Next Article in Journal
Nanotechnology-Based Strategies for Effective and Rapid Detection of SARS-CoV-2
Next Article in Special Issue
Investigation of Deoxidation Process of MoO3 Using Environmental TEM
Previous Article in Journal
Evaluation of Physical Characteristics and Sorption of Cement Mortars with Recycled Ceramic Aggregate
Previous Article in Special Issue
Monitoring Electrical Biasing of Pb(Zr0.2Ti0.8)O3 Ferroelectric Thin Films In Situ by DPC-STEM Imaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Structural Phase Transition and In-Situ Energy Storage Pathway in Nonpolar Materials: A Review

by
Xian-Kui Wei
1,*,
Rafal E. Dunin-Borkowski
1 and
Joachim Mayer
1,2
1
Ernst Ruska-Centre for Microscopy and Spectroscopy with Electrons, Research Centre Jülich, 52425 Jülich, Germany
2
Gemeinschaftslabor für Elektronenmikroskopie (GFE), RWTH Aachen University, 52074 Aachen, Germany
*
Author to whom correspondence should be addressed.
Materials 2021, 14(24), 7854; https://doi.org/10.3390/ma14247854
Submission received: 9 November 2021 / Revised: 9 December 2021 / Accepted: 16 December 2021 / Published: 18 December 2021
(This article belongs to the Special Issue Materials Characterizations Using In-Situ Techniques)

Abstract

:
Benefitting from exceptional energy storage performance, dielectric-based capacitors are playing increasingly important roles in advanced electronics and high-power electrical systems. Nevertheless, a series of unresolved structural puzzles represent obstacles to further improving the energy storage performance. Compared with ferroelectrics and linear dielectrics, antiferroelectric materials have unique advantages in unlocking these puzzles due to the inherent coupling of structural transitions with the energy storage process. In this review, we summarize the most recent studies about in-situ structural phase transitions in PbZrO3-based and NaNbO3-based systems. In the context of the ultrahigh energy storage density of SrTiO3-based capacitors, we highlight the necessity of extending the concept of antiferroelectric-to-ferroelectric (AFE-to-FE) transition to broader antiferrodistortive-to-ferrodistortive (AFD-to-FD) transition for materials that are simultaneously ferroelastic. Combining discussion of the factors driving ferroelectricity, electric-field-driven metal-to-insulator transition in a (La1−xSrx)MnO3 electrode is emphasized to determine the role of ionic migration in improving the storage performance. We believe that this review, aiming at depicting a clearer structure–property relationship, will be of benefit for researchers who wish to carry out cutting-edge structure and energy storage exploration.

Graphical Abstract

1. Introduction

Renewable energies harvested from solar, wind and chemical fuels are playing ever-greater roles in our lives [1,2]. However, their widespread utilization is largely impeded by the underdeveloped energy storage technologies. Thus far, popular electrical energy storage systems consist of the following categories: (1) solid oxide fuel cells (SOFCs), (2) batteries, (3) electrochemical capacitors and (4) dielectric capacitors. From the Ragone plot, one can see that the SOFCs have the highest energy density, while the electrostatic capacitors possess the highest power density, which is up to 107 W/kg. In between, the electrochemical capacitors show a tendency of partially replacing batteries due to their fast-growing energy density; see Figure 1. Specifically, the ultrafast charging/discharging rates, at microsecond level, make dielectric capacitors widely used in devices such as motor starters, high-power lasers, signal processing and sensors [3]. In retrospect, the technological advancement may date back to the ferroelectric (FE) phenomena of Rochelle salt found by Valasek in 1921 [4,5]. During World War II [6], the discovery of BaTiO3 and its usage in high-energy-density capacitors launched a research boom in the field of FE materials. Possibly enlightened by the whimsical concept of antimatter introduced by Schuster in 1898 [7], Néel and Kittel proposed the concepts of antiferromagnet and antiferroelectric (AFE) in 1936 and 1951, respectively, to explain magnetic and dielectric anomalies at Curie temperature (TC) [8,9].
Although receiving less attention than FEs, AFE materials have been investigated for their intriguing physical properties, including the origin of antiferroelectricity [10,11,12,13], electromechanical coupling [14], electrocaloric effect [15,16,17] and negative capacitance [18]. In fact, AFE-based materials have been proposed to design memory devices [19,20]. In recent years, both AFE-based and relaxor FE (RFE)-based capacitors [21,22,23,24,25,26] have been widely investigated due to their reducible energy loss [22,27], improvable maximum polarization (Pmax), breakdown electric field (EB) [28,29,30] and energy efficiency (η) [31,32]. Recently, a series of promising data have been reported [32,33,34,35,36]. Using a superparaelectric design, Pan et al. achieved very high recoverable energy density (Jrec = 152 J/cm3) in BiFeO3-BaTiO3-based RFE thin films [21,37,38]. By means of Sr doping, Acharya et al. reported ultrahigh energy efficiency (η = 97%) in AFE Pb1-xSrxHfO3 thin films [39], where the Jrec and EB were 77 J/cm3 and 5.12 MV/cm, respectively. More surprisingly, Hou et al. found the highest energy density reported thus far (Jrec = 307 J/cm3) in SrTiO3/La1-xSrxMnO3 (STO/LSMO, x ≈ 1/3) thin films, which have values of Pmax ≈ 125 μC/cm2, efficiency η ≈ 89% and EB ≈ 6.8 MV/cm [28]. For representative dielectric films and bulks, a comparison of their energy storage performance is shown in Table 1.
Given a “mismatched” structure–property relationship, one may naturally ask: how does a nonpolar SrTiO3 deliver such a high Pmax value and energy storage density? Although static structural and compositional characterizations are very helpful, in situ phase transition and storage pathway studies are more promising in unveiling the hidden mysteries. Compared with macroscopic property investigation, in situ dynamic structure study lags far behind in the field of dielectric capacitors. An important reason lies in the fact that there exists a large mismatch between the structural response time for energy storage [47] and the data collection time [48,49,50]. For the former, this is usually completed at the millisecond scale or within an even shorter time. Meanwhile, for the latter, e.g., X-ray diffraction (XRD) and selected area electron diffraction (SAED), they usually take several seconds or even minutes in collecting a dataset. In comparison with RFEs [21,23,51], AFEs are more widely investigated in structural phase transitions under in-situ conditions. The reason resides in the fact that the AFE-to-FE transitions inherently couple with the energy storage process [32,52,53,54].
Building on the established phase transition framework, this review seeks to broaden the research scope from AFEs to nonpolar materials, which are simultaneously ferroelastic or antiferroelastic, in the quest for promising energy storage materials. Under in situ conditions, this entails the extension of well-known AFE-to-FE transition to broader antiferrodistortive-to-ferrodistortive (AFD-to-FD) transition. By summarizing external field-driven phase transitions, research progress regarding in situ and atomic-scale structural characterization, achieved by using direct light-element imaging, is highlighted to understand the energy storage mechanism. Associated with ionic migration across the dielectric/electrode interface, we further discuss the electric field-driven metal-to-insulator (M-I) transition in electrodes and its potential impact on tuning the macroscopic energy storage density of dielectric capacitors.

2. Fundamentals of Capacitor Energy Storage

A dielectric capacitor is constructed in a parallel-plate form, i.e., a dielectric layer sandwiched by two conductive electrodes. The physical quantity that manifests the energy storage is capacitance (C), which can be described by the following equation:
C = ε 0 ε r A / d
where ε0 is the dielectric permittivity in vacuum (~8.85 × 10−12 F/m), εr (>> 1) is the relative dielectric constant of the dielectric layer, A is the overlapping area of the electrodes, and d is thickness of the dielectric layer. In principle, the capacitance only depends on the geometry of the capacitor and permittivity of the dielectric layer. However, this is not always the case, especially when electric field-driven electrochemical activities take place near the dielectric/electrode interface, as discussed below.
When an external voltage is applied, charges with opposite signs and equal magnitude accumulate at the electrodes. This is the so-called charging process. The charges form an internal electric field, whose direction is opposite to that of the external electric field. The charging process finishes when the internal electric field induced by the accumulated charges (Q = CV) is equal to the external field (E = V/d), where V is the applied voltage. During the charging process, the charges are moved by the external electric field, and electrostatic energy is stored in the dielectric layer. The stored energy can be calculated from the following expression:
W = 0 Q m a x V d q
where Qmax is the maximum charge when the charging process finishes, and dq is an increment of charge. One of the key figures of merit of the dielectric capacitor is its energy density (J), which measures its “capability” for storage performance and can be written in the following form:
J = W A d = 0 P m a x E d P
The energy density can therefore easily be obtained by integrating the area between the polarization and electric field axes in the P-E loop. Another key figure of merit for the capacitor is energy efficiency (η). It is equals to the ratio between the recoverable energy density (Jrec) and total energy density (Jtot), which can be expressed by
η = (Jrec/Jtot) × 100%
where the Jtot is a sum of the P-E loop (Jloss) and its left-side area (Jrec) at E ≥ 0 in the P-E relation chart (see Figure 2b). It appears that a linear dielectric, with Jrec 1 2 ε 0 ε r E 2 and very small Jloss, may offer higher storage density. In practice, phase transition materials offer much higher energy storage densities due to their much larger EB and maximum polarization (Pmax), which can be one order of magnitude higher than that of linear dielectrics [39,55]. As for more circuit-related details of energy storage measurement, this has been discussed elsewhere [24,40].
From a microstructure point of view, complex structural changes take place during the charging process. Inside the nonpolar dielectric layer, this involves inversion symmetry breaking and the emergence of electric dipoles from the originally centrosymmetric lattices. Irrelevant to either displacive or order–disorder phase transition [56,57], the electric dipoles are defined by the separation of positive and negative charge centers within each unit cell. At mesoscopic scale, this is affected by phase constitution, domain reorientation and growth, defect type and density. Near the dielectric/electrode interface, polarization screening, chemical diffusion, ionic migration and potential M–I transition significantly impact the energy storage and release. Therefore, unveiling the dynamic structural responses under in situ conditions may greatly deepen our mechanistic understanding, which is essential for the development of new materials and devices.

3. Structural Features and Phase Transitions in Antiferroelectrics

FEs undergo a nonpolar-to-polar transition at TC and the breaking of spatial inversion symmetry leads to the emergence of spontaneous polarization (PS) in the low-temperature phase. Under the application of an electric field (E), the direction of polarization (P) in domains can be reversed and its nonlinear response to E gives rise to a P-E hysteresis loop. Depending on the category of the FEs, the hysteresis loop may vary between a square shape (e.g., in BaTiO3) [58] and a slim shape (e.g., in relaxor Na0.5Bi0.5TiO3) [37]. Thus, the energy consumption is usually larger for the former than the latter due to its larger Pr (a value at E = 0) and loop areas. In contrast, AFEs undergo a phase transition between two nonpolar phases at TC, around which an FE phase may transiently exist in a narrow temperature window [59,60,61]; see Figure 2a. Despite zero spontaneous polarization at T < TC, a strong E can drive an AFE-to-FE phase transition [52,62], which gives rise to a double hysteresis loop; see Figure 2b. Due to the concave curvature of the P-E relationship, with a reflection point at coercive field EA or EB, the AFEs usually possess intrinsically high energy storage densities [63].
In a crystal structure, an important feature of AFE is that its unit cell volume is doubled with respect to its paraelectric (PE) phase (Figure 2c). Benefiting from the antiparallel arrangement of electric dipoles within sublattices, null polarization is obtained on the unit cell scale. Typical AFE systems include PbZrO3, AgNbO3, NaNbO3, HfO2 [64,65,66,67,68,69,70,71,72] and recently reported 2D van der Waals AFE β’-In2Se3 [73,74]. From the perspective of lattice dynamics, the emergence of antiferroelectricity is a consequence of competing lattice instabilities [10,11], which are manifested by delicate structural orders [66,75]. Taking PbZrO3 as an example, its AFE phase is dominated by antiparallel Pb displacements along the [100]O//[110]c direction (O: orthorhombic; c: pseudocubic) and antiferrodistortive (AFD) oxygen octahedra (a-a-c0 in Glazer’s notation) [66]. The corresponding Σ and R modes in the Brillouin zone (Figure 2d,f) are represented by wave vectors of kΣ = (2π/a)(1/4, 1/4, 0) (a is the cubic lattice constant) and kR = (2π/a)(1/2, 1/2, 1/2), respectively [10,11].
A polar instability, i.e., the Γ mode, manifested by opposite shifts of Pb cations and oxygen anions, is responsible for ferroelectricity (Figure 2e). In the ground state, the polar instability is suppressed due to its higher energy level. Nevertheless, it has been shown that the energy levels of different lattice instabilities can be subverted in the following situations. (1) Application of large E can destabilize the Σ and R modes and drive an AFE-to-FE transformation [76,77,78]. (2) Interruption of the cationic displacement order and AFD order at translational boundaries allows the emergence of local ferroelectricity with a bi-stable feature [79,80,81]. (3) An intrinsic surface effect may trigger an AFE-to-FE transition as the sample thickness is below a critical thickness, ~6.5 nm for PbZrO3 [82,83]. (4) Chemical doping can also alter the cationic displacement order and AFD order, e.g., stochastic stacking of Pb-displacement-based stripes in La-doped PbZrO3. Due to the uncompensated antiparallel arrangement of Pb cations, weak ferrielectricity emerges in doped systems [84,85,86].

