,
K.
Oguz
,
K.
Rode
,
H.
Kurt
, and
J. M. D.
Coey
,
IEEE Trans. Magn.
46
,
2116
(
2010
).
896.
C.
Fowley
,
K.
Rode
,
K.
Oguz
,
H.
Kurt
, and
J. M. D.
Coey
,
J. Phys. D: Appl. Phys.
44
,
305001
(
2011
).
897.
A.
Mardana
,
S.
Ducharme
, and
S.
Adenwalla
,
Nano Lett.
11
,
3862
(
2011
).
898.
J.
Heidler
,
M.
Fechner
,
R. V.
Chopdekar
,
C.
Piamonteze
,
J.
Dreiser
,
C. A.
Jenkins
,
E.
Arenholz
,
S.
Rusponi
,
H.
Brune
,
N. A.
Spaldin
, and
F.
Nolting
,
Phys. Rev. B
94
,
014401
(
2016
).
899.
L.
Shu
,
Z.
Li
,
J.
Ma
,
Y.
Gao
,
L.
Gu
,
Y.
Shen
,
Y.
Lin
, and
C. W.
Nan
,
Appl. Phys. Lett.
100
,
022405
(
2012
).
900.
F.
Bonell
,
S.
Murakami
,
Y.
Shiota
,
T.
Nozaki
,
T.
Shinjo
, and
Y.
Suzuki
,
Appl. Phys. Lett.
98
,
232510
(
2011
).
901.
Y.
Shiota
,
T.
Maruyama
,
T.
Nozaki
,
T.
Shinjo
,
M.
Shiraishi
, and
Y.
Suzuki
,
Appl. Phys. Express
2
,
063001
(
2009
).
902.
S.-S.
Ha
,
N.-H.
Kim
,
S.
Lee
,
C.-Y.
You
,
Y.
Shiota
,
T.
Maruyama
,
T.
Nozaki
, and
Y.
Suzuki
,
Appl. Phys. Lett.
96
,
142512
(
2010
).
903.
K. H.
He
,
J. S.
Chen
, and
Y. P.
Feng
,
Appl. Phys. Lett.
99
,
072503
(
2011
).
904.
T.
Nozaki
,
Y.
Shiota
,
M.
Shiraishi
,
T.
Shinjo
, and
Y.
Suzuki
,
Appl. Phys. Lett.
96
,
022506
(
2010
).
905.
Y.
Shiota
,
S.
Murakami
,
F.
Bonell
,
T.
Nozaki
,
T.
Shinjo
, and
Y.
Suzuki
,
Appl. Phys. Express
4
,
043005
(
2011
).
906.
W.-G.
Wang
,
M.
Li
,
S.
Hageman
, and
C. L.
Chien
,
Nat. Mater.
11
,
64
(
2012
).
907.
H.
Meng
,
R.
Sbiaa
,
M. A. K.
Akhtar
,
R. S.
Liu
,
V. B.
Naik
, and
C. C.
Wang
,
Appl. Phys. Lett.
100
,
122405
(
2012
).
908.
T.
Zhou
,
S. H.
Leong
,
Z. M.
Yuan
,
S. B.
Hu
,
C. L.
Ong
, and
B.
Liu
,
Appl. Phys. Lett.
96
,
012506
(
2010
).
909.
C.-G.
Duan
,
S. S.
Jaswal
, and
E. Y.
Tsymbal
,
Phys. Rev. Lett.
97
,
047201
(
2006
).
910.
M.
Fechner
,
I. V.
Maznichenko
,
S.
Ostanin
,
A.
Ernst
,
J.
Henk
,
P.
Bruno
, and
I.
Mertig
,
Phys. Rev. B
78
,
212406
(
2008
).
911.
H. L.
Meyerheim
,
F.
Klimenta
,
A.
Ernst
,
K.
Mohseni
,
S.
Ostanin
,
M.
Fechner
,
S.
Parihar
,
I. V.
Maznichenko
,
I.
Mertig
, and
J.
Kirschner
,
Phys. Rev. Lett.
106
,
087203
(
2011
).
912.
J.-Q.
Dai
,
Y.-M.
Song
, and
H.
Zhang
,
J. Appl. Phys.
111
,
114301
(
2012
).
913.
L.
Bocher
,
A.
Gloter
,
A.
Crassous
,
V.
Garcia
,
K.
March
,
A.
Zobelli
,
S.
Valencia
,
S.
Enouz-Vedrenne
,
X.
Moya
,
N. D.
Marthur
,
C.
Deranlot
,
S.
Fusil
,
K.
Bouzehouane
,
M.
Bibes
,
A.
Barthélémy
,
C.
Colliex
, and
O.
Stéphan
,
Nano Lett.
12
,
376
(
2012
).
914.
S.
Borek
,
I.
Maznichenko
,
G.
Fischer
,
W.
Hergert
,
I.
Mertig
,
A.
Ernst
,
S.
Ostanin
, and
A.
Chassé
,
Phys. Rev. B
85
,
134432
(
2012
).
915.
R.
Arras
and
S.
Cherifi-Hertel
,
ACS Appl. Mater. Interfaces
11
,
34399
(
2019
).
916.
S.
Valencia
,
A.
Crassous
,
L.
Bocher
,
V.
Garcia
,
X.
Moya
,
R. O.
Cherifi
,
C.
Deranlot
,
K.
Bouzehouane
,
S.
Fusil
,
A.
Zobelli
,
A.
Gloter
,
N. D.
Mathur
,
A.
Gaupp
,
R.
Abrudan
,
F.
Radu
,
A.
Barthélémy
, and
M.
Bibes
,
Nat. Mater.
10
,
753
(
2011
).
917.
Y.