4. Antiferrodistortive and Ferrodistortive Phase Transitions

In fact, both FE-to-PE and AFE-to-PE transitions can be understood from the viewpoint of ferroelastic phase transition, which is related to a change in lattice symmetry around TC [87,88]. In many cases, ferroelasticity or antiferroelasticity is a secondary or improper effect, since the driving force is related to cation or molecular ordering and the softening of optical phonon branches [89,90]. Thus, a large proportion of ferroelastic transitions depend on mode condensation at a position in the Brillouin zone. When modes condense at zone boundaries, the transition is termed AFD [91]. For a perovskite, the AFD order is manifested by in-phase and out-of-phase rotation of the oxygen octahedra [92,93]. Thus, the AFEs constitute a subgroup of AFD transitions (Figure 3a,b). In contrast, when modes condense at the zone center, the transition is termed FD. Correspondingly, the FEs constitute a subgroup of FD transitions (Figure 3c,d). It should be noted that the AFD phases are characterized by the rigid octahedral linking through out-of-phase tilting, in-phase tilting or their combination. In FD phases, associated with symmetry lowering, polar octahedral distortion can modify or destroy the rigid linking and response to the applied electric field together with A-site cations. In addition, there are also antipolar (AP) and pyroelectric (PyE) phases, which belong to the subgroups of AFD and FD transitions, respectively. In contrast to AFE and FE phases, their dipoles cannot be reversed, even by a strong electric field [91].
It should be noted that the zone-center mode condensation is not necessarily bound to the zone-boundary condensation in proper FEs. This means that the polar order can either be irrelevant to the AFD order (e.g., in BaTiO3 and PbTiO3) or couple cooperatively with the AFD order (e.g., in BiFeO3, ZnSnO3 and ScFeO3) [94,95,96]. Besides the routine manifestations, as a counterpart of AFD, the FD phase transition may place extra emphasis on a means of inducing ferroelectricity through the long-range polar distortion of oxygen octahedra. This broadens the perspectives of understanding the origins of ferroelectricity and implies that more nonpolar dielectrics with typical AFD order [92,93] can be transformed to novel FD phases under excitation of an external field [97]. Being compatible with chemical, defect and strain engineering [30,98], therefore, the range of candidate energy storage systems can be greatly expanded by the AFD-to-FD transition.
With reference to the above classifications, the STO/LSMO system, with ultrahigh energy density (307 J/cm3) at a thickness range of 410~710 nm for STO, provides a good example to analyze the AFD-to-FD transition [28,99]. It has been widely accepted that the cubic SrTiO3 (space group Pm 3 ¯ m) transforms to a tetragonal structure (I4/mcm) via a second-order AFD phase transition at TC = 105 K. Below TC, the tetragonal phase is characterized by the antiphase rotation of the octahedra along its c axis (a0a0c- in Glazer’s notation) [100]. The temperature–strain phase diagram [101,102,103,104] shows that the FE phase can be stabilized through strain engineering (Figure 4a). Nonetheless, the following results indicate that the mechanism leading to polarity in SrTiO3 is not that simple. Compared with strained SrTiO3, Jang et al. demonstrate that strain-free SrTiO3 is an RFE, with a temperature for the maximum εr at Tm ≈ 45 K [29]. The role of strain is to stabilize the long-range correlation of preexisting nanopolar regions; see Figure 4b,c. Another argument points out that the abnormal ferroelectricity at T < Tm arises from FE antiphase boundaries [105,106,107]. Similar to AFE PbZrO3, once the nonpolar lattice instabilities become destabilized, the “hidden” polar lattice instability prevails and gives rise to the ferroelectricity.
Excited by either an optical pump or a THz electric field (with a threshold field amplitude of ~300 kV/cm), two research groups independently reported that metastable ferroelectricity with TC ≈ 290 K can be achieved in SrTiO3 [102,108]. By introducing Ti/O-deficient nanoregions [30], Li et al. showed that defect engineering can also drive the occurrence of AFD-to-FD transition. Using this strategy, they realized a very large lattice tetragonality ratio (c/a = 1.038), strong room-temperature ferroelectricity (PS = 41.6 μC/cm2) and very high TC (~1098 K) in SrTiO3 films; see Figure 4d. Using first-principles calculations, Klyukin and Alexandrov reported that antisite defects, either Ti on a Sr site (TiSr) or vice versa (SrTi), can result in large electric polarization in SrTiO3 [109,110,111]. These outcomes suggest that structural defects should play an important role in achieving ultrahigh polarization in the STO/LSMO system (see inset of Figure 4d). More importantly, this sets a precedent for retrofitting other nonpolar perovskites into energy storage media via chemical and defect engineering, e.g., CaTiO3 and DyScO3 [55,98,112].

5. Anisotropic Energy Storage

The double P-E hysteresis loop shown in Figure 2b indicates that synergistically changing the following parameters may optimize the energy storage performance: (1) simultaneously increasing the critical electric fields EA and EB and minimizing the loop area [31]; (2) increasing the Pmax [28] and minimizing the Pr [32]. In practice, due to the dielectric anisotropy of the crystals [113], the energy storage density of identical materials also exhibits a direction-dependent nature [71,97,114,115]. Taking PbZrO3 as an example, multiscale first-principles computations show that as the E is applied along different crystallographic directions, its phase transition pathways are distinct (Figure 5a). On this basis, Lisenkov et al. found three high-strain polar phases: a monoclinic (Cc) phase, an orthorhombic (Ima2) phase and a tetragonal (I4cm) phase [114].
In experiments, the key electrical parameters relating to energy storage also show strong crystal orientation dependence. AFE (200 nm) PbHfO3 films grown on (100)C-plane, (110)C-plane and (111)C-plane terminated SrTiO3 substrates, which are buffered by SrRuO3 electrodes, are good examples that confirm this [71,72]. Together with large changes in critical field and Pr, the Pmax takes values of 30.54, 38.91 and 32.64 μC/cm2 in the three films, respectively (Figure 5b). Corresponding to a larger coercive field and Pmax, the (110)C oriented film delivers higher energy storage density (Jrec = 21.4 J/cm3) than the other two (Jrec ≈ 16 J/cm3) [71]. From a microstructural perspective, one may expect that the type and density of structural defects [116,117,118,119] may influence the energy storage performance. Through introducing local compressive pressure, Zhang et al. reported that enhanced critical fields may increase the Jrec [53,67,120], e.g., from 9 to 16.2 J/cm3 in Li+-La3+ co-doped PbZrO3 films. By constructing a ferrielectric (FiE) M2-M3 phase boundary, Luo et al. achieved an energy density of 6.3 J/cm3 with η = 90% in (1-x)AgNbO3-xAgTaO3 solid solution [46].

6. In Situ Structural Shase Sransition Sathway

In conjunction with macroscopic property measurement, in situ dynamic structural studies can help to construct structure–property relationships directly. Under the application of an electric field, answers to the following questions are expected to play critical roles in further improving the energy storage performance. Q1) How does electrical polarization emerge and evolve from a nonpolar lattice matrix? Q2) What are the key structural factors that control and limit critical fields (EA and EB), Pmax and Pr? Although the phase transition processes are complex, as reported in PbZrO3-based and NaNbO3-based systems [121,122,123,124], manipulative free dimensions about spatial, temporal and electric field offer a plethora of opportunities to unveil the unknowns. For example, in Pb(Zr0.57Sn0.43)O3-based ceramics, Fan et al. observed local depolarization field-assisted AFE-to-FE transition during monotonic E loading using in situ (scanning) transmission electron microscopy (S/TEM). In addition to this, they also found suppression of FE domain mobility after 103-time bipolar cycling, which indicates the electric fatigue of the FE phase [125].
By largely improving the temporal resolution under in situ biasing conditions, from ~30 s to 415 ms per data pattern, Zhang et al. investigated the dynamic structure evolution of polycrystalline NaNbO3 using high-energy XRD (Figure 6a). Associated with the disappearance and appearance of specific superlattice reflections, e.g., ½(312)c reflection, they found that an AFE-to-FE transition took place around E = 8 kV/mm [126]. At E > 12 kV/mm, abruptly enhanced polarization and longitudinal strain S33 indicated the transformation of the nonpolar P phase (orthorhombic AFE, Pbcm) to the polar Q phase (orthorhombic FE, P21ma) (Figure 6b–d). According to the mismatch of the P-phase fraction with changes in lattice parameter, volume, polarization and S33, they proposed a decoupled polarization switching process in the range of 8 < E < 12 kV/mm. Given that the P phase (a-a-b+/a-a-b-/a-a-b+) differs from the Q phase (a-a-b+) in octahedral tilting, the field-dependent structural data cannot exclude another possibility, i.e., the existence of an intermediate FiE phase. It adopts a Q-phase structural framework [98], but the oxygen octahedra are in a transitional state. Observation of weak ferrielectricity in PbZrO3-based ceramics supports this possible interpretation [85,86,127].
Compared with diffraction-based techniques, E-dependent structural transition, measured at atomic scale using S/TEM, provides more intuitive information about energy storage. However, even using a state-of-the-art imaging detector such as the K3 camera, which is capable of collecting 1500 data frames per second (0.67 ms per frame), the mismatch in time remains due to the requirement of sufficient signal intensity and image contrast for data analysis. Thus far, several advanced imaging techniques have been used to image light elements, e.g., negative spherical aberration imaging (NCSI) [128,129,130], annular bright field (ABF) imaging [131], integrated differential phase contrast (iDPC) imaging [132,133] and electron ptychography [134]. Despite this, imaging light elements such as oxygen under in situ biasing conditions remains a challenge due to their very low scattering power. As a compromise, slowing down the phase transition speed becomes a good solution to unravel the evolution of characteristic structural orders as a function of field strength and time [52].
Recently, Wei et al. found that an illumination electron beam in a TEM can act as an external electric field to trigger phase transitions in dielectric insulators. The principle lies in the fact that the electrons captured by the insulating sample surface are “equivalent” to the electrostatic charging effect [135,136]. By slowing down the phase transition speed to the level of seconds, the investigation of atomic-scale structural changes becomes possible [127,137,138]. For a [001]O-oriented PbZrO3 lamella sample, time- and atomic-resolution NCSI-TEM study [127] reveals that the AFE-to-FE transition involves the splitting of pseudocubic ap and bp axes. As a function of irradiation time, the unit cell volume undergoes a reduction-to-expansion transition as the AFE phase evolves into FE monoclinic (FEM) and rhombohedral (FER) phases (Figure 7a–e). During the in situ energy storage process, an intermediate transient FD phase was observed between the AFE and FE phases. With the preservation of antiparallel Pb displacements, polar octahedral distortion takes place and breaks the spatial inversion symmetry along the x direction. Associated with the extraction of atomic positions through quantitative TEM study, the authors found that the transient FD phase exhibits a cycloidal order of polarization, with PS ≈ 2.8 μC/cm2 (Figure 7c,f). The finding of this FiE phase suggests the origin of the linear polarization response at E < EA(B) in the P-E loop of PbZrO3. Synchrotron X-ray and neutron diffraction structural studies reported similar FiE behavior, characteristic of a wavy polarization order, in ternary PbZrO3-PbSnO3-PbTiO3 solid solutions [85,86,139].
Regarding the phase transition pathway, we should note the possible difference between calculated models and experimental observations (Figure 5 and Figure 7). For the latter, structural defects such as atomic vacancies and stoichiometry issues prevail in real samples [140,141], and they may potentially influence the transition pathway. By using time- and atomic-resolution NCSI-TEM, Wei et al. found that a defect core, induced by oxygen and Pb vacancies [118], can act as a seed to trigger unit-cell-wise AFD-to-FD transition in [001]O-oriented PbZrO3 (Figure 8a–e). This is in sharp contrast to the case with the absence of such a defect core, where a relatively uniform AFD-to-FD transition takes place in the region irradiated by the electron beam [127]. As a result of the seed effect, charged FD domains with head-to-head and tail-to-tail configurations are observed during the energy storage process.
Compared with neutral domain walls, we know that the formation energy of charged domain walls is much higher [117,142] and 2D electron gas may form at the walls to compensate for bound charges and lower the electrostatic energy [113,143]. The anisotropic structural transformation suggests that the defect cores probably lower the coercive field (EB). With the preservation of antiparallel Pb displacements, this further proves that the energy storage and transfer begin with distorting the oxygen octahedral network. Another point worth noting is the evolution of the polar configuration inside the defect core. At the initial AFE state, electric dipoles form a self-compensated vortex structure. Accompanied by energy injection, the local polarity is stabilized until immersion in the FD states. Strain analysis reveals that nanoscale local compressive strain leads to the seed effect of the defect core (Figure 8f–j). This points out that the rational selection of the doping element and concentration is crucial for optimizing the storage performance of dielectrics [144,145,146].

7. Ionic Migration across Dielectric/Metal Interfaces

In oxide-based dielectric capacitors, metallic (La1−xSrx)MnO3 [147], (La1−xSrx)CoO3 [148] and SrRuO3 [149] are standard choices for bottom electrodes. Previous transport property studies have reported that (La1−xSrx)MnO3 undergoes thickness-driven and electric-field-driven M–I transition [137,150,151]. When such a transition occurs in dielectric capacitors, “dead” layers appear at the dielectric/electrode interfaces and this may impact the energy storage performance. Driven by the depolarization effect, the “dead” layers usually lead to reduced capacitance near the interface [152,153,154]. The contrary phenomenon observed in the STO/LSMO system (PS ≈ 125 μC/cm2) suggests that the interface should contribute positively to the “defect-induced” ferroelectricity in SrTiO3 (Figure 4d). Hou et al. ascribe the enhanced EB to the modulation of the local electric field and redistribution of oxygen vacancies at the oxide interface [28]. However, the finite interface thickness (~5 nm), compared to that of ~560 nm for SrTiO3, casts doubt on the role of the interfacial contribution alone.
Rhombohedral LSMO (x ≈ 1/3, space group R 3 ¯ c) has a ferromagnetic metallic state at room temperature. Below TC ≈ 367 K, its structure is characterized by AFD octahedral rotations (a-a-a- in Glazer’s notation) [147,155]. By using in situ biasing STEM, Yao et al. revealed an E-induced M–I transition by measuring the E-dependent resistivity change. The insulating state was attributed to the formation of a brownmillerite (BM) phase (C2/c), which has a bandgap of Eg ≈ 0.63 eV [151,156,157]. In Hf0.5Zr0.5O2 (HZO, 5 nm)/LSMO films (Figure 9a), Nukala et al. directly observed oxygen vacancy-induced structural transition in bottom LSMO by using in situ biasing iDPC STEM [158]. In addition to an intermediate BM precursor phase, they found the intertwining of polarization switching with the migration of oxygen vacancies (Figure 9b–d). Specifically, the role of the dielectric oxide layer, either as a fast conduit or as a source/sink of oxygen migration depending on the oxygen reactivity of the top electrode [158,159], highlights the necessity of considering electrochemical activities for capacitors under working conditions. This may involve the migration of atomic vacancies and cations through antisite defects, and electronic structure changes in all atomic species. The ultrahigh energy density achieved in Na-based perovskite oxides, e.g., ionic conductor (Na0.5Bi0.5)TiO3 [160], also suggests the important role of ionic migration in enhancing the energy storage performance; see Table 1. In addition, possible FE metal states [161,162,163] should be considered at the dielectric/metal interfaces. These mechanisms may apply to the nonpolar Au/STO/LSMO and other systems [164,165].

8. Summary and Outlook

By generalizing the AFE-to-FE transition to a broader AFD-to-FD transition, we summarize in situ phase transition pathway studies in energy storage nonpolar materials. Aiming at identifying the underlying mechanism and improving the energy storage performance of dielectric capacitors, we highlight several key points below to inspire future research.
(I) Besides mesoscale domain structure and phase boundary evolution, light element sensitive imaging techniques should play a key role in unraveling the in situ dynamic phase transition during energy storage. The reason lies in the fact that the delicate energy transfer usually starts from light-element-based structural units, e.g., distortion of the oxygen octahedra and migration of oxygen vacancies. Specifically, transient FE or FiE phases may possibly emerge under non-equilibrium conditions. Apart from deepening our understanding of the energy storage process, such studies promise to expand the recognition scope of FE physics.
(II) Case studies of PbZrO3, SrTiO3 and (La1−xSrx)MnO3 highlight the significance of carrying out defect and interface structure investigation during the energy storage processes. This may include dopant-induced strain change, antisite mechanism and phase boundary evolution under in situ biasing conditions. Establishing the direct microstructure–property relationship may help to optimize the figures of merit of the P-E loops, e.g., increasing the Pmax and critical fields while reducing Pr and the loop area.
Finally, two recent works reported that mesoscale-domain engineering may greatly improve the energy storage density, e.g., from 12.2 to 18.5 J/cm3 in NaNbO3-based relaxor AFE ceramics [44,166]. Pertinent to the beneficial hierarchical domain structures, structure–property relationship study of FE Pb(Zr,Ti)O3 shows that this may arise from phase transition frustration near a tricritical point [167]. Together with the use of advanced computation and simulation methods, such as machine learning, we believe that carrying out cross-scale microstructure study may boost the development of energy storage materials and their device application.

Author Contributions

X.-K.W. conceived the review idea and wrote the manuscript. R.E.D.-B. and J.M. provided full support and revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

X.-K. Wei acknowledges the support of the Deutsche Forschungsgemeinschaft (DFG; German Research Foundation) under Germany’s Excellence Strategy-Cluster of Excellence Matter and Light for Quantum Computing (ML4Q) EXC 2004/1-390534769.

Conflicts of Interest

The authors declare no conflict of Interest.