Liu
,
J.
Tornos
,
S. G. E.
te Velthuis
,
J. W.
Freeland
,
H.
Zhou
,
P.
Steadman
,
P.
Bencok
,
C.
Leon
, and
J.
Santamaria
,
APL Mater.
4
,
046105
(
2016
).
918.
O.
Vlašín
,
R.
Jarrier
,
R.
Arras
,
L.
Calmels
,
B.
Warot-Fonrose
,
C.
Marcelot
,
M.
Jamet
,
P.
Ohresser
,
F.
Scheurer
,
R.
Hertel
,
G.
Herranz
, and
S.
Cherifi-Hertel
,
ACS Appl. Mater. Interfaces
8
,
7553
(
2016
).
919.
P. W.
Anderson
, “
Exchange in insulators: Superexchange, direct exchange, and double exchange
,” in
Magnetism
, edited by
G. T.
Rado
and
H.
Suhl
(
Academic Press
,
New York
,
1963
), Vol.
I
, p.
25
.
920.
P. W.
Anderson
and
H.
Hasegawa
,
Phys. Rev.
100
,
675
(
1955
).
921.
P.-G.
de Gennes
,
Phys. Rev.
118
,
141
(
1960
).
922.
K.
Kubo
and
N.
Ohata
,
J. Phys. Soc. Jpn.
33
,
21
(
1972
).
923.
924.
T. R.
McGuire
and
E. L.
Boyd
, in
NiFe3+ spinels (Ni ferrites) and NiFe3+ spinels with substitutions
, Volume 4b of Landolt-Börnstein - Group III: Crystal and Solid State Physics (
Springer-Verlag
,
Berlin
,
1970
), p.
75
.
925.
J. B.
Goodenough
and
J.-S.
Zhou
,
Chem. Mater.
10
,
2980
(
1998
).
926.
Y. A.
Izyumov
and
E. Z.
Kurmaev
,
Phys.-Usp.
51
,
23
(
2008
).
927.
E.
Dagotto
and
Y.
Tokura
,
MRS Bull.
33
,
1037
(
2008
).
928.
J. B.
Goodenough
,
Prog. Solid State Chem.
5
,
145
(
1971
).
929.
M.
Mochizuki
and
M.
Imada
,
New J. Phys.
6
,
154
(
2004
).
930.
M.
Johnsson
and
P.
Lemmens
, “
Crystallography and chemistry of perovskites
,” in
Handbook of Magnetism and Advanced Magnetic Materials
, edited by
H.
Kronmüller
and
S.
Parkin
(
John Wiley & Sons
,
2007
), Vol.
4
,
p
.
1
.
931.
M.
Johnsson
and
P.
Lemmens
,
J. Phys.: Condens. Matter
20
,
264001
(
2008
).
932.
I.
Solovyev
and
K.
Terakura
, “
Orbital degeneracy and magnetism of perovskite manganese oxides
,” in
Electronic Structure and Magnetism of Complex Materials
, edited by
D. J.
Singh
and
D. A.
Papaconstantopoulos
(
Springer-Verlag
,
Berlin
,
2003
), p.
253
.
933.
A.-M.
Haghiri-Gosnet
and
J.-P.
Renard
,
J. Phys. D: Appl. Phys.
36
,
R127
(
2003
).
934.
G. H.
Jonker
and
J. H.
van Santen
,
Physica
16
,
337
(
1950
).
935.
B.
Roldan Cuenya
,
M.
Doi
,
W.
Keune
,
S.
Hoch
,
D.
Reuter
,
A.
Wieck
,
T.
Schmitte
, and
H.
Zabel
,
Appl. Phys. Lett.
82
,
1072
(
2003
).
936.
R. P.
Borges
,
W.
Guichard
,
J. G.
Lunney
,
J. M. D.
Coey
, and
F.
Ott
,
J. Appl. Phys.
89
,
3868
(
2001
).
937.
M.
Izumi
,
Y.
Ogimoto
,
T.
Manako
,
M.
Kawasaki
, and
Y.
Tokura
,
J. Phys. Soc. Jpn.
71
,
2621
(
2002
).
938.
J. J.
Kavich
,
M. P.
Warusawithana
,
J. W.
Freeland
,
P.
Ryan
,
X.
Zhai
,
R. H.
Kodama
, and
J. N.
Eckstein
,
Phys. Rev. B
76
,
014410
(
2007
).
939.
J. W.
Freeland
,
J. J.
Kavich
,
K. E.
Gray
,
L.
Ozyuzer
,
H.
Zheng
,
J. F.
Mitchell
,
M. P.
Warusawithana
,
P.
Ryan
,
X.
Zhai
,
R. H.
Kodama
, and
J. N.
Eckstein
,
J. Phys.: Condens. Matter
19
,
315210
(
2007
).
940.
M.
Huijben
,
L. W.
Martin
,
Y.-H.
Chu
,
M. B.
Holcomb
,
P.
Yu
,
G.
Rijnders
,
D. H. A.
Blank
, and
R.
Ramesh
,
Phys. Rev. B
78
,
094413
(
2008
).
941.
L. F.
Kourkoutis
,
J. H.
Song
,
H. Y.
Hwang
, and
D. A.
Muller
,
Proc. Natl. Acad. Sci.
107
,
11682
(
2010
).
942.
J.-S.
Lee
,
D. A.
Arena
,
P.
Yu
,
C. S.
Nelson
,
R.
Fan
,
C. J.
Kinane
,
S.
Langridge
,
M. D.
Rossell
,
R.
Ramesh
, and
C.-C.
Kao
,
Phys. Rev. Lett.
105
,
257204
(
2010
).
943.
H.
Boschker
,
J.