References

  1. Chu, S.; Majumdar, A. Opportunities and Challenges for a Sustainable Energy Future. Nature 2012, 488, 294–303. [Google Scholar] [CrossRef] [PubMed]
  2. Zou, X.; Zhang, Y. Noble Metal-Free Hydrogen Evolution Catalysts for Water Splitting. Chem. Soc. Rev. 2015, 44, 5148–5180. [Google Scholar] [CrossRef] [PubMed]
  3. Applications of Capacitors. Available online: https://en.wikipedia.org/wiki/Applications_of_capacitors (accessed on 11 August 2021).
  4. Lüker, A. A Short History of Ferroelectricity. Sol-Gel Derived Ferroelectric Thin Films for Voltage Tunable Applications. Ph.D. Thesis, University of Cranfield, Cranfield, UK, 2009. [Google Scholar]
  5. Valasek, J. Piezo-Electric and Allied Phenomena in Rochelle Salt. Phys. Rev. 1921, 17, 475–481. [Google Scholar] [CrossRef]
  6. Randall, C.A.; Newnham, R.E.; Cross, L.E. History of the First Ferroelectric Oxide, BaTiO3. Available online: https://ceramics.org/ (accessed on 11 August 2021).
  7. Schuster, A. Potential Matter—A Holiday Dream. Nature 1989, 58, 1503. [Google Scholar] [CrossRef] [Green Version]
  8. Barbara, B. Louis Néel: His Multifaceted Seminal Work in Magnetism. Comptes Rendus Phys. 2019, 20, 631–649. [Google Scholar] [CrossRef]
  9. Kittel, C. Theory of Antiferroelectric Crystals. Phys. Rev. 1951, 82, 729–732. [Google Scholar] [CrossRef]
  10. Tagantsev, A.K.; Vaideeswaran, K.; Vakhrushev, S.; Filimonov, A.; Burkovsky, R.G.; Shaganov, A.; Andronikova, D.; Rudskoy, A.I.; Baron, A.; Uchiyama, H.; et al. The Origin of Antiferroelectricity in PbZrO3. Nat. Commun. 2013, 4, 2229. [Google Scholar] [CrossRef] [Green Version]
  11. Hlinka, J.; Ostapchuk, T.; Buixaderas, E.; Kadlec, C.; Kuzel, P.; Gregora, I.; Kroupa, J.; Savinov, M.; Klic, A.; Drahokoupil, J.; et al. Multiple Soft-Mode Vibrations of Lead Zirconate. Phys. Rev. Lett. 2014, 112, 197601. [Google Scholar] [CrossRef]
  12. Xu, B.; Hellman, O.; Bellaiche, L. Order-Disorder Transition in the Prototypical Antiferroelectric PbZrO3. Phys. Rev. B 2019, 100, 020102. [Google Scholar] [CrossRef] [Green Version]
  13. Iguez, J.; Stengel, M.; Prosandeev, S.; Bellaiche, L. First-Principles Study of the Multimode Antiferroelectric Transition in PbZrO3. Phys. Rev. B 2014, 90, 220103. [Google Scholar] [CrossRef] [Green Version]
  14. Zhuo, F.; Li, Q.; Zhou, Y.; Ji, Y.; Yan, Q.; Zhang, Y.; Xi, X.; Chu, X.; Cao, W. Large Field-Induced Strain, Giant Strain Memory Effect, and High Thermal Stability Energy Storage in (Pb,La)(Zr,Sn,Ti)O3 Antiferroelectric Single Crystal. Acta Mater. 2018, 148, 28–37. [Google Scholar] [CrossRef]
  15. Vales-Castro, P.; Faye, R.; Vellvehi, M.; Nouchokgwe, Y.; Perpiñà, X.; Caicedo, J.M.; Jordà, X.; Roleder, K.; Kajewski, D.; Perez-Tomas, A.; et al. Origin of Large Negative Electrocaloric Effect in Antiferroelectric PbZrO3. Phys. Rev. B 2021, 103, 054112. [Google Scholar] [CrossRef]
  16. Geng, W.; Liu, Y.; Meng, X.; Bellaiche, L.; Scott, J.F.; Dkhil, B.; Jiang, A. Giant Negative Electrocaloric Effect in Antiferroelectric La-Doped Pb(ZrTi)O3 Thin Films near Room Temperature. Adv. Mater. 2015, 27, 3165–3169. [Google Scholar] [CrossRef] [PubMed]
  17. Mischenko, A.S.; Zhang, Q.; Scott, J.F.; Whatmore, R.W.; Mathur, N.D. Giant Electrocaloric Effect in Thin-Film PbZr0.95Ti0.05O3. ChemInform 2006, 311, 1270–1271. [Google Scholar] [CrossRef] [Green Version]
  18. Qiao, L.; Song, C.; Sun, Y.; Fayaz, M.U.; Lu, T.; Yin, S.; Chen, C.; Xu, H.; Ren, T.-L.; Pan, F. Observation of Negative Capacitance in Antiferroelectric PbZrO3 Films. Nat. Commun. 2021, 12, 4215. [Google Scholar] [CrossRef] [PubMed]
  19. Zhou, Z.; Yang, Q.; Liu, M.; Zhang, Z.; Zhang, X.; Sun, D.; Nan, T.; Sun, N.; Chen, X. Antiferroelectric Materials, Applications and Recent Progress on Multiferroic Heterostructures. SPIN 2015, 05, 1530001. [Google Scholar] [CrossRef]
  20. Vopson, M.M.; Caruntu, G.; Tan, X. Polarization Reversal and Memory Effect in Anti-Ferroelectric Materials. Scr. Mater. 2017, 128, 61–64. [Google Scholar] [CrossRef] [Green Version]
  21. Pan, H.; Li, F.; Liu, Y.; Zhang, Q.; Wang, M.; Lan, S.; Zheng, Y.; Ma, J.; Gu, L.; Shen, Y.; et al. Ultrahigh-Energy Density Lead-Free Dielectric Films via Polymorphic Nanodomain Design. Science 2019, 365, 578–582. [Google Scholar] [CrossRef] [PubMed]
  22. Qi, H.; Zuo, R.; Xie, A.; Tian, A.; Fu, J.; Zhang, Y.; Zhang, S. Ultrahigh Energy-Storage Density in NaNbO3-Based Lead-Free Relaxor Antiferroelectric Ceramics with Nanoscale Domains. Adv. Funct. Mater. 2019, 29, 1903877. [Google Scholar] [CrossRef]
  23. Pan, H.; Ma, J.; Ma, J.; Zhang, Q.; Liu, X.; Guan, B.; Gu, L.; Zhang, X.; Zhang, Y.-J.; Li, L.; et al. Giant Energy Density and High Efficiency Achieved in Bismuth Ferrite-Based Film Capacitors via Domain Engineering. Nat. Commun. 2018, 9, 1813. [Google Scholar] [CrossRef] [Green Version]
  24. Veerapandiyan, V.; Benes, F.; Gindel, T.; DeLuca, M. Strategies to Improve the Energy Storage Properties of Perovskite Lead-Free Relaxor Ferroelectrics: A Review. Materials 2020, 13, 5742. [Google Scholar] [CrossRef] [PubMed]
  25. Mohapatra, P.; Fan, Z.; Cui, J.; Tan, X. Relaxor Antiferroelectric Ceramics with Ultrahigh Efficiency for Energy Storage Applications. J. Eur. Ceram. Soc. 2019, 39, 4735–4742. [Google Scholar] [CrossRef]
  26. Liu, Z.; Lu, T.; Ye, J.; Wang, G.; Dong, X.; Withers, R.; Liu, Y. Antiferroelectrics for Energy Storage Applications: A Review. Adv. Mater. Technol. 2018, 3, 1800111. [Google Scholar] [CrossRef]
  27. Li, J.; Li, F.; Xu, Z.; Zhang, S. Multilayer Lead-Free Ceramic Capacitors with Ultrahigh Energy Density and Efficiency. Adv. Mater. 2018, 30, 1802155. [Google Scholar] [CrossRef]
  28. Hou, C.; Huang, W.; Zhao, W.; Zhang, D.; Yin, Y.; Li, X. Ultrahigh Energy Density in SrTiO3 Film Capacitors. ACS Appl. Mater. Interfaces 2017, 9, 20484–20490. [Google Scholar] [CrossRef] [PubMed]
  29. Jang, H.W.; Kumar, A.; Denev, S.; Biegalski, M.D.; Maksymovych, P.; Bark, C.W.; Nelson, C.T.; Folkman, C.M.; Baek, S.H.; Balke, N.; et al. Ferroelectricity in Strain-Free SrTiO3 Thin Films. Phys. Rev. Lett. 2010, 104, 197601. [Google Scholar] [CrossRef] [Green Version]
  30. Li, T.; Deng, S.; Liu, H.; Sun, S.; Li, H.; Hu, S.; Liu, S.; Xing, X.; Chen, J. Strong Room-Temperature Ferroelectricity in Strained SrTiO3 Homoepitaxial Film. Adv. Mater. 2021, 33, 2008316. [Google Scholar] [CrossRef]
  31. Huang, K.; Ge, G.; Yan, F.; Shen, B.; Zhai, J. Ultralow Electrical Hysteresis along with High Energy-Storage Density in Lead-Based Antiferroelectric Ceramics. Adv. Electron. Mater. 2020, 6, 1901366. [Google Scholar] [CrossRef]
  32. Wang, H.; Liu, Y.; Yang, T.; Zhang, S. Ultrahigh Energy-Storage Density in Antiferroelectric Ceramics with Field-Induced Multiphase Transitions. Adv. Funct. Mater. 2019, 29, 1807321. [Google Scholar] [CrossRef]
  33. Palneedi, H.; Peddigari, M.; Hwang, G.-T.; Jeong, D.-Y.; Ryu, J. High-Performance Dielectric Ceramic Films for Energy Storage Capacitors: Progress and Outlook. Adv. Funct. Mater. 2018, 28, 1803665. [Google Scholar] [CrossRef]
  34. Dong, X.; Li, X.; Chen, X.; Chen, H.; Sun, C.; Shi, J.; Pang, F.; Zhou, H. High Energy Storage Density and Power Density Achieved Simultaneously in NaNbO3-Based Lead-Free Ceramics via Antiferroelectricity Enhancement. J. Mater. 2020, 7, 629–639. [Google Scholar] [CrossRef]
  35. Zhao, L.; Liu, Q.; Gao, J.; Zhang, S.; Li, J. Lead-Free Antiferroelectric Silver Niobate Tantalate with High Energy Storage Performance. Adv. Mater. 2017, 29, 1701824. [Google Scholar] [CrossRef] [PubMed]
  36. Li, J.; Shen, Z.; Chen, X.; Yang, S.; Zhou, W.; Wang, M.; Wang, L.; Kou, Q.; Liu, Y.; Li, Q.; et al. Grain-Orientation-Engineered Multilayer Ceramic Capacitors for Energy Storage Applications. Nat. Mater. 2020, 19, 999–1005. [Google Scholar] [CrossRef] [PubMed]
  37. Peng, B.; Zhang, Q.; Li, X.; Sun, T.; Fan, H.; Ke, S.; Ye, M.; Wang, Y.; Lu, W.; Niu, H.; et al. Giant Electric Energy Density in Epitaxial Lead-Free Thin Films with Coexistence of Ferroelectrics and Antiferroelectrics. Adv. Electron. Mater. 2015, 1, 1500052. [Google Scholar] [CrossRef] [Green Version]
  38. Pan, H.; Lan, S.; Xu, S.; Zhang, Q.; Yao, H.; Liu, Y.; Meng, F.; Guo, E.-J.; Gu, L.; Yi, D.; et al. Ultrahigh Energy Storage in Superparaelectric Relaxor Ferroelectrics. Science 2021, 374, 100–104. [Google Scholar] [CrossRef] [PubMed]
  39. Acharya, M.; Banyas, E.; Ramesh, M.; Jiang, Y.; Fernandez, A.; Dasgupta, A.; Ling, H.; Hanrahan, B.; Persson, K.; Neaton, J.B.; et al. Exploring the Pb1-xSrxHfO3 System and Potential for High Capacitive Energy Storage Density and Efficiency. Adv. Mater. 2021, 33, 2105967. [Google Scholar] [CrossRef]
  40. Yang, L.; Kong, X.; Li, F.; Hao, H.; Cheng, Z.; Liu, H.; Li, J.-F.; Zhang, S. Perovskite Lead-Free Dielectrics for Energy Storage Applications. Prog. Mater. Sci. 2018, 102, 72–108. [Google Scholar] [CrossRef]
  41. Fan, Q.; Ma, C.; Li, Y.; Liang, Z.; Cheng, S.; Guo, M.; Dai, Y.; Ma, C.; Lu, L.; Wang, W.; et al. Realization of High Energy Density in an Ultra-Wide Temperature Range through Engineering of Ferroelectric Sandwich Structures. Nano Energy 2019, 62, 725–733. [Google Scholar] [CrossRef]
  42. Yang, C.; Lv, P.; Qian, J.; Han, Y.; Ouyang, J.; Lin, X.; Huang, S.; Cheng, Z. Fatigue-Free and Bending-Endurable Flexible Mn-Doped Na0.5Bi0.5TiO3-BaTiO3-BiFeO3 Film Capacitor with an Ultrahigh Energy Storage Performance. Adv. Energy Mater. 2019, 9, 1803949. [Google Scholar] [CrossRef]
  43. Ma, B.; Hu, Z.; Koritala, R.E.; Lee, T.H.; Dorris, S.E.; Balachandran, U. PLZT Film Capacitors for Power Electronics and Energy Storage Applications. J. Mater. Sci. Mater. Electron. 2015, 26, 9279–9287. [Google Scholar] [CrossRef]
  44. Jiang, J.; Meng, X.; Li, L.; Guo, S.; Huang, M.; Zhang, J.; Wang, J.; Hao, X.; Zhu, H.; Zhang, S.-T. Ultrahigh Energy Storage Density in Lead-Free Relaxor Antiferroelectric Ceramics via Domain Engineering. Energy Storage Mater. 2021, 43, 383–390. [Google Scholar] [CrossRef]
  45. Qi, H.; Xie, A.; Tian, A.; Zuo, R. Superior Energy-Storage Capacitors with Simultaneously Giant Energy Density and Efficiency Using Nanodomain Engineered BiFeO3-BaTiO3-NaNbO3 Lead-Free Bulk Ferroelectrics. Adv. Energy Mater. 2019, 10, 1903338. [Google Scholar] [CrossRef]
  46. Luo, N.; Han, K.; Cabral, M.J.; Liao, X.; Zhang, S.; Liao, C.; Zhang, G.; Chen, X.; Feng, Q.; Li, J.-F.; et al. Constructing Phase Boundary in AgNbO3 Antiferroelectrics: Pathway Simultaneously Achieving High Energy Density and Efficiency. Nat. Commun. 2020, 11, 4824. [Google Scholar] [CrossRef] [PubMed]
  47. Ge, G.; Huang, K.; Wu, S.; Yan, F.; Li, X.; Shen, B.; Zhai, J. Synergistic Optimization of Antiferroelectric Ceramics with Superior Energy Storage Properties via Phase Structure Engineering. Energy Storage Mater. 2020, 35, 114–121. [Google Scholar] [CrossRef]
  48. Blue, C.T.; Hicks, J.C.; Park, S.E.; Yoshikawa, S.; Cross, L.E. In Situ X-ray Diffraction Study of the Antiferroelectric-Ferroelectric Phase Transition in PLSnZT. Appl. Phys. Lett. 1996, 68, 2942–2944. [Google Scholar] [CrossRef]
  49. Cai, Y.; Phillipp, F.; Zimmermann, A.; Zhou, L.; Aldinger, F.; Rühle, M. TEM Study of Superstructure in a Perovskite Lead Lanthanum Zirconate Stannate Titanate Ceramic. Acta Mater. 2003, 51, 6429–6436. [Google Scholar] [CrossRef]
  50. He, H.; Tan, X. Electric-Field-Induced Transformation of Incommensurate Modulations in Antiferroelectric Pb0.99Nb0.02[(Zr1-xSnx)1-yTiy]0.98O3. Phys. Rev. B 2005, 72, 024102. [Google Scholar] [CrossRef] [Green Version]
  51. Li, F.; Lin, D.; Chen, Z.; Cheng, Z.; Wang, J.; Li, C.; Xu, Z.; Huang, Q.; Liao, X.; Chen, L.-Q.; et al. Ultrahigh Piezoelectricity in Ferroelectric Ceramics by Design. Nat. Mater. 2018, 17, 349–354. [Google Scholar] [CrossRef]
  52. Tan, X.; Ma, C.; Frederick, J.; Beckman, S.; Webber, K.G. The Antiferroelectric ↔ Ferroelectric Phase Transition in Lead-Containing and Lead-Free Perovskite Ceramics. J. Am. Ceram. Soc. 2011, 94, 4091–4107. [Google Scholar] [CrossRef] [Green Version]
  53. Zhang, T.; Zhao, Y.; Li, W.; Fei, W. High Energy Storage Density at Low Electric Field of ABO3 Antiferroelectric Films with Ionic Pair Doping. Energy Storage Mater. 2018, 18, 238–245. [Google Scholar] [CrossRef]
  54. Liu, X.; Li, Y.; Sun, N.; Hao, X. High Energy-Storage Performance of PLZS Antiferroelectric Multilayer Ceramic Capacitors. Inorg. Chem. Front. 2019, 7, 756–764. [Google Scholar] [CrossRef]
  55. Zhou, H.Y.; Liu, X.Q.; Zhu, X.L.; Chen, X.M. CaTiO3 Linear Dielectric Ceramics with Greatly Enhanced Dielectric Strength and Energy Storage Density. J. Am. Ceram. Soc. 2017, 101, 1999–2008. [Google Scholar] [CrossRef]
  56. Ravel, B.; Slcron, N.; Yacoby, Y.; Stern, E.A.; Dogan, F.; Rehr, J.J. Order-Disorder Behavior in the Phase Transition of PbTiO3. Ferroelectrics 1995, 164, 265–277. [Google Scholar] [CrossRef]
  57. Salmani-Rezaie, S.; Ahadi, K.; Strickland, W.M.; Stemmer, S. Order-Disorder Ferroelectric Transition of Strained SrTiO3. Phys. Rev. Lett. 2020, 125, 087601. [Google Scholar] [CrossRef]
  58. Aksel, E.; Jones, J.L. Advances in Lead-Free Piezoelectric Materials for Sensors and Actuators. Sensors 2010, 10, 1935–1954. [Google Scholar] [CrossRef] [PubMed]
  59. Whatmore, R.W.; Glazer, A.M. Structural Phase Transitions in Lead Zirconate. J. Phys. C Solid State Phys. 1979, 12. [Google Scholar] [CrossRef]
  60. Scott, B.A.; Burns, G. Crystal Growth and Observation of the Ferroelectric Phase of PbZrO3. J. Am. Ceram. Soc. 1972, 55, 331–333. [Google Scholar] [CrossRef]
  61. Kobayashi, K.; Ryu, M.; Doshida, Y.; Mizuno, Y.; Randall, C. Novel High-Temperature Antiferroelectric-Based Dielectric NaNbO3-NaTaO3 Solid Solutions Processed in Low Oxygen Partial Pressures. J. Am. Ceram. Soc. 2012, 96, 531–537. [Google Scholar] [CrossRef]
  62. Sawaguchi, E. Ferroelectricity versus Antiferroelectricity in the Solid Solutions of PbZrO3 and PbTiO3. J. Phys. Soc. Jpn. 1953, 8, 615–629. [Google Scholar] [CrossRef]
  63. Chauhan, A.; Patel, S.; Vaish, R.; Bowen, C.R. Anti-Ferroelectric Ceramics for High Energy Density Capacitors. Materials 2015, 8, 8009–8031. [Google Scholar] [CrossRef] [Green Version]
  64. Chen, D.; Nelson, C.T.; Zhu, X.; Serrao, C.R.; Clarkson, J.D.; Wang, Z.; Gao, Y.; Hsu, S.-L.; Dedon, L.R.; Chen, Z.; et al. A Strain-Driven Antiferroelectric-to-Ferroelectric Phase Transition in La-Doped BiFeO3 Thin Films on Si. Nano Lett. 2017, 17, 5823–5829. [Google Scholar] [CrossRef] [PubMed]
  65. Borisevich, A.; Eliseev, E.; Morozovska, A.N.; Cheng, C.-J.; Lin, J.-Y.; Chu, Y.-H.; Kan, D.; Takeuchi, I.; Nagarajan, V.; Kalinin, S. Atomic-Scale Evolution of Modulated Phases at the Ferroelectric-Antiferroelectric Morphotropic Phase Boundary Controlled by Flexoelectric Interaction. Nat. Commun. 2012, 3, 775. [Google Scholar] [CrossRef] [PubMed]
  66. Yamasaki, K.; Soejima, Y.; Fischer, K.F. Superstructure Determination of PbZrO3. Acta Crystallogr. Sect. B Struct. Sci. 1998, 54, 524–530. [Google Scholar] [CrossRef]
  67. Lu, Z.; Bao, W.; Wang, G.; Sun, S.-K.; Li, L.; Li, J.; Yang, H.; Ji, H.; Feteira, A.; Li, D.; et al. Mechanism of Enhanced Energy Storage Density in AgNbO3-Based Lead-Free Antiferroelectrics. Nano Energy 2020, 79, 105423. [Google Scholar] [CrossRef]
  68. Xie, A.; Qi, H.; Zuo, R.; Tian, A.; Chen, J.; Zhang, S. An Environmentally-Benign NaNbO3 Based Perovskite Antiferroelectric Alternative to Traditional Lead-Based Counterparts. J. Mater. Chem. C 2019, 7, 15153–15161. [Google Scholar] [CrossRef]
  69. Xu, X.; Huang, F.-T.; Qi, Y.; Singh, S.; Rabe, K.M.; Obeysekera, D.; Yang, J.; Chu, M.-W.; Cheong, S.-W. Kinetically Stabilized Ferroelectricity in Bulk Single-Crystalline HfO2:Y. Nat. Mater. 2021, 20, 826–832. [Google Scholar] [CrossRef]
  70. Silva, J.P.B.; Sekhar, K.C.; Pan, H.; MacManus-Driscoll, J.L.; Pereira, M. Advances in Dielectric Thin Films for Energy Storage Applications, Revealing the Promise of Group IV Binary Oxides. ACS Energy Lett. 2021, 6, 2208–2217. [Google Scholar] [CrossRef]
  71. Tsai, M.; Zheng, Y.; Lu, S.; Zheng, J.; Pan, H.; Duan, C.; Yu, P.; Huang, R.; Chu, Y. Antiferroelectric Anisotropy of Epitaxial PbHfO3 Films for Flexible Energy Storage. Adv. Funct. Mater. 2021, 31, 2105060. [Google Scholar] [CrossRef]
  72. Ko, D.-L.; Hsin, T.; Lai, Y.-H.; Ho, S.-Z.; Zheng, Y.; Huang, R.; Pan, H.; Chen, Y.-C.; Chu, Y.-H. High-Stability Transparent Flexible Energy Storage Based on PbZrO3/Muscovite Heterostructure. Nano Energy 2021, 87, 106149. [Google Scholar] [CrossRef]
  73. Xu, C.; Chen, Y.; Cai, X.; Meingast, A.; Guo, X.; Wang, F.; Lin, Z.; Lo, T.W.; Maunders, C.; Lazar, S.; et al. Two-Dimensional Antiferroelectricity in Nanostripe-Ordered In2Se3. Phys. Rev. Lett. 2020, 125, 047601. [Google Scholar] [CrossRef]
  74. Xu, C.; Mao, J.; Guo, X.; Yan, S.; Chen, Y.; Lo, T.W.; Chen, C.; Lei, D.; Luo, X.; Hao, J.; et al. Two-Dimensional Ferroelasticity in van der Waals Beta′-In2Se3. Nat. Commun. 2021, 12, 3665. [Google Scholar] [CrossRef] [PubMed]
  75. Corker, D.L.; Glazer, A.M.; Dec, J.; Roleder, K.; Whatmore, R. A Re-investigation of the Crystal Structure of the Perovskite PbZrO3 by X-ray and Neutron Diffraction. Acta Crystallogr. Sect. B Struct. Sci. 1997, 53, 135–142. [Google Scholar] [CrossRef]
  76. Mishra, S.K.; Choudhury, N.; Chaplot, S.L.; Krishna, P.S.R.; Mittal, R. Competing Antiferroelectric and Ferroelectric Interactions in NaNbO3: Neutron Diffraction and Theoretical Studies. Phys. Rev. B 2007, 76, 024110. [Google Scholar] [CrossRef] [Green Version]
  77. Shimizu, H.; Guo, H.; Reyes-Lillo, S.E.; Mizuno, Y.; Rabe, K.M.; Randall, C.A. Lead-Free Antiferroelectric: xCaZrO3-(1-x)NaNbO3 system (0 ≤ x ≤ 0.10). Dalton Trans. 2015, 44, 10763. [Google Scholar] [CrossRef] [PubMed]
  78. Kagimura, R.; Singh, D.J. First-Principles Investigations of Elastic Properties and Energetics of Antiferroelectric and Ferroelectric Phases of PbZrO3. Phys. Rev. B 2008, 77, 104113. [Google Scholar] [CrossRef]
  79. Wei, X.-K.; Tagantsev, A.K.; Kvasov, A.; Roleder, K.; Jia, C.-L.; Setter, N. Ferroelectric Translational Antiphase Boundaries in Nonpolar Materials. Nat. Commun. 2014, 5, 3031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Wei, X.-K.; Jia, C.-L.; Roleder, K.; Setter, N. Polarity of Translation Boundaries in Antiferroelectric PbZrO3. Mater. Res. Bull. 2015, 62, 101–105. [Google Scholar] [CrossRef]
  81. Wei, X.-K.; Vaideeswaran, K.; Sandu, C.S.; Jia, C.-L.; Setter, N. Preferential Creation of Polar Translational Boundaries by Interface Engineering in Antiferroelectric PbZrO3Thin Films. Adv. Mater. Interfaces 2015, 2, 1500349. [Google Scholar] [CrossRef]
  82. Mani, B.K.; Chang, C.M.; Lisenkov, S.; Ponomareva, I. Critical Thickness for Antiferroelectricity in PbZrO3. Phys. Rev. Lett. 2015, 115, 097601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Chaudhuri, A.R.; Arredondo, M.; Hähnel, A.; Morelli, A.; Becker, M.; Alexe, M.; Vrejoiu, I. Epitaxial Strain Stabilization of a Ferroelectric Phase in PbZrO3 Thin films. Phys. Rev. B 2011, 84, 054112. [Google Scholar] [CrossRef] [Green Version]
  84. MacLaren, I.; Villaurrutia, R.; Schaffer, B.; Houben, L.; Peláiz-Barranco, A. Atomic-Scale Imaging and Quantification of Electrical Polarisation in Incommensurate Antiferroelectric Lanthanum-Doped Lead Zirconate Titanate. Adv. Funct. Mater. 2011, 22, 261–266. [Google Scholar] [CrossRef]
  85. Fu, Z.; Chen, X.; Li, Z.; Hu, T.; Zhang, L.; Lu, P.; Zhang, S.; Wang, G.; Dong, X.; Xu, F. Unveiling the Ferrielectric Nature of PbZrO3-Based Antiferroelectric Materials. Nat. Commun. 2020, 11, 3809. [Google Scholar] [CrossRef]
  86. Ma, T.; Fan, Z.; Xu, B.; Kim, T.-H.; Lu, P.; Bellaiche, L.; Kramer, M.J.; Tan, X.; Zhou, L. Uncompensated Polarization in Incommensurate Modulations of Perovskite Antiferroelectrics. Phys. Rev. Lett. 2019, 123, 217602. [Google Scholar] [CrossRef] [Green Version]
  87. Carpenter, M.A.; Salje, E.K.H.; Graeme-Barber, A. Spontaneous Strain as a Determinant of Thermodynamic Properties for Phase Transitions in Minerals. Eur. J. Miner. 1998, 10, 621–691. [Google Scholar] [CrossRef]
  88. Puchberger, S.; Soprunyuk, V.; Schranz, W. Diverging Relaxation Times of Domain Wall Motion Indicating Glassy Dynamics in Ferroelastics. Mater. Res. 2018, 21, e20170842. [Google Scholar] [CrossRef]
  89. Salje, E.K. Phase Transitions in Ferroelastic and Co-Elastic Crystals; Cambridge University Press: Cambridge, UK, 1993; ISBN 0521429366. [Google Scholar]