Kautz
,
E. P.
Houwman
,
W.
Siemons
,
D. H. A.
Blank
,
M.
Huijben
,
G.
Koster
,
A.
Vailionis
, and
G.
Rijnders
,
Phys. Rev. Lett.
109
,
157207
(
2012
).
944.
L.
Chen
,
Z.
Wang
,
G.
Wang
,
H.
Guo
,
M.
Saghayezhian
,
Z.
Liao
,
Y.
Zhu
,
E. W.
Plummer
, and
J.
Zhang
,
Phys. Rev. Mater.
3
,
044407
(
2019
).
945.
C. A. F.
Vaz
,
J.
Hoffman
,
Y.
Segal
,
M. S. J.
Marshall
,
J. W.
Reiner
,
Z.
Zhang
,
R. D.
Grober
,
F. J.
Walker
, and
C. H.
Ahn
,
J. Appl. Phys.
109
,
07D905
(
2011
).
946.
T.
Kanki
,
H.
Tanaka
, and
T.
Kawai
,
Appl. Phys. Lett.
89
,
242506
(
2006
).
947.
Z.
Fang
,
I. V.
Solovyev
, and
K.
Terakura
,
Phys. Rev. Lett.
84
,
3169
(
2000
).
948.
J. D.
Burton
and
E. Y.
Tsymbal
,
Phys. Rev. B
80
,
174406
(
2009
).
949.
S.
Dong
,
X.
Zhang
,
R.
Yu
,
J.-M.
Liu
, and
E.
Dagotto
,
Phys. Rev. B
84
,
155117
(
2011
).
950.
N. C.
Bristowe
,
M.
Stegel
,
P. B.
Littlewood
,
J. M.
Pruneda
, and
E.
Artacho
,
Phys. Rev. B
85
,
024106
(
2012
).
951.
H.
Chen
and
S.
Ismael-Beigi
,
Phys. Rev. B
86
,
024433
(
2012
).
952.
S.
Dong
and
E.
Dagotto
,
Phys. Rev. B
88
,
140404(R)
(
2013
).
953.
M.
Hammouri
,
E.
Fohtung
, and
I.
Vasiliev
,
J. Phys.: Condens. Matter
28
,
396004
(
2016
).
954.
C. A. F.
Vaz
,
Y.
Segal
,
J.
Hoffman
,
R. D.
Grober
,
F. J.
Walker
, and
C. H.
Ahn
,
Appl. Phys. Lett.
97
,
042506
(
2010
).
955.
C. A. F.
Vaz
,
J.
Hoffman
,
Y.
Segal
,
F. J.
Walker
, and
C. H.
Ahn
,
Proc. SPIE
7760
,
776013
(
2010
).
956.
H.
Lu
,
T. A.
George
,
Y.
Wang
,
I.
Ketsman
,
J. D.
Burton
,
C.-W.
Bark
,
S.
Ryu
,
D. J.
Kim
,
J.
Wang
,
C.
Binek
,
P. A.
Dowben
,
A.
Sokolov
,
C.-B.
Eom
,
E. Y.
Tsymbal
, and
A.
Gruverman
,
Appl. Phys. Lett.
100
,
232904
(
2012
).
957.
A. M.
Kane
,
I.-T.
Chiu
,
N. J.
Ahlm
,
R. V.
Chopdekar
,
A. T.
N'Diaye
,
E.
Arenholz
,
A.
Mehta
,
V.
Lauter
, and
Y.
Takamura
,
ACS Appl. Mater. Interfaces
12
,
45437
(
2020
).
958.
P. M.
Leufke
,
R.
Kruk
,
R. A.
Brand
, and
H.
Hahn
,
Phys. Rev. B
87
,
094416
(
2013
).
959.
D.
Preziosi
,
I.
Fina
,
E.
Pippel
,
D.
Hesse
,
X.
Marti
,
F.
Bern
,
M.
Ziese
, and
M.
Alexe
,
Phys. Rev. B
90
,
125155
(
2014
).
960.
T. L.
Meyer
,
A.
Herklotz
,
V.
Lauter
,
J. W.
Freeland
,
J.
Nichols
,
E.-J.
Guo
,
S.
Lee
,
T. Z.
Ward
,
N.
Balke
,
S. V.
Kalinin
,
M. R.
Fitzsimmons
, and
H. N.
Lee
,
Phys. Rev. B
94
,
174432
(
2016
).
961.
J.
Ma
,
Y.
Tian
,
M.
Chen
,
J.
Wang
,
J.
Ma
,
J.
Zhang
, and
C.-W.
Nan
,
J. Appl. Phys.
126
,
075301
(
2019
).
962.
D.
Yi
,
P.
Yu
,
Y.-C.
Chen
,
H.-H.
Lee
,
Q.
He
,
Y.-H.
Chu
, and
R.
Ramesh
,
Adv. Mater.
31
,
1806335
(
2019
).
963.
B.
Paudel
,
I.
Vasiliev
,
M.
Hammouri
,
D.
Karpov
,
A.
Chen
,
V.
Lauter
, and
E.
Fohtung
,
RSC Adv.
9
,
13033
(
2019
).
964.
C.
Liu
,
Y.
Liu
,
B.
Zhang
,
C.-J.
Sun
,
D.
Lan
,
P.
Chen
,
X.
Wu
,
P.
Yang
,
X.
Yu
,
T.
Charlton
,
M. R.
Fitzsimmons
,
J.
Ding
,
J.
Chen
, and
G. M.
Chow
,
ACS Appl. Mater. Interfaces
13
,
30137
(
2021
).
965.
X.
Ma
,
A.
Kumar
,
S.
Dussan
,
H.
Zhai
,
F.
Fang
,
H. B.
Zhao
,
J. F.
Scott
,
R. S.
Katiyar
, and
G.
Lüpke
,
Appl. Phys. Lett.
104
,
132905
(
2014
).
966.
J. F.
Ankner
and
G. P.
Felcher
,
J. Magn. Magn. Mater.
200
,
741
(
1999
).
967.
968.
C.
Fermon
,
F.
Ott
, and
A.
Menelle
, “
Neutron reflectometry
,” in
X-Ray and Neutron Reflectivity: Principles and Applications
, edited by
J.
Daillant
and
A.
Gibaud
(
Springer
,
Berlin
,
1999
), p.
163
.
969.
J. A. C.
Bland
and
C. A. F.
Vaz
, “
Polarised neutron reflection studies of thin magnetic films
,” in
Ultrathin Magnetic Structures III
, edited by
J. A. C.
Bland
and
B.
Heinrich
(
Springer-Verlag
,
Berlin
,
2005
), p.
233
.
970.
H.
Zabel
,
K.
Theis-Bröhl
, and
B. P.
Toperverg
, “
Polarized neutron reflectivity and scattering of magnetic nanostructures and spintronic materials
,” in
Handbook of Magnetism and Advanced Magnetic Materials
, edited by
H.
Kronmüller
and
S.
Parkin
(
John Wiley & Sons
,
Hoboken
,
2007
), p.
1237
.
971.
M.
Varela
,
M. P.
Oxley
,
W.
Luo
,
J.
Tao
,
M.
Watanabe
,
A. R.
Lupini
,
S. T.
Pantelides
, and
S. J.
Pennycook
,
Phys. Rev. B
79
,
085117
(
2009
).
972.
W.
Huang
,
Y.
Yin
, and
X.
Li
,
Appl. Phys. Rev.
5
,
041110
(
2018
).
973.
A. Y.
Borisevich
,
H. J.
Chang
,
M.
Huijben
,
M. P.
Oxley
,
S.
Okamoto
,
M. K.
Niranjan
,
J. D.
Burton
,
E. Y.
Tsymbal
,
Y. H.
Chu
,
P.
Yu
,
R.
Ramesh
,
S. V.
Kalinin
, and
S. J.
Pennycook
,
Phys. Rev. Lett.
105
,
087204
(
2010
).
974.
S.
Zhang
,
X.
Guo
,
Y.
Tang
,
D.
Ma
,
Y.
Zhu
,
Y.
Wang
,
S.
Li
,
M.
Han
,
D.
Chen
,
J.
Ma
,
B.
Wu
, and
X.
Ma
,
ACS Nano
12
,
3681
(
2018
).
975.
F. Y.
Bruno
,
S.
Boyn
,
S.
Fusil
,
S.
Girod
,
C.
Carrétéro
,
M.
Marinova
,
A.
Gloter
,
S.
Xavier
,
C.
Deranlot
,
M.
Bibes
,
A.
Barthélémy
, and
V.
Garcia
,
Adv. Electron. Mater.
2
,
1500245
(
2016
).
976.
R. A.
Lefever
, in Fe3+ Spinels, Volume 12b of Landolt-Börnstein—Group III: Crystal and Solid State Physics (
Springer-Verlag
,
Berlin
,
1980
).
977.
H.
Ji
,
Y. G.
Wang
, and
Y.
Li
,
J. Magn. Magn. Mater.
442
,
242
(
2017
).
978.
H.
Ji
,
Y. G.
Wang
, and
Y.
Li
,
J. Phys.: Condens. Matter
30
,
11LT01
(
2018
).
979.
H.
Wu
,
S. D.
Zhou
,
Y.
Li
,
Y. G.
Wang
, and
F. M.
Pan
,
Appl. Phys. Lett.
113
,
102404
(
2018
).
980.
Y. L.
Lu
,
H.
Wu
,
L.
Zhu
,
Y.
Li
,
Y. G.
Wang
, and
F. M.
Pan
,
Appl. Phys. Lett.
114
,
052401
(
2019
).
981.
C.
Moutafis
,
S.
Komineas
,
C. A. F.
Vaz
,
J. A. C.
Bland
,
T.
Shima
,
T.
Seki
, and
K.
Takanashi
,
Phys. Rev. B
76
,
104426
(
2007
).
982.
F.
Büttner
,
C.
Moutafis
,
M.
Schneider
,
B.
Krüger
,
C. M.
Günther
,
J.
Geilhufe
,
C. v
Korff Schmising
,
J.
Mohanty
,
B.
Pfau
,
S.
Schaffert
,
A.
Bisig
,
M.
Foerster
,
T.
Schulz
,
C. A. F.
Vaz
,
J. H.
Franken
,
H. J. M.
Swagten
,
M.
Kläui
, and
S.
Eisebitt
,
Nat. Phys.
11
,
225
(
2015
).
983.
984.
S.
Mühlbauer
,
B.
Binz
,
F.
Jonietz
,
C.
Pfleiderer
,
A.
Rosch
,
A.
Neubauer
,
R.
Georgii
, and
P.
Böni
,
Science
323
,
915
(
2009
).
985.
X. Z.
Yu
,
Y.
Onose
,
N.
Kanazawa
,
J. H.
Park
,
J. H.
Han
,
Y.
Matsui
,
N.
Nagaosa
, and
Y.
Tokura
,
Nature
465
,
901
(
2010
).
986.
S.
Heinze
,
K.
von Bergmann
,
M.
Menzel
,
J.
Brede
,
A.
Kubetzka
,
R.
Wiesendanger
,
G.
Bihlmayer
, and
S.
Blügel
,
Nat. Phys.
7
,
713
(
2011
).
987.
J.
Sampaio
,
V.
Cros
,
S.