  90. Watanabe, S.; Hidaka, M.; Yoshizawa, H.; Wanklyn, B.M. Antiferroelastic Structural Transitions in PrAlO3 by Means of Neutron Diffraction. Phys. Status Solidi B 2005, 243, 424–434. [Google Scholar] [CrossRef]
  91. Lines, M.E.; Glass, A.M. Principles and Applications of Ferroelectrics and Related Materials; Oxford University Press: Oxford, UK, 1977. [Google Scholar]
  92. Glazer, A.M. The Classification of Tilted Octahedra in Perovskites. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1972, 28, 3384–3392. [Google Scholar] [CrossRef]
  93. Woodward, D.I.; Reaney, I. Electron Diffraction of Tilted Perovskites. Acta Crystallogr. Sect. B Struct. Sci. 2005, 61, 387–399. [Google Scholar] [CrossRef]
  94. Karpinsky, D.V.; Eliseev, E.; Xue, F.; Silibin, M.V.; Franz, A.; Glinchuk, M.; Troyanchuk, I.O.; Gavrilov, S.; Gopalan, V.; Chen, L.-Q.; et al. Thermodynamic Potential and Phase Diagram for Multiferroic Bismuth Ferrite (BiFeO3). NPJ Comput. Mater. 2017, 3, 20. [Google Scholar] [CrossRef]
  95. Gu, T.; Scarbrough, T.; Yang, Y.; Íñiguez, J.; Bellaiche, L.; Xiang, H.J. Cooperative Couplings between Octahedral Rotations and Ferroelectricity in Perovskites and Related Materials. Phys. Rev. Lett. 2018, 120, 197602. [Google Scholar] [CrossRef] [Green Version]
  96. Son, J.Y.; Lee, G.; Jo, M.-H.; Kim, H.; Jang, H.M.; Shin, Y.-H. Heteroepitaxial Ferroelectric ZnSnO3 Thin Film. J. Am. Chem. Soc. 2009, 131, 8386–8387. [Google Scholar] [CrossRef] [PubMed]
  97. Xu, B.; Iñiguez, J.; Bellaiche, L. Designing Lead-Free Antiferroelectrics for Energy Storage. Nat. Commun. 2017, 8, 15682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Kim, J.R.; Jang, J.; Go, K.-J.; Park, S.Y.; Roh, C.J.; Bonini, J.; Kim, J.; Lee, H.G.; Rabe, K.M.; Lee, J.S.; et al. Stabilizing Hidden Room-Temperature Ferroelectricity via a Metastable Atomic Distortion Pattern. Nat. Commun. 2020, 11, 4944. [Google Scholar] [CrossRef]
  99. Li, Z.; Song, D.; Yu, R.; Ge, B.; Liao, Z.; Li, Y.; Dong, S.; Zhu, J. Competing Interfacial Reconstruction Mechanisms in La0.7Sr0.3MnO3/SrTiO3 Heterostructures. ACS Appl. Mater. Interfaces 2016, 8, 24192–24197. [Google Scholar] [CrossRef]
  100. El-Mellouhi, F.; Brothers, E.N.; Lucero, M.J.; Scuseria, G.E. Modeling of the Cubic and Antiferrodistortive Phases of SrTiO3 with Screened Hybrid Density Functional Theory. Phys. Rev. B 2011, 84, 115122. [Google Scholar] [CrossRef] [Green Version]
  101. Li, Y.L.; Choudhury, S.; Haeni, J.H.; Biegalski, M.D.; Vasudevarao, A.; Sharan, A.; Ma, H.Z.; Levy, J.; Gopalan, V.; Trolier-McKinstry, S.; et al. Phase Transitions and Domain Structures in Strained Pseudocubic (100) SrTiO3 Thin Films. Phys. Rev. B 2006, 73, 184112. [Google Scholar] [CrossRef]
  102. Nova, T.F.; Disa, A.S.; Fechner, M.; Cavalleri, A. Metastable Ferroelectricity in Optically Strained SrTiO3. Science 2019, 364, 1075–1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Haeni, J.H.; Irvin, P.; Chang, W.; Uecker, R.; Reiche, P.; Li, Y.; Choudhury, S.; Tian, W.; Hawley, M.E.; Craigo, B.; et al. Room-Temperature Ferroelectricity in Strained SrTiO3. Nature 2004, 430, 758–761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Xu, R.; Huang, J.; Barnard, E.S.; Hong, S.S.; Singh, P.; Wong, E.K.; Jansen, T.; Harbola, V.; Xiao, J.; Wang, B.Y.; et al. Strain-Induced Room-Temperature Ferroelectricity in SrTiO3 Membranes. Nat. Commun. 2020, 11, 3141. [Google Scholar] [CrossRef] [PubMed]
  105. Weaver, H. Dielectric Properties of Single Crystals of SrTiO3 at Low Temperatures. J. Phys. Chem. Solids 1959, 11, 274–277. [Google Scholar] [CrossRef]
  106. Tagantsev, A.K.; Courtens, E.; Arzel, L. Prediction of a Low-Temperature Ferroelectric Instability in Antiphase Domain Boundaries of Strontium Titanate. Phys. Rev. B 2001, 64, 224107. [Google Scholar] [CrossRef]
  107. Kvasov, A.; Tagantsev, A.K.; Setter, N. Structure and Pressure-Induced Ferroelectric Phase Transition in Antiphase Domain Boundaries of Strontium Titanate from First Principles. Phys. Rev. B 2016, 94, 054102. [Google Scholar] [CrossRef]
  108. Li, X.; Qiu, T.; Zhang, J.; Baldini, E.; Lu, J.; Rappe, A.M.; Nelson, K.A. Terahertz Field-Induced Ferroelectricity in Quantum Paraelectric SrTiO3. Science 2019, 364, 1079–1082. [Google Scholar] [CrossRef] [Green Version]
  109. Lee, D.; Lu, H.; Gu, Y.; Choi, S.-Y.; Li, S.-D.; Ryu, S.; Paudel, T.R.; Song, K.; Mikheev, E.; Lee, S.; et al. Emergence of Room-Temperature Ferroelectricity at Reduced Dimensions. Science 2015, 349, 1314–1317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Klyukin, K.; Alexandrov, V. Effect of Intrinsic Point Defects on Ferroelectric Polarization Behavior of SrTiO3. Phys. Rev. B 2017, 95, 035301. [Google Scholar] [CrossRef] [Green Version]
  111. Zhang, X.; Shen, Y.; Xu, B.; Zhang, Q.; Gu, L.; Jiang, J.; Ma, J.; Lin, Y.; Nan, C.-W. Giant Energy Density and Improved Discharge Efficiency of Solution-Processed Polymer Nanocomposites for Dielectric Energy Storage. Adv. Mater. 2016, 28, 2055–2061. [Google Scholar] [CrossRef] [PubMed]
  112. Kužel, P.; Kadlec, F.; Petzelt, J.; Schubert, J.; Panaitov, G. Highly Tunable SrTiO3/DyScO3 Heterostructures for Applications in the Terahertz Range. Appl. Phys. Lett. 2007, 91, 232911. [Google Scholar] [CrossRef]
  113. Sluka, T.; Tagantsev, A.K.; Damjanovic, A.; Gureev, M.; Setter, N. Enhanced Electromechanical Response of Ferroelectrics Due to Charged Domain Walls. Nat. Commun. 2012, 3, 748. [Google Scholar] [CrossRef] [Green Version]
  114. Lisenkov, S.; Yao, Y.; Bassiri-Gharb, N.; Ponomareva, I. Prediction of High-Strain Polar Phases in Antiferroelectric PbZrO3 from a Multiscale Approach. Phys. Rev. B 2020, 102, 104101. [Google Scholar] [CrossRef]
  115. Ge, J.; Rémiens, D.; Costecalde, J.; Chen, Y.; Dong, X.; Wang, G. Effect of Residual Stress on Energy Storage Property in PbZrO3 Antiferroelectric Thin Films with Different Orientations. Appl. Phys. Lett. 2013, 103, 162903. [Google Scholar] [CrossRef]
  116. Chandrasekaran, A.; Wei, X.-K.; Feigl, L.; Damjanovic, D.; Setter, N.; Marzari, N. Asymmetric Structure of 90° Domain Walls and Interactions with Defects in PbTiO3. Phys. Rev. B 2016, 93, 144102. [Google Scholar] [CrossRef]
  117. Wei, X.-K.; Sluka, T.; Fraygola, B.; Feigl, L.; Du, H.; Jin, L.; Jia, C.-L.; Setter, N. Controlled Charging of Ferroelastic Domain Walls in Oxide Ferroelectrics. ACS Appl. Mater. Interfaces 2017, 9, 6539–6546. [Google Scholar] [CrossRef]
  118. Wei, X.K.; Jia, C.L.; Roleder, K.; Dunin-Borkowski, R.E.; Mayer, J. In Situ Observation of Point-Defect-Induced Unit-Cell-Wise Energy Storage Pathway in Antiferroelectric PbZrO3. Adv. Funct. Mater. 2021, 31, 2008609. [Google Scholar] [CrossRef]
  119. Zhang, L.; Fu, Z.; Chen, X.; Li, Z.; Hu, T.; Yu, Z.; Wang, G.; Dong, X.; Xu, F. Chemically Tunable Textured Interfacial Defects in PbZrO3-Based Antiferroelectric Perovskite Oxides. Chem. Mater. 2021, 33, 6743–6751. [Google Scholar] [CrossRef]
  120. Luo, N.; Han, K.; Zhuo, F.; Xu, C.; Zhang, G.; Liu, L.; Chen, X.; Hu, C.; Zhou, H.; Wei, Y. Aliovalent A-Site Engineered AgNbO3 Lead-Free Antiferroelectric Ceramics toward Superior Energy Storage Density. J. Mater. Chem. A 2019, 7, 14118–14128. [Google Scholar] [CrossRef]
  121. Liu, H.; Fan, L.; Sun, S.; Lin, K.; Ren, Y.; Tan, X.; Xing, X.; Chen, J. Electric-Field-Induced Structure and Domain Texture Evolution in PbZrO3-Based Antiferroelectric by In-Situ High-Energy Synchrotron X-ray Diffraction. Acta Mater. 2019, 184, 41–49. [Google Scholar] [CrossRef]
  122. Lu, T.; Studer, A.J.; Noren, L.; Hu, W.; Yu, D.; McBride, B.; Feng, Y.; Withers, R.; Chen, H.; Xu, Z.; et al. Electric-Field-Induced AFE-FE Transitions and Associated Strain/Preferred Orientation in Antiferroelectric PLZST. Sci. Rep. 2016, 6, 23659. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, J.; Liu, Y.; Withers, R.L.; Studer, A.; Li, Q.; Norén, L.; Guo, Y. A Correlated Electron Diffraction, in Situ Neutron Diffraction and Dielectric Properties Investigation of Poled (1-x)Bi0.5Na0.5TiO3-xBaTiO3 Ceramics. J. Appl. Phys. 2011, 110, 084114. [Google Scholar] [CrossRef] [Green Version]
  124. Ciuchi, I.V.; Chung, C.C.; Fancher, C.M.; Guerrier, J.; Forrester, J.S.; Jones, J.L.; Mitoseriu, L.; Galassi, C. Field-Induced Antiferroelectric to Ferroelectric Transitions in (Pb1xLax)(Zr0.90Ti0.10)1−x/4O3 investigated by in Situ X-ray Diffraction. J. Europ. Ceram. Soc. 2017, 37, 4631–4636. [Google Scholar] [CrossRef]
  125. Fan, Z.; Xue, F.; Tutuncu, G.; Chen, L.-Q.; Tan, X. Interaction Dynamics between Ferroelectric and Antiferroelectric Domains in a PbZrO3-Based Ceramic. Phys. Rev. Appl. 2019, 11, 064050. [Google Scholar] [CrossRef] [Green Version]
  126. Zhang, M.-H.; Zhao, C.; Fulanović, L.; Rödel, J.; Novak, N.; Schökel, A.; Koruza, J. Revealing the Mechanism of Electric-Field-Induced Phase Transition in Antiferroelectric NaNbO3 by in Situ High-Energy X-ray Diffraction. Appl. Phys. Lett. 2021, 118, 132903. [Google Scholar] [CrossRef]
  127. Wei, X.; Jia, C.; Du, H.; Roleder, K.; Mayer, J.; Dunin-Borkowski, R. An Unconventional Transient Phase with Cycloidal Order of Polarization in Energy-Storage Antiferroelectric PbZrO3. Adv. Mater. 2020, 32, e1907208. [Google Scholar] [CrossRef] [Green Version]
  128. Jia, C.L.; Lentzen, M.; Urban, K. Atomic-Resolution Imaging of Oxygen in Perovskite Ceramics. Science 2003, 299, 870–873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Wei, X.-K.; Jia, C.-L.; Sluka, T.; Wang, B.-X.; Ye, Z.-G.; Setter, N. Néel-like Domain Walls in Ferroelectric Pb(Zr,Ti)O3 Single Crystals. Nat. Commun. 2016, 7, 12385. [Google Scholar] [CrossRef] [PubMed]
  130. Urban, K.W.; Jia, C.-L.; Houben, L.; Lentzen, M.; Mi, S.-B.; Tillmann, K. Negative Spherical Aberration Ultrahigh-Resolution Imaging in Corrected Transmission Electron Microscopy. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2009, 367, 3735–3753. [Google Scholar] [CrossRef]
  131. Findlay, S.; Shibata, N.; Sawada, H.; Okunishi, E.; Kondo, Y.; Ikuhara, Y. Dynamics of Annular Bright Field Imaging in Scanning Transmission Electron Microscopy. Ultramicroscopy 2010, 110, 903–923. [Google Scholar] [CrossRef] [PubMed]
  132. Lazić, I.; Bosch, E.G.; Lazar, S. Phase contrast STEM for Thin Samples: Integrated Differential Phase Contrast. Ultramicroscopy 2016, 160, 265–280. [Google Scholar] [CrossRef] [PubMed]