Rohart
,
A.
Thiaville
, and
A.
Fert
,
Nat. Nanotechnol.
8
,
839
(
2013
).
988.
A.
Hrabec
,
N. A.
Porter
,
A.
Wells
,
M. J.
Benitez
,
G.
Burnell
,
S.
McVitie
,
D.
McGrouther
,
T. A.
Moore
, and
C. H.
Marrows
,
Phys. Rev. B
90
,
020402
(
2014
).
989.
C.
Moreau-Luchaire
,
C.
Moutafis
,
N.
Reyren
,
J.
Sampaio
,
C. A. F.
Vaz
,
N.
Van Horne
,
K.
Bouzehouane
,
K.
Garcia
,
C.
Deranlot
,
P.
Warnicke
,
P.
Wohlhueter
,
J.-M.
George
,
M.
Weigand
,
J.
Raabe
,
V.
Cros
, and
A.
Fert
,
Nat. Nanotechnol.
11
,
444
(
2016
).
990.
W.
Jiang
,
G.
Chen
,
K.
Liu
,
J.
Zang
,
S. G. E.
te Velthuis
, and
A.
Hoffmann
,
Phys. Rep.
704
,
1
(
2017
).
991.
A.
Soumyanarayanan
,
M.
Raju
,
A. L. G.
Oyarce
,
A. K. C.
Tan
,
M. Y.
Im
,
A. P.
Petrovic
,
P.
Ho
,
K. H.
Khoo
,
M.
Tran
,
C. K.
Gan
,
F.
Ernult
, and
C.
Panagopoulos
,
Nat. Mater.
16
,
898
(
2017
).
992.
K. Y.
Meng
,
A. S.
Ahmed
,
M.
Bacani
,
A. O.
Mandru
,
X.
Zhao
,
N.
Bagues
,
B. D.
Esser
,
J.
Flores
,
D. W.
McComb
,
H. J.
Hug
, and
F.
Yang
,
Nano Lett.
19
,
3169
(
2019
).
993.
J.
Matsuno
,
N.
Ogawa
,
K.
Yasuda
,
F.
Kagawa
,
W.
Koshibae
,
N.
Nagaosa
,
Y.
Tokura
, and
M.
Kawasaki
,
Sci. Adv.
2
,
e1600304
(
2016
).
994.
Q.
Qin
,
L.
Liu
,
W.
Lin
,
X.
Shu
,
Q.
Xie
,
Z.
Lim
,
C.
Li
,
S.
He
,
G. M.
Chow
, and
J.
Chen
,
Adv. Mater.
31
,
e1807008
(
2019
).
995.
M.
Ziese
,
L.
Jin
, and
I.
Lindfors-Vrejoiu
,
J. Phys.: Mater.
2
,
34008
(
2019
).
996.
L. F.
Mattheiss
,
Phys. Rev. B
13
,
2433
(
1976
).
997.
L.
Wang
,
Q.
Feng
,
Y.
Kim
,
R.
Kim
,
K. H.
Lee
,
S. D.
Pollard
,
Y. J.
Shin
,
H.
Zhou
,
W.
Peng
,
D.
Lee
,
W.
Meng
,
H.
Yang
,
J. H.
Han
,
M.
Kim
,
Q.
Lu
, and
T. W.
Noh
,
Nat. Mater.
17
,
1087
(
2018
).
998.
Y.
Gu
,
C.
Song
,
Q.
Zhang
,
F.
Li
,
H.
Tan
,
K.
Xu
,
J.
Li
,
M. S.
Saleem
,
M. U.
Fayaz
,
J.
Peng
,
F.
Hu
,
L.
Gu
,
W.
Liu
,
Z.
Zhang
, and
F.
Pan
,
ACS Appl. Mater. Interfaces
12
(
5
),
6707
(
2020
).
999.
T.
Dietl
,
H.
Ohno
,
F.
Matsukura
,
J.
Cibert
, and
D.
Ferrand
,
Science
287
,
1019
(
2000
).
1000.
T.
Dietl
,
H.
Ohno
, and
F.
Matsukura
,
Phys. Rev. B
63
,
195205
(
2001
).
1001.
T.
Dietl
,
Semicond. Sci. Technol.
17
,
377
(
2002
).
1002.
P.
Mahadevan
,
A.
Zunger
, and
D. D.
Sarma
,
Phys. Rev. Lett.
93
,
177201
(
2004
).
1003.
T.
Jungwirth
,
J.
Sinova
,
J.
Mašek
,
J.
Kučera
, and
A. H.
MacDonald
,
Rev. Mod. Phys.
78
,
809
(
2006
).
1004.
1005.
A.
Zunger
,
S.
Lany
, and
H.
Raebiger
,
Physics
3
,
53
(
2010
).
1006.
K.
Sato
,
L.
Bergqvist
,
J.
Kudrnovský
,
P. H.
Dederichs
,
O.
Eriksson
,
I.
Turek
,
B.
Sanyal
,
G.
Bouzerar
,
H.
Katayama-Yoshida
,
V. A.
Dinh
,
T.
Fukushima
,
H.
Kizaki
, and
R.
Zeller
,
Rev. Mod. Phys.
82
,
1633
(
2010
).
1007.
H.
Ohno
,
D.
Chiba
,
F.
Matsukura
,
T.
Omiya
,
E.
Abe
,
T.
Dietl
,
Y.
Ohno
, and
K.
Ohtani
,
Nature
408
,
944
(
2000
).
1008.
D.
Chiba
,
M.
Yamanouchi
,
F.
Matsukura
, and
H.
Ohno
,
Science
301
,
943
(
2003
).
1009.
Y. D.
Park
,
A. T.
Hanbicki
,
S. C.
Erwin
,
C. S.
Hellberg
,
J. M.
Sullivan
,
J. E.
Mattson
,
T. F.
Ambrose
,
A.
Wilson
,
G.
Spanos
, and
B. T.
Jonker
,
Science
295
,
651
(
2002
).
1010.
J.
Chen
,
K. L.
Wang
, and
K.
Galatsis
,
Appl. Phys. Lett.
90
,
012501
(
2007
).
1011.
F.
Xiu
,
Y.
Wang
,
J.
Kim
,
A.
Hong
,
J.
Tang
,
A. P.
Jacob
,
J.
Zou
, and
K. L.
Wang
,
Nat. Mater.
9
,
337
(
2010
).
1012.
M. K.
Kneip
,
D. R.
Yakovlev
,
M.
Bayer
,
T.
Slobodskyy
,
G.
Schmidt
, and
L. W.
Molenkamp
,
Appl. Phys. Lett.
88
,
212105
(
2006
).
1013.
A. M.
Nazmul
,
S.
Kobayashi
,
S.
Sugahara
, and
M.
Tanaka
,
Physica E
21
,
937
(
2004
).
1014.
I.
Stolichnov
,
S. W. E.
Riester
,
H. J.
Trodahl
,
N.
Setter
,
A. W.
Rushforth
,
K. W.
Edmonds
,
R. P.
Campion
,
C. T.
Foxon
,
B. L.
Gallagher
, and
T.
Jungwirth
,
Nat. Mater.
7
,
464
(
2008
).
1015.
D.
Chiba
,
M.
Sawicki
,
Y.
Nishitani
,
Y.
Nakatani
,
F.
Matsukura
, and
H.
Ohno
,
Nature
455
,
515
(
2008
).
1016.
M. H. S.
Owen
,
J.
Wunderlich
,
V.
Novák
,
K.
Olejník
,
J.
Zemen
,
K.
Výborný
,
S.
Ogawa
,
A. C.
Irvine
,
A. J.
Ferguson
,
H.
Sirringhaus
, and
T.
Jungwirth
,
New J. Phys.
11
,
023008
(
2009
).
1017.
S. W. E.
Riester
,
I.
Stolichnov
,
H. J.
Trodahl
,
N.
Setter
,
A. W.
Rushforth
,
K. W.
Edmonds
,
R. P.
Campion
,
C. T.
Foxon
,
B. L.
Gallagher
, and
T.
Jungwirth
,
Appl. Phys. Lett.
94
,
063504
(
2009
).
1018.
M.
Endo
,
D.
Chiba
,
H.
Shimotani
,
F.
Matsukura
,
Y.
Iwasa
, and
H.
Ohno
,
Appl. Phys. Lett.
96
,
022515
(
2010
).
1019.
I.
Stolichnov
,
S. W. E.
Riester
,
E.
Mikheev
,
N.
Setter
,
A. W.
Rushforth
,
K. W.
Edmonds
,
R. P.
Campion
,
C. T.
Foxon
,
B. L.
Gallagher
,
T.
Jungwirth
, and
H. J.
Trodahl
,
Phys. Rev. B
83
,
115203
(
2011
).
1020.
T.
Zhao
,
S. R.
Shinde
,
S. B.
Ogale
,
H.
Zheng
,
T.
Venkatesan
,
R.
Ramesh
, and
S. D.
Sarma
,
Phys. Rev. Lett.
94
,
126601
(
2005
).
1021.
A.
Kaminski
and
S. D.
Sarma
,
Phys. Rev. Lett.
88
,
247202
(
2002
).
1022.
S.
Das Sarma
,
E. H.
Hwang
, and
A.
Kaminski
,
Phys. Rev. B
67
,
155201
(
2003
).
1023.
A.
Kaminski
and
S. D.
Sarma
,
Phys. Rev. B
68
,
235210
(
2003
).
1024.
J. M. D.
Coey
,
M.
Venkatesan
, and
C. B.
Fitzgerald
,
Nat. Mater.
4
,
173
(
2005
).
1025.
L.
Esaki
,
R. B.
Laibowitz
, and
P. J.
Stiles
,
IBM. Tech. Disclosure Bull.
13
,
2161
(
1971
).
1026.
I.
Giaever
,
Phys. Rev. Lett.
5
,
147
(
1960
).
1027.
I.
Giaever
,
Phys. Rev. Lett.
5
,
464
(
1960
).
1028.
J.
Moodera
,
L.
Kinder
,
T.
Wong
, and
R.
Meservey
,
Phys. Rev. Lett.
74
,
3273
(
1995
).
1029.
C.
Lichtensteiger
,
J.-M.
Triscone
,
J.
Junquera
, and
P.
Ghosez
,
Phys. Rev. Lett.
94
,
047603
(
2005
).
1030.
D. A.
Tenne
,
A.
Bruchhausen
,
N. D.
Lanzillotti-Kimura
,
A.
Fainstein
,
R. S.
Katiyar
,
A.
Cantarero
,
A.
Soukiassian
,
V.
Vaithyanathan
,
J. H.
Haeni
,
W.
Tian
,
D. G.
Schlom
,
K. J.
Choi
,
D. M.
Kim
,
C. B.
Eom
,
H. P.
Sun
,
X. Q.
Pan
,
Y. L.
Li
,
L. Q.
Chen
,
Q. X.
Jia
,
S. M.
Nakhmanson
,
K. M.
Rabe
, and
X. X.
Xi
,
Science
313
,
1614
(
2006
).
1031.
M. Y.
Zhuravlev
,
R. F.
Sabirianov
,
S. S.
Jaswal
, and
E. Y.
Tsymbal
,
Phys. Rev. Lett.