  133. Yücelen, E.; Lazić, I.; Bosch, E.G.T. Phase Contrast Scanning Transmission Electron Microscopy Imaging of Light and Heavy Atoms at the Limit of Contrast and Resolution. Sci. Rep. 2018, 8, 2676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Chen, Z.; Jiang, Y.; Shao, Y.-T.; Holtz, M.E.; Odstrčil, M.; Guizar-Sicairos, M.; Hanke, I.; Ganschow, S.; Schlom, D.G.; Muller, D.A. Electron Ptychography Achieves Atomic-Resolution Limits Set by Lattice Vibrations. Science 2021, 372, 826–831. [Google Scholar] [CrossRef] [PubMed]
  135. Chen, Q.; Dwyer, C.; Sheng, G.; Zhu, C.; Li, X.; Zheng, C.; Zhu, Y. Imaging Beam-Sensitive Materials by Electron Microscopy. Adv. Mater. 2020, 32, e1907619. [Google Scholar] [CrossRef]
  136. Russo, C.J.; Henderson, R. Charge Accumulation in Electron Cryomicroscopy. Ultramicroscopy 2018, 187, 43–49. [Google Scholar] [CrossRef]
  137. Yao, L.; Majumdar, S.; Akaslompolo, L.; Inkinen, S.; Qin, Q.H.; van Dijken, S. Electron-Beam-Induced Perovskite-Brownmillerite-Perovskite Structural Phase Transitions in Epitaxial La2/3Sr1/3MnO3 Films. Adv. Mater. 2014, 26, 2789–2793. [Google Scholar] [CrossRef]
  138. Zheng, H.; Rivest, J.B.; Miller, T.A.; Sadtler, B.; Lindenberg, A.; Toney, M.F.; Wang, L.-W.; Kisielowski, C.; Alivisatos, A.P. Observation of Transient Structural-Transformation Dynamics in a Cu2S Nanorod. Science 2011, 333, 206–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Liu, H.; Zhou, Z.; Qiu, Y.; Gao, B.; Sun, S.; Lin, K.; Ding, L.; Li, Q.; Cao, Y.; Ren, Y.; et al. An Intriguing Intermediate State as a Bridge between Antiferroelectric and Ferroelectric Perovskites. Mater. Horiz. 2020, 7, 1912–1918. [Google Scholar] [CrossRef]
  140. Bittner, R.; Humer, K.; Weber, H.W.; Kundzins, K.; Sternberg, A.; Lesnyh, D.A.; Kulikov, D.V.; Trushin, Y.V. Oxygen Vacancy Defects in Antiferroelectric PbZrO3 Thin Film Heterostructures after Neutron Irradiation. J. Appl. Phys. 2004, 96, 3239–3246. [Google Scholar] [CrossRef]
  141. Gao, R.; Reyes-Lillo, S.E.; Xu, R.; Dasgupta, A.; Dong, Y.; Dedon, L.R.; Kim, J.; Saremi, S.; Chen, Z.; Serrao, C.R.; et al. Ferroelectricity in Pb1+δZrO3 Thin Films. Chem. Mater. 2017, 29, 6544–6551. [Google Scholar] [CrossRef] [Green Version]
  142. Bednyakov, P.S.; Sturman, B.I.; Sluka, T.; Tagantsev, A.K.; Yudin, P.V. Physics and Applications of Charged Domain Walls. NPJ Comput. Mater. 2018, 4, 4. [Google Scholar] [CrossRef] [Green Version]
  143. Sluka, T.; Tagantsev, A.K.; Bednyakov, P.; Setter, N. Free-Electron Gas at Charged Domain Walls in Insulating BaTiO3. Nat. Commun. 2013, 4, 1808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Hao, X.; Zhai, J.; Zhou, J.; Yue, Z.; Yang, J.; Zhao, W.; An, S. Structure and Electrical Properties of PbZrO3 Antiferroelectric Thin Films Doped with Barium and Strontium. J. Alloy. Compd. 2011, 509, 271–275. [Google Scholar] [CrossRef]
  145. Zhao, L.; Gao, J.; Liu, Q.; Zhang, S.; Li, J.-F. Silver Niobate Lead-Free Antiferroelectric Ceramics: Enhancing Energy Storage Density by B-Site Doping. ACS Appl. Mater. Interfaces 2017, 10, 819–826. [Google Scholar] [CrossRef] [PubMed]
  146. Luo, N.; Han, K.; Zhuo, F.; Liu, L.; Chen, X.; Peng, B.; Wang, X.; Feng, Q.; Wei, Y. Design for High Energy Storage Density and Temperature-Insensitive Lead-Free Antiferroelectric Ceramics. J. Mater. Chem. C 2019, 7, 4999–5008. [Google Scholar] [CrossRef]
  147. Chmaissem, O.; Dabrowski, B.; Kolesnik, S.; Mais, J.; Jorgensen, J.D.; Short, S. Structural and Magnetic Phase Diagrams of La1−xSrxMnO3 and Pr1−ySryMnO3. Phys. Rev. B 2003, 67, 094431. [Google Scholar] [CrossRef]
  148. Terasaki, I.; Ikuta, M.; Yamamoto, T.D.; Taniguchi, H. Impurity-Induced Spin-State Crossover in La0.8Sr0.2Co1−xAlxO3. Crystals 2018, 8, 411. [Google Scholar] [CrossRef] [Green Version]
  149. Allen, P.B.; Berger, H.; Chauvet, O.; Forro, L.; Jarlborg, T.; Junod, A.; Revaz, B.; Santi, G. Transport Properties, Thermodynamic Properties, and Electronic Structure of SrRuO3. Phys. Rev. B 1996, 53, 4393. [Google Scholar] [CrossRef] [Green Version]
  150. Liao, Z.; Li, F.; Gao, P.; Li, L.; Guo, J.; Pan, X.; Jin, R.; Plummer, E.W.; Zhang, J. Origin of the Metal-Insulator Transition in Ultrathin Films of La2/3Sr2/3MnO3. Phys. Rev. B 2015, 92, 125123. [Google Scholar] [CrossRef] [Green Version]
  151. Yao, L.; Inkinen, S.; van Dijken, S. Direct Observation of Oxygen Vacancy-Driven Structural and Resistive Phase Transitions in La2/3Sr1/3MnO3. Nat. Commun. 2017, 8, 14544. [Google Scholar] [CrossRef] [PubMed]
  152. Stengel, M.; Spaldin, N.A. Origin of the Dielectric Dead Layer in Nanoscale Capacitors. Nature 2006, 443, 679–682. [Google Scholar] [CrossRef]
  153. Stengel, M.; Nicola, D.V.; Spaldin, A. Enhancement of Ferroelectricity at Metal-Oxide Interfaces. Nat. Mater. 2009, 8, 392. [Google Scholar] [CrossRef] [Green Version]
  154. Wei, X.-K.; Yang, Y.; McGilly, L.J.; Feigl, L.; Dunin-Borkowski, R.E.; Jia, C.-L.; Bellaiche, L.; Setter, N. Flexible Polarization Rotation at the Ferroelectric/Metal Interface as a Seed for Domain Nucleation. Phys. Rev. B 2018, 98, 020102. [Google Scholar] [CrossRef] [Green Version]
  155. Vailionis, A.; Boschker, H.; Siemons, W.; Houwman, E.P.; Blank, D.H.A.; Rijnders, G.; Koster, G. Misfit Strain Accommodation in Epitaxial ABO3 Perovskites: Lattice Rotations and Lattice Modulations. Phys. Rev. B 2011, 83, 064101. [Google Scholar] [CrossRef] [Green Version]
  156. Parsons, T.G.; D’Hondt, H.; Hadermann, J.; Hayward, M. Synthesis and Structural Characterization of La1−xAxMnO2.5 (A = Ba, Sr, Ca) Phases: Mapping the Variants of the Brownmillerite Structure. Chem. Mater. 2009, 21, 5527–5538. [Google Scholar] [CrossRef]
  157. Lee, H.-S.; Park, H.-H. Band Structure Analysis of La0.7Sr0.3MnO3 Perovskite Manganite Using a Synchrotron. Adv. Condens. Matter Phys. 2015, 2015, 746475. [Google Scholar] [CrossRef] [Green Version]
  158. Nukala, P.; Ahmadi, M.; Wei, Y.; de Graaf, S.; Stylianidis, E.; Chakrabortty, T.; Matzen, S.; Zandbergen, H.W.; Björling, A.; Mannix, D.; et al. Reversible Oxygen Migration and Phase Transitions in Hafnia-Based Ferroelectric Devices. Science 2021, 372, 630–635. [Google Scholar] [CrossRef] [PubMed]
  159. Bauer, U.; Yao, L.; Tan, A.J.; Agrawal, P.; Emori, S.; Tuller, H.L.; Van Dijken, S.; Beach, G. Magneto-Ionic Control of Interfacial Magnetism. Nat. Mater. 2014, 14, 174–181. [Google Scholar] [CrossRef] [PubMed]
  160. Li, M.; Pietrowski, M.J.; De Souza, R.A.; Zhang, H.; Reaney, I.; Cook, S.N.; Kilner, J.; Sinclair, D.C. A Family of Oxide Ion Conductors Based on the Ferroelectric Perovskite Na0.5Bi0.5TiO3. Nat. Mater. 2013, 13, 31–35. [Google Scholar] [CrossRef] [PubMed]
  161. Kim, T.H.; Puggioni, D.; Yuan, Y.; Xie, L.; Zhou, H.; Campbell, N.; Ryan, P.J.; Choi, Y.; Kim, J.-W.; Patzner, J.; et al. Polar Metals by Geometric Design. Nature 2016, 533, 68–72. [Google Scholar] [CrossRef]
  162. Puggioni, D.; Rondinelli, J.M. Designing a Robustly Metallic Noncenstrosymmetric Ruthenate Oxide with Large Thermopower Anisotropy. Nat. Commun. 2014, 5, 3432. [Google Scholar] [CrossRef] [Green Version]
  163. Wei, X.; Bihlmayer, G.; Zhou, X.; Feng, W.; Kolen’Ko, Y.V.; Xiong, D.; Liu, L.; Blügel, S.; Dunin-Borkowski, R.E. Discovery of Real-Space Topological Ferroelectricity in Metallic Transition Metal Phosphides. Adv. Mater. 2020, 32, e2003479. [Google Scholar] [CrossRef]
  164. Ali, F.; Liu, X.; Zhou, D.; Yang, X.; Xu, J.; Schenk, T.; Müller, J.; Schroeder, U.; Cao, F.; Dong, X. Silicon-Doped Hafnium Oxide Anti-Ferroelectric Thin Films for Energy Storage. J. Appl. Phys. 2017, 122, 144105. [Google Scholar] [CrossRef]
  165. Park, M.H.; Kim, H.J.; Kim, Y.J.; Moon, T.; Kim, K.D.; Hwang, C.S. Thin HfxZr1−xO2 Films: A New Lead-Free System for Electrostatic Supercapacitors with Large Energy Storage Density and Robust Thermal Stability. Adv. Energy Mater. 2014, 4, 1400610. [Google Scholar] [CrossRef]
  166. Yuan, Q.; Yao, F.-Z.; Cheng, S.-D.; Wang, L.; Wang, Y.; Mi, S.-B.; Wang, Q.; Wang, X.; Wang, H. Bioinspired Hierarchically Structured All-Inorganic Nanocomposites with Significantly Improved Capacitive Performance. Adv. Funct. Mater. 2020, 30. [Google Scholar] [CrossRef]
  167. Wei, X.-K.; Prokhorenko, S.; Wang, B.-X.; Liu, Z.; Xie, Y.-J.; Nahas, Y.; Jia, C.-L.; Dunin-Borkowski, R.E.; Mayer, J.; Bellaiche, L.; et al. Ferroelectric Phase-Transition Frustration near a Tricritical Composition Point. Nat. Commun. 2021, 12, 5322. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Log-scale Ragone plot showing a performance comparison between different energy storage devices, including superconducting magnetic energy storage (SMES). Conceptually, the energy density (Wh/kg, vertical axis) describes how much energy is available, while the power density (W/kg, horizontal axis) shows how quickly the energy can be delivered. The sloping lines indicate the required time to get the charge in or out of a device [40].
Figure 1. Log-scale Ragone plot showing a performance comparison between different energy storage devices, including superconducting magnetic energy storage (SMES). Conceptually, the energy density (Wh/kg, vertical axis) describes how much energy is available, while the power density (W/kg, horizontal axis) shows how quickly the energy can be delivered. The sloping lines indicate the required time to get the charge in or out of a device [40].
Materials 14 07854 g001
Figure 2. Dielectric constant (εr), P-E loop and structural phases of PbZrO3. (a,b) Temperature-dependent εr and hysteresis loop of AFEs. The green and blue shaded areas in (b) denote recoverable (Jrec) and loss (Jloss) energy during the AFE-to-FE transition. (c) Unit cell of the cubic PbZrO3 phase (space group Pm 3 ¯ m). (d–f) Antiparallel Pb displacements (Σ mode), polar displacement (Γ mode) and oxygen octahedral rotations (R mode) in AFE PbZrO3 (space group Pbam), respectively. The dotted lines show the cubic crystallographic axes in the ab plane [10].
Figure 2. Dielectric constant (εr), P-E loop and structural phases of PbZrO3. (a,b) Temperature-dependent εr and hysteresis loop of AFEs. The green and blue shaded areas in (b) denote recoverable (Jrec) and loss (Jloss) energy during the AFE-to-FE transition. (c) Unit cell of the cubic PbZrO3 phase (space group Pm 3 ¯ m). (d–f) Antiparallel Pb displacements (Σ mode), polar displacement (Γ mode) and oxygen octahedral rotations (R mode) in AFE PbZrO3 (space group Pbam), respectively. The dotted lines show the cubic crystallographic axes in the ab plane [10].