94
,
246802
(
2005
).
1032.
V.
Garcia
and
M.
Bibes
,
Nat. Commun.
5
,
4289
(
2014
).
1033.
Z.
Wen
and
D.
Wu
,
Adv. Mater.
32
,
1904123
(
2019
).
1034.
S.
Boyn
,
V.
Garcia
,
S.
Fusil
,
C.
Carrétéro
,
K.
Garcia
,
S.
Xavier
,
S.
Collin
,
C.
Deranlot
,
M.
Bibes
, and
A.
Barthélémy
,
APL Mater.
3
,
061101
(
2015
).
1035.
Z.
Wen
,
D.
Wu
,
A.
Li
, and
N.
Ming
,
Nat. Mater.
12
,
617
(
2013
).
1036.
S.
Boyn
,
A. M.
Douglas
,
C.
Blouzon
,
P.
Turner
,
A.
Barthélémy
,
M.
Bibes
,
S.
Fusil
,
J. M.
Gregg
, and
V.
Garcia
,
Appl. Phys. Lett.
109
,
232902
(
2016
).
1037.
Z.
Li
,
X.
Guo
,
H.-B.
Lu
,
Z.
Zhang
,
D.
Song
,
S.
Cheng
,
M.
Bosman
,
J.
Zhu
,
Z.
Dong
, and
W.
Zhu
,
Adv. Mater.
26
,
7185
(
2014
).
1038.
F.
Ambriz-Vargas
,
G.
Kolhatkar
,
R.
Thomas
,
R.
Nouar
,
A.
Sarkissian
,
C.
Gomez-Yáñez
,
M. A.
Gauthier
, and
A.
Ruediger
,
Appl. Phys. Lett.
110
,
093106
(
2017
).
1039.
F.
Ambriz-Vargas
,
G.
Kolhatkar
,
M.
Broyer
,
A.
Hadj-Youssef
,
R.
Nouar
,
A.
Sarkissian
,
R.
Thomas
,
C.
Yáñez
,
M. A.
Gauthier
, and
A.
Ruediger
,
ACS Appl. Mater. Interfaces
9
,
13262
(
2017
).
1040.
Y.
Wei
,
S.
Matzen
,
T.
Maroutian
,
G.
Agnus
,
M.
Salverda
,
P.
Nukala
,
J.
Ye
,
Q.
Chen
,
P.
Lecoeur
, and
B.
Noheda
,
Phys. Rev. Appl.
12
,
031001
(
2019
).
1041.
A.
Chanthbouala
,
V.
Garcia
,
R. O.
Cherifi
,
K.
Bouzehouane
,
S.
Fusil
,
X.
Moya
,
S.
Xavier
,
H.
Yamada
,
C.
Deranlot
,
N. D.
Mathur
,
M.
Bibes
,
A.
Barthélémy
, and
J.
Grollier
,
Nat. Mater.
11
,
860
(
2012
).
1042.
L.
Chua
,
IEEE Trans. Circuit Theory
18
,
507
(
1971
).
1043.
1044.
1045.
V.
Garcia
,
M.
Bibes
,
L.
Bocher
,
S.
Valencia
,
F.
Kronast
,
A.
Crassous
,
X.
Moya
,
S.
Enouz-Vedrenne
,
A.
Gloter
,
D.
Imhoff
,
C.
Deranlot
,
N. D.
Mathur
,
S.
Fusil
,
K.
Bouzehouane
, and
A.
Barthélémy
,
Science
327
,
1106
(
2010
).
1046.
D.
Pantel
,
S.
Goetze
,
D.
Hesse
, and
M.
Alexe
,
Nat. Mater.
11
,
289
(
2012
).
1047.
Y. W.
Yin
,
J. D.
Burton
,
Y.-M.
Kim
,
A. Y.
Borisevich
,
S. J.
Pennycook
,
S. M.
Yang
,
T. W.
Noh
,
A.
Gruverman
,
X. G.
Li
,
E. Y.
Tsymbal
, and
Q.
Li
,
Nat. Mater.
12
,
397
(
2013
).
1048.
Y. A.
Bychkov
and
E. I.
Rashba
,
JETP Lett.
39
,
78
(
1984
).
1049.
A.
Manchon
,
H. C.
Koo
,
J.
Nitta
,
S. M.
Frolov
, and
R. A.
Duine
,
Nat. Mater.
14
,
871
(
2015
).
1050.
J.
Varignon
,
L.
Vila
,
A.
Barthélémy
, and
M.
Bibes
,
Nat. Phys.
14
,
322
(
2018
).
1051.
V. M. M.
Edelstein
,
Solid State Commun.
73
,
233
(
1990
).
1052.
J. C.
Rojas-Sánchez
,
L.
Vila
,
G.
Desfonds
,
S.
Gambarelli
,
J. P.
Attané
,
J. M.
De Teresa
,
C.
Magén
, and
A.
Fert
,
Nat. Commun.
4
,
2944
(
2013
).
1053.
K.
Ishizaka
,
M. S.
Bahramy
,
H.
Murakawa
,
M.
Sakano
,
T.
Shimojima
,
T.
Sonobe
,
K.
Koizumi
,
S.
Shin
,
H.
Miyahara
,
A.
Kimura
,
K.
Miyamoto
,
T.
Okuda
,
H.
Namatame
,
M.
Taniguchi
,
R.
Arita
,
N.
Nagaosa
,
K.
Kobayashi
,
Y.
Murakami
,
R.
Kumaiv
,
Y.
Kaneko
,
Y.
Onose
, and
Y.
Tokura
,
Nat. Mater.
10
,
521
(
2011
).
1054.
1055.
D.
Di Sante
,
R.
Bertacco
,
P.
Barone
, and
S.
Picozzi
,
Adv. Mater.
25
,
509
(
2013
).
1056.
A. V.
Kolobov
,
D. J.
Kim
,
A.
Giussani
,
P.
Fons
,
J.
Tominaga
,
R.
Calarco
, and
A.
Gruverman
,
APL Mater.
2
,
066101
(
2014
).
1057.
S.
Varotto
,
L.
Nessi
,
S.
Cecchi
,
J.
Sławińska
,
P.
Noël
,
S.
Petrò
,
F.
Fagiani
,
A.
Novati
,
M.
Cantoni
,
D.
Petti
,
E.
Albisetti
,
M.
Costa
,
R.
Calarco
,
M. B.
Nardelli
,
M.
Bibes
,
S.
Picozzi
,
J.-P.
Attané
,
L.
Vila
,
R.
Bertacco
, and
C.
Rinaldi
, arXiv:2103.07646 [cond-mat.mtrl-sci].
1058.
C.
Rinaldi
,
S.
Varotto
,
M.
Asa
,
J.
Sławińska
,
J.
Fujii
,
G.
Vinai
,
S.
Cecchi
,
D.
Di Sante
,
R.
Calarco
,
I.
Vobornik
,
G.
Panaccione
,
S.
Picozzi
, and
R.
Bertacco
,
Nano Lett.
18
,
2751
(
2018
).
1059.
J.
Krempaský
,
S.
Muff
,
J.
Minár
,
N.
Pilet
,
M.
Fanciulli
,
A. P.
Weber
,
E. B.
Guedes
,
M.
Caputo
,
E.
Müller
,
V. V.
Volobuev
,
M.
Gmitra
,
C. A. F.
Vaz
,
V.
Scagnoli
,
G.
Springholz
, and
J. H.
Dil
,
Phys. Rev. X
8
,
021067
(
2018
).
1060.
L. G. D.
da Silveira
,
P.
Barone
, and
S.
Picozzi
,
Phys. Rev. B
93
,
245159
(
2016
).
1061.
L. L.
Tao
and
J.
Wang
,
J. Appl. Phys.
120
,
234101
(
2016
).
1062.
J.
Varignon
,
J.
Santamaria
, and
M.
Bibes
, arXiv:1808.03123 (
2018
).
1063.
H.
Djani
,
A. C.
Garcia-Castro
,
W.-Y.
Tong
,
P.
Barone
,
E.
Bousquet
,
S.
Picozzi
, and
P.
Ghosez
,
npj Quantum Mater.
4
,
51
(
2019
).
1064.
P.
Noël
,
F.
Trier
,
L. M. V.
Arche
,
J.
Bréhin
,
D. C.
Vaz
,
V.
Garcia
,
S.
Fusil
,
A.
Barthélémy
,
L.
Vila
,
M.
Bibes
, and
J.-P.
Attané
,
Nature
580
,
483
(
2020
).
1065.
H.
Mirhosseini
,
I. V. V.
Maznichenko
,
S.
Abdelouahed
,
S.
Ostanin
,
A.
Ernst
,
I.
Mertig
, and
J.
Henk
,
Phys. Rev. B
81
,
073406
(
2010
).
1066.
Z. C.
Zhong
,
L.
Si
,
Q. F.
Zhang
,
W. G.
Yin
,
S.
Yunoki
, and
K.
Held
,
Adv. Mater. Interfaces
2
,
1400445
(
2015
).
1067.
Y.
Song
,
D.
Zhang
,
B.
Xu
,
K.
Chang
, and
C. W.
Nan
, arXiv:1909.03727 (
2019
).
1068.
E.
Lesne
,
Y.
Fu
,
S.
Oyarzun
,
J. C.
Rojas-Sánchez
,
D. C.
Vaz
,
H.
Naganuma
,
G.
Sicoli
,
J.-P.
Attané
,
M.
Jamet
,
E.
Jacquet
,
J.-M.
George
,
A.
Barthélémy
,
H.
Jaffrès
,
A.
Fert
,
M.
Bibes
, and
L.
Vila
,
Nat. Mater.
15
,
1261
(
2016
).
1069.
D. C.
Vaz
,
P.
Noël
,
A.
Johansson
,
B.
Göbel
,
F. Y.
Bruno
,
G.
Singh
,
S.
McKeown-Walker
,
F.
Trier
,
L. M.
Vicente-Arche
,
A.
Sander
,
S.
Valencia
,
P.
Bruneel
,
M.
Vivek
,
M.
Gabay
,
N.
Bergeal
,
F.
Baumberger
,
H.
Okuno
,
A.
Barthélémy
,
A.
Fert
,
L.
Vila
,
I.
Mertig
,
J.-P.
Attané
, and
M.
Bibes
,
Nat. Mater.
18
,
1187
(
2019
).
1070.
D. C.
Vaz
,
F.
Trier
,
A.
Dyrdał
,
A.
Johansson
,
K.
Garcia
,
A.
Barthélémy
,
I.
Mertig
,
J.
Barnaś
,
A.
Fert
, and
M.
Bibes
,
Phys. Rev. Mater.
4
,
071001(R)
(
2020
).
1071.
F.
Trier
,
D. C.
Vaz
,
P.
Bruneel
,
P.
Noël
,
A.
Fert
,
L.
Vila
,
J.-P.
Attané
,
A.
Barthélémy
,
M.
Gabay
,
H.
Jaffrès
, and
M.
Bibes
,
Nano Lett.
20
,
395
(
2020
).
1072.
J.
Hemberger
,
P.
Lunkenheimer
,
R.
Viana
,
R.
Bohmer
, and
A.
Loidl
,
Phys. Rev. B
52
,
13159
(
1995
).