Materials 14 07854 g002
Figure 3. Structural phase transitions of a centrosymmetric perovskite oxide (ABO3). (a,c) Antiferrodistortive (AFD) and ferrodistortive (FD) transitions caused by temperature or stress, respectively. Out-of-phase or in-phase octahedral rotation is linked in (a) but can be unlinked in (c) due to polar distortion. (b,d) Antiferroelectric (AFE) and ferroelectric (FE) transitions caused by temperature and electric field E, respectively. Associated with symmetry lowering, polar octahedra in an FD phase may respond to applied electric field together with A-site cations.
Figure 3. Structural phase transitions of a centrosymmetric perovskite oxide (ABO3). (a,c) Antiferrodistortive (AFD) and ferrodistortive (FD) transitions caused by temperature or stress, respectively. Out-of-phase or in-phase octahedral rotation is linked in (a) but can be unlinked in (c) due to polar distortion. (b,d) Antiferroelectric (AFE) and ferroelectric (FE) transitions caused by temperature and electric field E, respectively. Associated with symmetry lowering, polar octahedra in an FD phase may respond to applied electric field together with A-site cations.
Materials 14 07854 g003
Figure 4. Structural phase transition and FE properties of SrTiO3. (a) Temperature–strain phase diagram showing an AFD transition and a further FE transition [102]. (b,c) Piezoresponse force microscopy (PFM) amplitude and phase images of strain-free SrTiO3 films recorded at 50 K at biasing voltages of ±10 V [29]. (d) Comparisons of Pr and maximum polarization (Pmax) between Ti/O-deficient SrTiO3 (Pr = 6.6 μC/cm2, Pmax = 41.6 μC/cm2) and the literature cited therein [30]. The inset shows the P-E loops measured from the Au/SrTiO3/LSMO (x ≈ 1/3) capacitors [28].
Figure 4. Structural phase transition and FE properties of SrTiO3. (a) Temperature–strain phase diagram showing an AFD transition and a further FE transition [102]. (b,c) Piezoresponse force microscopy (PFM) amplitude and phase images of strain-free SrTiO3 films recorded at 50 K at biasing voltages of ±10 V [29]. (d) Comparisons of Pr and maximum polarization (Pmax) between Ti/O-deficient SrTiO3 (Pr = 6.6 μC/cm2, Pmax = 41.6 μC/cm2) and the literature cited therein [30]. The inset shows the P-E loops measured from the Au/SrTiO3/LSMO (x ≈ 1/3) capacitors [28].
Materials 14 07854 g004
Figure 5. Anisotropic energy storage pathways. (a) Predicted phase transition pathways for PbZrO3 associated with ionic and full structural relaxation after applying electric fields along different crystallographic directions. The multiscale structure prediction combines classical and first-principles density functional theory [114]. (b) Structural anisotropy-dependent energy storage performance of AFE PbHfO3 thin films grown on different crystal-plane-terminated SrTiO3 substrates [71].
Figure 5. Anisotropic energy storage pathways. (a) Predicted phase transition pathways for PbZrO3 associated with ionic and full structural relaxation after applying electric fields along different crystallographic directions. The multiscale structure prediction combines classical and first-principles density functional theory [114]. (b) Structural anisotropy-dependent energy storage performance of AFE PbHfO3 thin films grown on different crystal-plane-terminated SrTiO3 substrates [71].
Materials 14 07854 g005
Figure 6. Time-dependent phase transition in NaNbO3 under excitation of electric field. (a) Evolution of XRD superlattice reflections with increasing electric field for the AFE P (stage I) and FE Q (stage III) phases. (b,c) Pseudocubic lattice parameters and unit cell volume plotted as a function of electric field amplitude. (d) Evolution of the P-phase fraction, macroscopic polarization and longitudinal strain S33 as a function of electric field [126]. Instead of the decoupled polarization switching process, stage II may also be interpreted as the emergence of an intermediate FiE phase.
Figure 6. Time-dependent phase transition in NaNbO3 under excitation of electric field. (a) Evolution of XRD superlattice reflections with increasing electric field for the AFE P (stage I) and FE Q (stage III) phases. (b,c) Pseudocubic lattice parameters and unit cell volume plotted as a function of electric field amplitude. (d) Evolution of the P-phase fraction, macroscopic polarization and longitudinal strain S33 as a function of electric field [126]. Instead of the decoupled polarization switching process, stage II may also be interpreted as the emergence of an intermediate FiE phase.
Materials 14 07854 g006
Figure 7. Structural phase transitions of PbZrO3 induced by electron-beam irradiation in a TEM. (a,b) Changes in lattice parameter (√2ap, √2bp) and volume (V = √2ap × √2bp × cp) plotted as a function of irradiation time, respectively. (c–e) Atomic-resolution NCSI-TEM images of an AFD-FD phase boundary and associated structural phase transitions recorded along [001]O direction, respectively. The oxygen displacements in (c) are indexed by (+, 0, –) symbols and colorful solid lines. The yellow arrows denote antiparallel Pb displacements. The atom types are Pb—yellow, Zr—green, O—red circles, respectively. (f) 2D cycloidal order of polarization for the transient FE-FD phase obtained using quantitative TEM study [127].
Figure 7. Structural phase transitions of PbZrO3 induced by electron-beam irradiation in a TEM. (a,b) Changes in lattice parameter (√2ap, √2bp) and volume (V = √2ap × √2bp × cp) plotted as a function of irradiation time, respectively. (c–e) Atomic-resolution NCSI-TEM images of an AFD-FD phase boundary and associated structural phase transitions recorded along [001]O direction, respectively. The oxygen displacements in (c) are indexed by (+, 0, –) symbols and colorful solid lines. The yellow arrows denote antiparallel Pb displacements. The atom types are Pb—yellow, Zr—green, O—red circles, respectively. (f) 2D cycloidal order of polarization for the transient FE-FD phase obtained using quantitative TEM study [127].
Materials 14 07854 g007
Figure 8. Defect-core-induced unit-cell-wise energy storage pathway in [001]O-oriented PbZrO3. (a–e) Phase and domain structure evolution as a function of electron-beam irradiation time (dose rate = 3.8 × 106 e/nm2⋅s). The cyan shadow and thick yellow arrows denote the defect core and polarization, respectively. (f–j) Corresponding lattice strain exx (horizontal [100]O direction) maps obtained using geometric phase analysis (GPA) [118].
Figure 8. Defect-core-induced unit-cell-wise energy storage pathway in [001]O-oriented PbZrO3. (a–e) Phase and domain structure evolution as a function of electron-beam irradiation time (dose rate = 3.8 × 106 e/nm2⋅s). The cyan shadow and thick yellow arrows denote the defect core and polarization, respectively. (f–j) Corresponding lattice strain exx (horizontal [100]O direction) maps obtained using geometric phase analysis (GPA) [118].
Materials 14 07854 g008
Figure 9. Deoxygenation of the bottom (La1-xSrx)MnO3 (LSMO, x ≈ 1/3) electrode with increasing positive bias in an LSMO/HZO/LSMO capacitor. (a) Polarity-dependent oxygen voltammetry process (VO: oxygen vacancy) and morphology of the dielectric capacitor. (b–d) The iDPC-STEM images of a representative region of the bottom LSMO layer, showing (b) the virgin 0.0 V state, (c) the BM precursor phase (with mixed MnO5 and MnO6 polyhedra) at 2.0 V and (d) the BM phase at 4.0 V viewed along the [110]c zone axis. The atom types in the insets are La/Sr—green, Mn—red and O—brown [158].
Figure 9. Deoxygenation of the bottom (La1-xSrx)MnO3 (LSMO, x ≈ 1/3) electrode with increasing positive bias in an LSMO/HZO/LSMO capacitor. (a) Polarity-dependent oxygen voltammetry process (VO: oxygen vacancy) and morphology of the dielectric capacitor. (b–d) The iDPC-STEM images of a representative region of the bottom LSMO layer, showing (b) the virgin 0.0 V state, (c) the BM precursor phase (with mixed MnO5 and MnO6 polyhedra) at 2.0 V and (d) the BM phase at 4.0 V viewed along the [110]c zone axis. The atom types in the insets are La/Sr—green, Mn—red and O—brown [158].
Materials 14 07854 g009
Table 1. Comparison of energy storage performance between representative dielectric films and between representative dielectric bulks.
Table 1. Comparison of energy storage performance between representative dielectric films and between representative dielectric bulks.
Dielectric FilmsJrec (J/cm3)η (%)Pmax (μC/cm2)EB (MV/cm)Ref.
SrTiO3(1)30789~1256.8[28]
La-Ba-Zr-doped (Na0.5Bi0.5)TiO3 (1)15497113.53.5[37]
Sm-doped BiFeO3-BaTiO3(2)152> 90~605.2[38]
BaZr0.35Ti0.65O3 multilayer(2)13073.8528.75[41]
Mn-doped Na0.5Bi0.5TiO3-BaTiO3-BiFeO3(3)102601242.86[42]
La-doped Pb(Zr,Ti)O3(4)85651154.5[43]
Pb0.5Sr0.5HfO3(5)7797~535.12[39]
Dielectric bulksJrec (J/cm3)η (%)Pmax (μC/cm2)EB (MV/cm)Ref.
Na0.5Bi0.5TiO3-Sr0.7Bi0.2TiO3(3)21.5~8067103[36]
0.90NaNbO3-0.10BiFeO3(6)18.578.764~1.0[44]
La-doped Pb(Zr0.55Sn0.45)0.995O3(7)10.48741.30.4[32]
(Pb0.91Ba0.045La0.03)(Zr0.6Sn0.4)O3(7)8.1692.1400.34[31]
BiFeO3-BaTiO3-NaNbO3(6)8.1290~520.36[45]
AgNbO3-AgTaO3(6)7.586320.53[46]
Note: The bottom electrode types are (1) La0.67Sr0.33MnO3, (2) Nb-SrTiO3, (3) Pt, (4) LaNiO3, (5) SrRuO3, (6) silver paste, (7) Au.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wei, X.-K.; Dunin-Borkowski, R.E.; Mayer, J. Structural Phase Transition and In-Situ Energy Storage Pathway in Nonpolar Materials: A Review. Materials 2021, 14, 7854. https://doi.org/10.3390/ma14247854

AMA Style

Wei X-K, Dunin-Borkowski RE, Mayer J. Structural Phase Transition and In-Situ Energy Storage Pathway in Nonpolar Materials: A Review. Materials. 2021; 14(24):7854. https://doi.org/10.3390/ma14247854

Chicago/Turabian Style

Wei, Xian-Kui, Rafal E. Dunin-Borkowski, and Joachim Mayer. 2021. "Structural Phase Transition and In-Situ Energy Storage Pathway in Nonpolar Materials: A Review" Materials 14, no. 24: 7854. https://doi.org/10.3390/ma14247854

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop