Next Article in Journal
Light-Triggered Polymersome-Based Anticancer Therapeutics Delivery
Next Article in Special Issue
The Influence of Magnetic Field and Nanoparticle Concentration on the Thin Film Colloidal Deposition Process of Magnetic Nanoparticles: The Search for High-Efficiency Hematite Photoanodes
Previous Article in Journal
A First-Principle Study of Interactions between Magnesium and Metal-Atom-Doped Graphene
Previous Article in Special Issue
Tantalum Oxide as an Efficient Alternative Electron Transporting Layer for Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Strain Engineering: A Pathway for Tunable Functionalities of Perovskite Metal Oxide Films

1
Department of Materials Design and Innovation, University at Buffalo—The State University of New York, Buffalo, NY 14260, USA
2
Center for Integrated Nanotechnologies, Los Alamos National Laboratory, Los Alamos, NM 87545, USA
3
Division of Quantum Phases & Devices, Department of Physics, Konkuk University, Seoul 05029, Korea
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(5), 835; https://doi.org/10.3390/nano12050835
Submission received: 19 January 2022 / Revised: 14 February 2022 / Accepted: 24 February 2022 / Published: 1 March 2022
(This article belongs to the Special Issue Structure, Properties and Applications of Nanocrystalline Thin Films)

Abstract

:
Perovskite offers a framework that boasts various functionalities and physical properties of interest such as ferroelectricity, magnetic orderings, multiferroicity, superconductivity, semiconductor, and optoelectronic properties owing to their rich compositional diversity. These properties are also uniquely tied to their crystal distortion which is directly affected by lattice strain. Therefore, many important properties of perovskite can be further tuned through strain engineering which can be accomplished by chemical doping or simply element substitution, interface engineering in epitaxial thin films, and special architectures such as nanocomposites. In this review, we focus on and highlight the structure–property relationships of perovskite metal oxide films and elucidate the principles to manipulate the functionalities through different modalities of strain engineering approaches.

1. Introduction

Perovskite has emerged as an important class of material in technologically important sectors. The research interests and application development of such materials have grown exponentially. Since many elements from the periodic table can be fitted into such a unique crystal structure [1], it is not an overstatement for us to envision that perovskite can be used as a framework to study a wide range of properties and functionalities including ferroic orderings, superconductivity, colossal magnetoresistance, optoelectronics, etc. [2]. Oxide perovskite thin films have been of particular interest in terms of both fundamental research and technological applications for oxide electronics [3]. Since the physical properties of such materials can be further affected by the chemical doping [4,5,6,7] and epitaxial strain [8,9,10,11,12,13], one can strain the materials to tune the properties of the materials towards specific applications. Recently, more approaches for strain engineering have come to the fore. Examples include mechanically straining free-standing single-crystalline membranes [8,14] grown on the buffered substrate where the buffer layer can be etched off thereafter, and the ion implantation of species such as helium [15] and nitrogen [16] to strain the lattice. Perovskite materials have also garnered interest in a wide range of applications such as solar cells, photo/electrocatalysis, photopolymerization, thermoelectrics, resistive switching devices, etc. [17,18,19,20,21,22,23,24,25]. In this review, we will focus on the role of chemical pressure and epitaxial strain in perovskite metal oxide films, superlattices, and vertically aligned nanocomposites.
The ideal structure of perovskite with a chemical formula ABO3 is cubic, as shown in Figure 1a, and composed of a three-dimensional network of corner-sharing [BO6] octahedra, where the A cation sits in the cubo-octahedral site between adjacent octahedra. However, the structures of most perovskite materials exhibit a lower symmetry, such as rhombohedral, orthorhombic, or tetragonal [2]. The geometry and symmetry of the crystal structure are directly affected by the relative size of the comprising ions and often viewed through the proxy of Goldschmidt tolerance factor, t :
t = R A + R X 2 ( R B + R X )
For the ideal cubic aristotype, t = 1 . As t lowers further, the structures distort via octahedral tilts and rotations, as shown in Figure 1b, to lower the symmetry until the perovskite phase stops being stable [2,26]. From the viewpoint of chemical compositions, the structure can therefore be distorted by doping isovalent elements of different sizes to exert a chemical pressure on the lattice.
Another way to affect the crystal structure is through epitaxial strain. While bulk crystals are brittle and fragile to mechanical strain, epitaxial thin films can tolerate strains of the order of several percent [27,28]. Such epitaxial thin films have a rich history in perovskite metal oxides facilitated by the advances in thin-film growth techniques such as molecular beam epitaxy, pulse laser deposition, sputtering, metal–organic chemical vapor deposition, and chemical solution deposition. Biaxial strains in perovskite metal oxide films can be used as a lever to tune physical properties such as electric polarization, magnetoresistance and magnetic anisotropy, and stabilize metastable phases [9]. Strain can distort the perovskite structure by tilting, rotating, and elongating the oxygen octahedra, and therefore significantly affect the properties of given materials such as Curie temperature (TC) [10,29,30,31], magnetic anisotropy [32,33,34,35,36,37], and magnetotransport properties of ferromagnetic materials [38,39,40,41]. In ferroelectric perovskite oxides, for example, the epitaxial strain can enhance the TC for several hundreds of degrees [11,42,43,44,45]. One of the best illustrations is probably SrTiO3, where the lattice strain can make SrTiO3 ferroelectric at room temperature, despite the unstrained SrTiO3 being paraelectric throughout all temperatures [13]. Similarly, epitaxial strain-engineering can also be leveraged in multilayer epitaxial heterostructures, i.e., epitaxial layered structures, superlattices, and vertically aligned nanocomposites (VANs). For example, ferroelectric SrTiO3 has been experimentally demonstrated in fully strained SrTiO3 films [13], superlattices SrTiO3/BaZrO3 [45] and SrTiO3/BaTiO3 [11], and vertically aligned epitaxial SrTiO3:MgO and SrTiO3:Sm2O3 nanocomposite films [46,47].

2. Strain Engineering through Chemical Substitution or Chemical Pressure

The substitution of ions via chemical pressure is one of the common approaches to strain the crystal structure. As the A-site cations prefer to form close-shelled electronic configurations with fixed valency, they largely play a structural role. The cation size influences lattice constants, bond lengths and angles, and octahedral rotations and tilts. In other words, a distortion of the central [BO6] octahedra can directly affect the material properties. Both the mean size and the size mismatch of the cations on the A-site affect the electronic and structural phase transitions [48,49,50].
To illustrate strain engineering to tune the functionalities of perovskites through chemical substitution or chemical pressure, we use AMnO3 manganates as the model system. Mixed-valent perovskite manganates, AMnO3, are one of the most widely studied systems for their colossal magnetoresistance (CMR) property. The Curie temperature of these materials, TC, is very sensitive to the change in chemical pressure which can be introduced via the substitution of trivalent rare earth metals of different sizes. For example, as the atomic radius of the A cation changes from a relatively larger to a smaller size in the case of La, Pr, Nd and Eu, and the Neel temperature, TN, for the antiferromagnetic insulator phase is lowered with the change in the B-O-B angle away from 180°, as shown in Figure 2a [51]. In Ln1−xMxMnO3 systems, where Ln = trivalent lanthanide (La, Pr, Nd, …), M = divalent cation (Ca, Sr, Ba, …), the disorder due to size disparity between the two A-site cations plays a role in determining and tuning their electronic properties. This is evident from the fact that different Ln1−xMxMnO3 perovskites with the same doping level and tolerance factor can have quite different metal-to-insulator transition temperatures [48]. As the x value increases, the hole doping increases linearly. The average A cation size also changes linearly with x value, but the change in disorder is non-linear, with it being minimum at x = 0 and 1, and maximum at x = 0.5. The statistical variance between the distribution of the radii can be considered as the quantity to parametrize this disorder [48]. La1−xSrxMnO3 is the most well studied material in this class. Here, the bond lengths and angles are lowered with an increase in Sr content. The effect of Sr content on lattice parameters and bond angle is depicted in Figure 2b [52]. The change here is even more dramatic compared to the prior example of switching trivalent rare-earth metals. As shown in Figure 2c [52], the TN first begins to decrease until a transition to a ferromagnetic insulator phase, and eventually to a ferromagnetic metal with TC of the ferromagnetic phase well above room temperature. The pressure exerted in the lattice modifies local parameters such as the Mn-O bond length and the O-Mn-O bond angle which, in turn, affects the balance between the co-existing metallic and insulating states, and thus their CMR properties [53]. Mixed-valent perovskite manganates also exhibit many charge-ordered phases that are affected by factors such as the size of the A cations as well as isotopic and chemical substitutions [54,55,56].
Another well studied system is the titanate, ATiO3 (A = Ba, Sr, Ca), where the ferroelectric Curie temperature, TC, was seen to be manipulatable by the size variance of the A-site cation mixture [57]. Ba1−xSrxTiO3 shows a decrease in TC with an increase in Sr2+ as the smaller cation stabilizes the more symmetric cubic phase with tolerance factor decreasing from that of the ferroelectric BaTiO3, which has a tolerance factor of 1.06. However, in Ba1−xCaxTiO3, the size mismatch between the A cations is bigger and only x up to 0.24 is stable without phase segregation. The TC rises to x = 0.08 from 403 K to 410 K, and then decreases to the solution limit. This behavior cannot be explained by simple size arguments, where the initial increase has been rationalized by the strain effect due to the mismatch between the two cations and the subsequent decrease due to the size effect [49,57].
One of the major limitations of chemical substitution to exert chemical pressure over epitaxial strain engineering is the introduction of disorder and heterogeneity associated with chemical substitution in such films. Disorder, for example, can broaden the phase transition by hundreds of degrees [58,59,60]. We will discuss the epitaxial control of strain in the following sections.

3. Epitaxial Strain Engineering

A range of commercially available single-crystal substrates with different lattice parameters makes it possible to epitaxially grow perovskite oxide thin films with different strain states. The lattice mismatch and the strain between the substrate and the film can then be designed by selecting an appropriate substrate. Other considerations besides the lattice parameters are chemical and thermal expansion compatibility. The most commonly used and commercially available substrates and thin-film materials of interest are shown in the lower part of Figure 3d. Lots of materials of interest (top part of Figure 3d) can be epitaxially grown on these substrates with certain lattice mismatches.
For a simple cubic lattice, the lattice mismatch, f, between the substrate and the film can be defined as:
f = a s a 0 a 0
where a s and a 0 are the lattice parameters of the substrate and film in an unstrained state, respectively. In a fully strained epitaxial film, the in-plane lattice parameters of the film are constrained to the lattice of the substrate. However, as the thickness of the film increases above a critical film thickness, defects such as misfit dislocations become energetically favorable, and the lattice strain relaxes.
For isotropic films, the biaxial strain is:
ε x x = ε y y = a a 0 a 0
and the out-of-plane strain is:
ε z z = a a 0 a 0
where a and a are the in-plane and out-of-plane lattice parameters of the strained epitaxial film. A substantial portion of the change in lattice parameters due to epitaxial strain is through changes in the relative magnitude of the [BO6] octahedral rotations and the B–O bond lengths [51,61].
The strain on the film evolves with the increase in thickness of the film, as the top of the film starts to relax with dislocation defects and the strained in-plane lattice parameter starts to converge to the relaxed bulk value, deviating from that of the film–substrate interface, eventually reaching full relaxation [62]. The strain in the film leads to the storage of elastic energy and at the critical thickness, this energy equals the formation energy for misfit dislocation defects in the People–Bean model [62]. Critical thickness, therefore, is inversely correlated with the lattice mismatch [63]. However, the critical thickness can also differ depending on experimental growth conditions such as growth temperature, post growth annealing, stoichiometry, etc., due to the influence of thermal strains and defect formation. Thermal strains arise due to a mismatch between the in-plane thermal expansion coefficients between the two lattices during the cool-down process [64,65,66]. In cases of post-growth annealing, which are often used to decrease the number of oxygen vacancies and improve the crystallinity of the film, the detailed annealing protocol could also change the strain state in the as-deposited films. For instance, in the deposition of epitaxial films of La0.7Sr0.3MnO3 on Al2O3/MgO substrate–buffer platform, it has been reported that the post-growth annealing temperature beyond the deposition temperature of ~900 °C could create an irreversible strain relaxation and degrade the magnetic saturation of the films. The properties of the films were only partially recovered by a second annealing step at 700 °C [67].
Vailionis et al. studied the strain accommodation in the rhombohedral La0.67Sr0.33MnO3 (LSMO) via lattice modulations and rotations [38,68]. Under compressive strain, LSMO has an (110) out-of-plane-oriented monoclinic unit cell with space group P21/m (No. 11), while under tensile strain it exhibits an (001) out-of-plane-oriented tetragonal unit cell with space group Cmcm (No. 63). Under compressive strain, out-of-phase octahedral rotation around the (001) direction occurs, while under tensile strain these rotations are absent. The octahedra are rotated in phase around the (100)-axis of the pseudocubic unit cell and out of phase around the (010) direction in both cases (Figure 4a–d). The additional strain along the (100)-direction is accommodated by periodic lattice modulations, as shown in Figure 4f. The changes in octahedral rotations owing to stress and the dissimilar in-plane rotational patterns affect the in-plane magnetic anisotropy in LSMO films. Similar results were seen for the orthorhombic SrRuO3, and the authors argued that this could be extrapolated to other rhombohedral and orthorhombic perovskite oxides [38]. Strain accommodation via similar lattice modulations has also been experimentally observed by high-resolution transmission electron microscopy in epitaxial YBa2Cu3O7−δ with the presence of twin boundaries and intergrowths [69,70]. High-resolution transmission electron microscopy has emerged as a powerful tool towards enabling the direct observation of such structural changes [71]. In ultrathin La2/3Sr1/3MnO3 films grown on NdGaO3 substrates, strong oxygen octahedral coupling is found to transfer the octahedral rotation in the perovskite substrate to the perovskite thin film near the interface [72]. An unexpected realignment of the magnetic easy axis along the short axis of the unit cell as well as the presence of a giant anisotropic transport in these ultrathin La2/3Sr1/3MnO3/ NdGaO3 films was observed. Similar control over octahedral tilts was also demonstrated in SrRuO3 by using a 0–4 unit cell thick Ca0.5Sr0.5TiO3 buffer layer on GdScO3 substrates [73]. The Ru–O–Ti and the Ru–O–Ru bond angles at the interface could be tuned via changing the thickness of the buffer layer, along with affecting magnetic anisotropy in the entire SrRuO3 layer.
In (001) LSMO, the resistivity of thin films has been shown to be controlled via epitaxial strain. The effect on the temperature dependence of resistivity can be seen in Figure 5a [30]. Metal-to-insulator transition can be lowered considerably by using substrates that can lead to a higher tensile strain. This can be clearly seen from a systematic comparison of the resistivity vs. temperature characteristics of films on different substrates such as SrTiO3 (0.5%), DyScO3 (1.6%) GdScO3 (2.3%), SmScO3 (2.7%) and NdScO3 (3.2%). The Curie temperature and magnetization were also seen to be dependent on substrate choice (Figure 5b). In accordance with theoretical predictions by Mills et al. [74], a best-fit plane of the Curie temperature’s dependence on the bulk compressive strain ε B   (which tends to increase electron hopping probability and reduce the effects of electron–phonon coupling) and biaxial distortion ε* (which increases the Jahn–Teller splitting in the eg orbitals and acts only to reduce TC) is shown in Figure 5c.
Interestingly, in lightly doped manganites, the in-plane compressive strain has been found to increase the TC significantly and produce metal-insulator-transition (MIT) in the compounds that are supposed to be ferromagnetic insulators (FMI) [60]. For example, Bulk La0.9Sr0.1MnO3 is a ferromagnetic insulator with a Curie temperature (TC) of 145 K. When grown epitaxially on LAO, thinner La0.9Sr0.1MnO3 (~15 nm) films are metallic with a greatly enhanced TC, which is 97 K higher than the bulk value. In-plane compressive strain (−1.5%) was reported to be partially responsible for the TC enhancement. Strain-induced stoichiometry modification also plays a role in modulating the TC and strain relaxation in these thin films, accommodated both by misfit dislocations and La deficiency.
The opposite effect of compressive in-plane strain on optimally doped and lightly doped manganites is not surprising. In optimally doped manganites, the in-plane compressive strain induces the change in the O–Mn–O bond angle and length away from the ideal case. This, therefore, reduces TC and MIT. It is noted that this strain in lightly doped manganites induces oxygen octahedral tilt and straightens the out-of-plane O–Mn–O chain which promotes MIT.
Lattice strain has also been implicated in oxygen vacancy formation which can lead to radical changes in the physical properties of these films. For instance, oxygen vacancies may act as pinning sites for ferroelectric and magnetic domain wall movements, create structural disorder and introduce electronic defects [75,76,77]. In the prototypical ferroelectric perovskite system, PbTiO3, ab initio study showed that oxygen vacancies that act as strong domain pinning centers can be modulated to a non-pinning center via compressive misfit strain by changing the formation energies of the various types of oxygen vacancies, as shown in Figure 6a,b [78]. Ferroelectric distortion stabilized in SrTiO3 by epitaxial strain has also been shown to promote the formation of oxygen vacancies [79]. In half-doped manganite, La0.5Sr0.5MnO3, the epitaxial strain was shown to modulate ferromagnetic and antiferromagnetic phase proportions by manipulating the oxygen nonstochiometry as evidenced by the change in Mn3+ and Mn4+ ratio with the increase in thickness (Figure 6d). This was seen to lead to depressed magnetization and enhanced exchange bias (Figure 6c) [80]. In La0.67Ca0.33MnO3, the epitaxial strain was implicated in the formation of a dead layer near the interface with a higher concentration of oxygen vacancies with reduced Mn valence and the unidirectional displacement of Mn ions [81].
The tuning of TC in ferroelectric systems via epitaxial strain can be well illustrated in the titanates. The much enhanced ferroelectric property is most explicitly demonstrated by the case of SrTiO3, which is normally not ferroelectric at any temperature without any lattice strain. However, when SrTiO3 is grown epitaxially on DyScO3, with an in-plane tensile strain of ~0.94%, fully strained SrTiO3 film shows a TC as high as 293 K [13]. This strained SrTiO3 film was later demonstrated to be an orthorhombic phase [82], and showed an antiferrodistortive phase transition that exhibited ferroelastic–ferroelectric multiferroicity [83]. The strain vs. Curie temperature phase diagram of SrTiO3 predicted from thermodynamic principles using a Landau–Ginsburg–Devonshire-type theory by Persev et al. is shown in Figure 7a. It illustrates the complex interconnection between the Curie temperature and the lattice strain [13,84].
The coupling between the lattice strain and the ferroelectric properties such as Curie temperature is common in perovskite metal oxides. For instance, the Curie temperature and the ferroelectric polarization of the classical ferroelectric perovskites BaTiO3 and PbTiO3 can also be influenced by lattice strain. BaTiO3 thin films coherently grown on GdScO3 and DyScO3 substrates with a misfit strain of about −1.0% and −1.7%, respectively, show a large increase in ferroelectric transition temperature with a TC 400 °C on GdScO3 and 540 °C on DyScO3. In comparison, the TC is only 120 °C for bulk BaTiO3. It was also observed that the strained thin film had a remnant polarization 250% higher than bulk BaTiO3 single crystals (Figure 7b) [11]. These results are comparable to unstrained Pb(ZrxTi1−x)O3, but with a lead-free composition, which is preferable due to environmental and human health implications of the processing and disposal of the toxic element. Theoretical studies also suggest that there are temperature and strain regions, in particular, under tensile strain, where the system decomposes into multi-domain structures [85,86,87]. Despite both BaTiO3 and PbTiO3 being titanates, the primary drivers of ferroelectric polarization differ. For BaTiO3, the polarization arises from the off-centering of the Ti+4 cation, while the 6s2 lone pair on Pb2+ in PbTiO3 is stereoactive and contributes to the ferroelectric polarization too. Different approximations and assumptions about domain configurations in thermodynamic analysis and phase-field simulations of strain-phase diagrams of PbTiO3 lead to markedly different phase diagrams, as shown in Figure 8. All the diagrams involve the paraelectric phase, the tetragonal c-phase, and are under the assumption of a single-domain film. The orthorhombic aa-phase exists at higher in-plane strains with a ‘‘monoclinic gap’’ r-phase between the tetragonal and orthorhombic phases (Figure 8a). However, real films always have more than one type of domain and extended phase diagrams that account for two-dimensional domains are subsequently derived from thermodynamic calculations of Pertsev et al. (Figure 8b,c) [88,89]. Subsequently, phase-field models, which could predict multi-domain states without making assumptions about domain wall orientation, were used to derive the three-dimensional phase diagram shown in Figure 8d. Experimental and theoretical calculations concur that the TC can be increased via both tensile and compressive strains. It is noted that the TC is also evidently affected by the film thickness. The increase in TC with film thickness indicates that the ferroelectric property is further affected by grain size in addition to the lattice strain [90,91].

4. Superlattices

In layered structures and superlattices, strain relaxation behaviors are different from single-layer systems. The lattice mismatch between the layers and their thicknesses can all contribute to the strain state of the whole structure. For instance, strain engineering of superlattices such as BaTiO3/SrRuO3 can be achieved by choosing suitable substrates and controlling the thickness of each component layer [63]. One of the remarkable examples of superlattices and their unique interplay with strain is seen in the system of (SrTiO3)/(BaTiO3)/(CaTiO3), where the BaTiO3 layer shows strain-induced ferroelectric polarization with 50% enhancement of the superlattice global polarization compared to a pure BaTiO3 layer grown similarly (Figure 9) [94]. The BaTiO3 layer remains fully strained as long as the number of unit cells does not exceed that of the combined SrTiO3 and CaTiO3. Figure 9c shows how the lattice parameter of the superlattice shifts as a function of the stacking sequence. Their polarization is depicted in Figure 9d. Polarization is strongest when the BaTiO3 layer is thin enough to be fully strained but thick enough to contain enough non-interfacial TiO6 octahedra. The advantage of such superlattices is the possibility of breaking inversion symmetry that persists with most two-component or symmetric superlattices [95], and provides additional freedom in tuning the average lattice parameter [94]. The interrogation of ε(E) curves shown in Figure 9b illustrates asymmetric response due to compositionally broken symmetry.

5. Vertically Aligned Nanocomposites

As discussed above, epitaxial strain provides an alternative pathway to chemical doping for manipulating the physical properties of a range of perovskite metal oxide films. Experimental results have indeed demonstrated that much improved physical properties and/or emergent behaviors could be accomplished by growing lattice-strained epitaxial perovskite metal oxide thin films [11,13,94,96,97,98]. However, enhanced functionalities are typically achieved in such epitaxial metal oxide films with a thickness of less than a few tens of nanometers at which the lattice strain could be maintained [11,13,97,98,99].
For many technological applications, lattice-strained thick epitaxial perovskite metal oxide films (in the range of ~µm or above) are needed. In a conventional lattice strain framework, heteroepitaxial strain resulting from the lattice mismatch between the film and the substrate exists only below a critical film thickness. In other words, films are strained to the substrate lattice up to a critical thickness, hc. Above hc, the strain energy becomes so large that the nucleation of misfit dislocations at the interface is then energetically favorable to relieve the strain. The critical thickness, hc, at which the formation of dislocations becomes favorable, can be expressed as:
h c = b 4 π f ( 1 + v ) [ ln ( h c b ) + 1 ]
where v is the Poisson ratio, b is the magnitude of the Burgers vector, and f is the lattice mismatch between the film and the substrate. As can be seen from Equation (5), reducing the lattice mismatch could circumvent the upper limit of critical film thickness to some extent.
Great success in the lattice strain of much thicker films has been demonstrated in the vertically aligned nanocomposite, where the lattice strain is predominantly controlled by the vertical interface between the two individual constituents instead of the lateral interface between the film and the substrate. As shown in Figure 10a, vertically aligned nanocomposite film is composed of a regularly arranged A:B network from an array of vertically aligned nanopillar-like material A with a feature size d entrenched in a matrix of material B. Since both constituents A and B are laterally coupled across the vertical interface through the whole film thickness, as shown in Figure 10b, the vertical strain can exist in thick films. For example, lead-free, self-assembled BaTiO3:Sm2O3 nanocomposite films (up to 1.25 µm thick) were grown that exhibited tetragonality up to at least 800 °C and strong remanent polarization to at least 330 °C [100]. Similar nanocomposites, such as Ba0,6Sr0.4TiO3:Sm2O3, with film thickness in the range of 1.0 µm were also demonstrated that showed tetragonality up to at least 400 °C [101].
Characteristics of lateral epitaxial strain are well studied. However, the relationship between vertical strain and microstructure in vertically aligned nanocomposites is less known. It is important to know the parameters that control the vertical strain and microstructure of the nanocomposites. The final microstructure (pillar size, shape, and lateral spacing) of a vertically aligned nanocomposite film is determined by the minimization of the total free energy, which includes the elastic and interfacial energies of the system composed of the two individual phases (A and B) and the substrate. At the given growth conditions, both the size and volume of pillars predominantly determine the interfacial area at the vertical interface. It is accepted that the pillar size d, the density of the nanopillars, as well as the defects along the vertical interface determine the overall strain state of vertically aligned nanocomposites [32]. Figure 11 shows the general relationship between the vertical strain, the pillar feature size, and the volume of pillars [102]. Figure 11 also shows the general correlation among the growth temperature, pillar size, and strain state. In a system with a fixed second-phase volume, a higher growth temperature will generally produce a larger pillar size. This will result in a smaller interface area, which usually leads to a lower vertical strain. On the other hand, a lower growth temperature generally results in a smaller pillar size and larger interface area, which can lead to a larger vertical strain. In a system with a fixed pillar size, the increase in the pillar density can increase the total vertical interface area and thus lead to a larger strain. It was reported that the vertical strain in vertically aligned nanocomposites is ultimately related to the vertical interfacial area and interfacial dislocation density [12,32]. This can be understood by the following. In a nanocomposite with uniform nanopillars embedded in the film matrix, the total volume of nanopillar phase is V = n × πd2/4 × h, where n is the number of nanopillars, and h is the film thickness or pillar height. The total vertical surface area can be expressed as S = n × πd × h = πd × 4V/(πd2) = 4V/d. Since Figure 11 shows strain ε is proportional to 1/d and V then, ultimately, strain is directly controlled by the vertical interface area. The efficiency of the vertical interface coupling is related to the coherence of the interface.
Table 1 and Table 2 have summarized some vertically aligned nanocomposite films with the inclusion of perovskite metal oxides reported in the literature [32]. For most material systems, the assembly occurs as a result of nucleation and growth, with embedded pillars of the elastically stiffer phase in the matrix of the softer phase or the perovskite oxides. Pseudospinodal decomposition is the other growth mode which can lead to a checkerboard configuration which is associated with symmetry-lifting crystal lattice rearrangement [103,104].
Such vertical interface strain has been used to tune functional properties in oxides. Room-temperature ferroelectricity in vertically strained SrTiO3 nanocomposite with MgO nanopillars embedded in the strained SrTiO3 matrix has been demonstrated [46]. An out-of-plane strain of ~1.5% was observed in the film matrix. Elastic coupling between MgO and SrTiO3 can be confirmed by etching away the MgO pillars, after which the lattice parameter of the SrTiO3 relaxes to 3.911 Å, as shown in Figure 12 [12]. The measured strain was corroborated by phase-field modeling to explore the spatial distribution of the strain and the resultant polarization. The TC was extracted from optical second-harmonic generation (SHG) measurements. The phase transition was found to be very broad, which was consistent with the non-uniform strain distribution and the non-uniformity shown in the piezoelectric force microscopy measurement.
A multiferroic epitaxial nanocomposite of CoFe2O4 nanopillars in a matrix of BaTiO3 was demonstrated [118]. The CoFe2O4 nanopillars experienced an out-of-plane compressive strain of 0.8%. The coupling between the ferroelectric perovskite BaTiO3 and the ferrimagnetic spinel CoFe2O4 allowed the interconversion of energies stored in electric and magnetic fields to create an emergent ferroelectromagnet. The magnetoelectric coupling could be explained by the strong elastic interactions between the two phases by thermodynamic analyses, which was also further supported by the fact that the coupling could not be reproduced in multilayer structures due to the substrate clamping effect.
Chen et al. explored the vertical strain engineering of the La0.7Sr0.3MnO3 system via incorporating nanopillars of ZnO and MgO in the nanocomposites [32]. The strain in the La0.7Sr0.3MnO3:MgO nanocomposites was much larger than that in the La0.7Sr0.3MnO3:ZnO with the same volume of the second phase due to the larger lattice mismatch and elastic modulus mismatch. Figure 13 shows the comparative studies of the nanocomposites with different second-phase materials and volume ratios. The rise in resistivity for the La0.7Sr0.3MnO3:ZnO at ~50% of second-phase (ZnO) volume can be explained by a second-phase-induced volume effect. On the other hand, the sharp drop of over 10 times in resistivity of La0.7Sr0.3MnO3:MgO happens at only 20% second-phase (MgO) volume which is quite below the volume-induced percolation threshold. The experimental results, therefore, could predominantly be explained by the vertical strain effects. Vertical strain effects could also be the cause of the large drop in the metal-to-insulator transition temperature from 350 K of the bulk value of La0.7Sr0.3MnO3 to 150 K in the La0.7Sr0.3MnO3:MgO nanocomposite.
It is also noted that the strain on the LSMO matrix could be tuned via the density and size of the MgO nanopillars [32]. This, consequently, will affect the magnetic properties of the LSMO. For example, the LSMO:MgO nanocomposite with 5% MgO inclusion showed an in-plane easy axis. With increasing the density of MgO nanopillars (15% MgO inclusion), the easy axis switches to the out-of-plane direction. This switching of the magnetic anisotropy is correlated by the strain on the LSMO matrix exerted by the MgO nanopillars.

6. Future Prospects

As discussed above, perovskite metal oxides have been considered as an extremely important class of materials for technological applications due to their versatile and tunable physical properties. At the same time, the unique interplay between the charge, spin, orbital and lattice degrees of freedom across similar energy scales also makes the physical properties of such materials very sensitive to both microstructural and compositional variations. This is understandable considering that the distortions of crystal structure resulting from the change in chemistry and/or the eight corner-shared [BO6] octahedra will lead to the change in the electronic structure of the materials. Over the years, tremendous progress in controlled synthesis, advanced characterization at different length and time scales, and sophisticated modeling/simulation have helped us to fundamentally understand the processing–structure–property–performance relationship of perovskite metal oxides. Even so, more effort is much needed to further explore and design perovskites with desired properties, given their rich compositional diversity where the properties are closely tied to the crystal distortion such as the lattice strain. Below are some interesting and yet challenging research activities for the development of perovskite materials with enhanced, desired, and/or emergent properties.
(1)
An in-depth and quantitative understanding of the role of interface and defect on the properties is critical for the rational design of novel perovskite materials. This is especially important where superlattice and/or vertically aligned nanocomposite are the building blocks [11];
(2)
Substitution of ions via chemical pressure is one of the common approaches to strain the crystal structure. Much improved properteis could be achieved by forming multicomponent or high entropy perovskite oxides [119];
(3)
Epitaxial strain resulting from the lattice mismatch between the substrate and the film has been widely investigated for a range of perovskite metal oxide films. The exploration of dynamic strain by laminating freestanding perovskite metal oxide films or membranes onto a stretchable substrate could provide possibilities for strain-tunable devices [8];
(4)
Quantitative analysis of oxygen vacancies on the structural and physical properties of perovskite metal oxides is much needed. This is especially important for perovskites that are composed of cations with multiple valance states at the B site [120];
(5)
The formation of nanocomposite represents an attractive strategy to achieve tunable properties. More effort is needed to address the challenges related to the controlled synthesis to achieve long-range ordered structures [97];
(6)
While the most commonly studied perovskites are complex metal oxides, other perovskite materials, namely the halide perovskite [121], chalcogenide perovskite [122], and nitride perovskite [123], have attracted much attention lately. Halide perovskites, in particular, have attracted quite a lot of attention due to their potential applications for optoelectronic devices such as solar cells. However, the strain analysis and engineering in these thin films and devices have significantly different considerations than that of metal oxide perovskites, due to the soft ionic nature of the crystal lattice, generally non-epitaxial, and the complex interfaces of different charge transport layers [124,125,126,127,128]. While many studies has been undertaken and tremendous progress has been made in the synthesis and application of such materials, there are still many open questions related to the roles of strain, defect, interface, microstructure, and heterogeneity at different length scales on the physical properties of the materials. The recent review article published by Liu et al. summarizes the progress and challenges on this front admirably [129].

7. Conclusions

Perovskite metal oxides exhibit a multitude of important physical properties, such as metal-to-insulator transitions, superconductivity, ferroelectric, ferromagnetic, and magnetoresistive properties. As these properties are intimately linked with the structure of the perovskite, factors such as octahedral rotations, tilts and distortions, and chemical substitution with different sized cations can be also used to exert a chemical pressure to tailor these properties. Epitaxial strain is another lever that can be used to tune these properties in thin films by choosing an appropriate substrate with an appropriate lattice mismatch. Multilayers and/or superlattices take the advantage of layer thickness and multiple lateral interfaces to control the strain state of the assembly. Vertically aligned nanocomposites are an exciting way to construct strained thick films that is not possible in lateral heterostructures. Importantly, novel or improved functionalities, often not accessible from the individual constituents, could be produced through synergistic coupling interactions of known materials and lattice strain along the vertical directions. In the vertically aligned nanocomposite, it is possible to select one constituent as the active phase (typically perovskite metal oxide) with targeted functionality, and another as the passive phase (usually binary metal oxide), where the passive constituent is used to improve and/or tune the functionalities of the active phase. Such a strategy makes it possible to choose a much broader range of materials with targeted electrical, magnetic, optical, thermal, and/or mechanical properties for specific applications.

Author Contributions

Writing—original draft preparation, S.D. and Q.J.; writing—review and editing, S.D., A.C., W.N., B.P. and Q.J.; supervision, Q.J.; All authors have read and agreed to the published version of the manuscript.

Funding

The work at the University at Buffalo was partially supported by the U.S. National Science Foundation under award number ECCS-1902623. The work at Los Alamos National Laboratory was supported by the NNSA’s Laboratory Directed Research and Development Program. Los Alamos National Laboratory, an affirmative action equal opportunity employer, is managed by Triad National Security, LLC for the U.S. Department of Energy’s NNSA, under contract 89233218CNA000001.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

The work at Los Alamos National Laboratory was performed, in part, at the Center for Integrated Nanotechnologies (CINT), an Office of Science User Facility operated for the U.S. Department of Energy Office of Science. S.D. and Q.X.J. acknowledge the CINT Users Program.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Peña, M.A.; Fierro, J.L.G. Chemical structures and performance of perovskite oxides. Chem. Rev. 2001, 101, 1981–2018. [Google Scholar] [CrossRef] [PubMed]
  2. Tilley, R.J.D. Perovskites: Structure-Property Relationships; John Wiley & Sons: Hoboken, NJ, USA, 2016; ISBN 978-1-118-93566-8. [Google Scholar]
  3. Coll, M.; Fontcuberta, J.; Althammer, M.; Bibes, M.; Boschker, H.; Calleja, A.; Cheng, G.; Cuoco, M.; Dittmann, R.; Dkhil, B.; et al. Towards oxide electronics: A roadmap. Appl. Surf. Sci. 2019, 482, 1–93. [Google Scholar] [CrossRef]
  4. Liu, R.S.; Shen, C.H.; Hu, S.-F.; Chen, J.M. Chemical pressure control of curie temperature in La0.6(Ba0.4-XCax) MnO3. Mater. Chem. Phys. 2002, 75, 26–31. [Google Scholar] [CrossRef]
  5. Mishra, D.K.; Sahu, D.R.; Mishra, P.K.; Behera, D.; Pradhan, A.K.; Roul, B.K. Chemical pressure effect on CMR behavior of Sr substituted La0. 67Ca0. 33-XSrxMnO3. Phys. B Condens. Matter 2007, 400, 155–162. [Google Scholar] [CrossRef]
  6. Mumtaz, F.; Nasir, S.; Jaffari, G.H.; Shah, S.I. Chemical pressure exerted by rare earth substitution in BiFeO3: Effect on crystal symmetry, band structure and magnetism. J. Alloys Compd. 2021, 876, 160178. [Google Scholar] [CrossRef]
  7. Tiittanen, T.; Karppinen, M. Structure evolution upon chemical and physical pressure in (Sr1−XBax)2FeSbO6. J. Solid State Chem. 2017, 246, 245–251. [Google Scholar] [CrossRef] [Green Version]
  8. 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]
  9. Schlom, D.G.; Chen, L.-Q.; Fennie, C.J.; Gopalan, V.; Muller, D.A.; Pan, X.; Ramesh, R.; Uecker, R. Elastic strain engineering of ferroic oxides. MRS Bull. 2014, 39, 118–130. [Google Scholar] [CrossRef] [Green Version]
  10. Tsui, F.; Smoak, M.C.; Nath, T.K.; Eom, C.B. Strain-dependent magnetic phase diagram of epitaxial La0.67Sr0.33MnO3 thin films. Appl. Phys. Lett. 2000, 76, 2421–2423. [Google Scholar] [CrossRef]
  11. Choi, K.J.; Biegalski, M.; Li, Y.L.; Sharan, A.; Schubert, J.; Uecker, R.; Reiche, P.; Chen, Y.B.; Pan, X.Q.; Gopalan, V.; et al. Enhancement of ferroelectricity in strained BaTiO3 thin films. Science 2004, 306, 1005–1009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Chen, A.; Su, Q.; Han, H.; Enriquez, E.; Jia, Q.X. Metal oxide nanocomposites: A perspective from strain, defect, and interface. Adv. Mater. 2019, 31, 1803241. [Google Scholar] [CrossRef] [PubMed]
  13. Haeni, J.H.; Irvin, P.; Chang, W.; Uecker, R.; Reiche, P.; Li, Y.L.; 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]
  14. Pesquera, D.; Parsonnet, E.; Qualls, A.; Xu, R.; Gubser, A.J.; Kim, J.; Jiang, Y.; Velarde, G.; Huang, Y.; Hwang, H.Y.; et al. Beyond substrates: Strain engineering of ferroelectric membranes. Adv. Mater. 2020, 32, 2003780. [Google Scholar] [CrossRef] [PubMed]
  15. Guo, H.; Dong, S.; Rack, P.D.; Budai, J.D.; Beekman, C.; Gai, Z.; Siemons, W.; Gonzalez, C.M.; Timilsina, R.; Wong, A.T.; et al. Strain doping: Reversible single-axis control of a complex oxide lattice via helium implantation. Phys. Rev. Lett. 2015, 114, 256801. [Google Scholar] [CrossRef] [PubMed]
  16. Bhogra, A.; Masarrat, A.; Meena, R.; Hasina, D.; Bala, M.; Dong, C.-L.; Chen, C.-L.; Som, T.; Kumar, A.; Kandasami, A. Tuning the electrical and thermoelectric properties of N ion implanted SrTiO3 thin films and their conduction mechanisms. Sci. Rep. 2019, 9, 14486. [Google Scholar] [CrossRef]
  17. Jena, A.K.; Kulkarni, A.; Miyasaka, T. Halide perovskite photovoltaics: Background, status, and future prospects. Chem. Rev. 2019, 119, 3036–3103. [Google Scholar] [CrossRef]
  18. Dey, A.; Ye, J.; De, A.; Debroye, E.; Ha, S.K.; Bladt, E.; Kshirsagar, A.S.; Wang, Z.; Yin, J.; Wang, Y.; et al. State of the art and prospects for halide perovskite nanocrystals. ACS Nano 2021, 15, 10775–10981. [Google Scholar] [CrossRef]
  19. Grabowska, E. Selected perovskite oxides: Characterization, preparation and photocatalytic properties—A review. Appl. Catal. B Environ. 2016, 186, 97–126. [Google Scholar] [CrossRef]
  20. Bresolin, B.-M.; Park, Y.; Bahnemann, D.W. Recent progresses on metal halide perovskite-based material as potential photocatalyst. Catalysts 2020, 10, 709. [Google Scholar] [CrossRef]
  21. Hwang, J.; Rao, R.R.; Giordano, L.; Katayama, Y.; Yu, Y.; Shao-Horn, Y. Perovskites in catalysis and electrocatalysis. Science 2017, 358, 751–756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Wong, Y.-C.; Ng, J.D.A.; Tan, Z.-K. Perovskite-initiated photopolymerization for singly dispersed luminescent nanocomposites. Adv. Mater. 2018, 30, 1800774. [Google Scholar] [CrossRef] [PubMed]
  23. Haque, M.A.; Kee, S.; Villalva, D.R.; Ong, W.-L.; Baran, D. Halide perovskites: Thermal transport and prospects for thermoelectricity. Adv. Sci. 2020, 7, 1903389. [Google Scholar] [CrossRef] [Green Version]
  24. Wang, H.; Su, W.; Liu, J.; Wang, C. Recent development of N-type perovskite thermoelectrics. J. Mater. 2016, 2, 225–236. [Google Scholar] [CrossRef] [Green Version]
  25. Panda, D.; Tseng, T.-Y. Perovskite oxides as resistive switching memories: A review. Ferroelectrics 2014, 471, 23–64. [Google Scholar] [CrossRef]
  26. Rondinelli, J.M.; Spaldin, N.A. Structure and properties of functional oxide thin films: Insights from electronic-structure calculations. Adv. Mater. 2011, 23, 3363–3381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Klokholm, E.; Matthews, J.W.; Mayadas, A.F.; Angilello, J. Epitaxial strains and fracture in garnet films. AIP Conf. Proc. 1972, 5, 105–109. [Google Scholar] [CrossRef]
  28. Freund, L.B.; Suresh, S. Thin Film Materials: Stress, Defect Formation and Surface Evolution; Cambridge University Press: Cambridge, UK, 2004; ISBN 978-1-139-44982-3. [Google Scholar]
  29. Chaluvadi, S.K.; Ajejas, F.; Orgiani, P.; Lebargy, S.; Minj, A.; Flament, S.; Camarero, J.; Perna, P.; Méchin, L. Epitaxial strain and thickness dependent structural, electrical and magnetic properties of La0.67Sr0.33MnO3 films. J. Phys. Appl. Phys. 2020, 53, 375005. [Google Scholar] [CrossRef]
  30. Adamo, C.; Ke, X.; Wang, H.Q.; Xin, H.L.; Heeg, T.; Hawley, M.E.; Zander, W.; Schubert, J.; Schiffer, P.; Muller, D.A. Effect of biaxial strain on the electrical and magnetic properties of (001) La0.7Sr0.3MnO3 thin films. Appl. Phys. Lett. 2009, 95, 112504. [Google Scholar] [CrossRef]
  31. Yin, L.; Wang, C.; Shen, Q.; Zhang, L. Strain-induced curie temperature variation in La0.9Sr0.1MnO3 thin films. RSC Adv. 2016, 6, 96093–96102. [Google Scholar] [CrossRef]
  32. Chen, A.; Hu, J.-M.; Lu, P.; Yang, T.; Zhang, W.; Li, L.; Ahmed, T.; Enriquez, E.; Weigand, M.; Su, Q.; et al. Role of scaffold network in controlling strain and functionalities of nanocomposite films. Sci. Adv. 2016, 2, e1600245. [Google Scholar] [CrossRef] [Green Version]
  33. Berndt, L.M.; Balbarin, V.; Suzuki, Y. Magnetic anisotropy and strain states of (001) and (110) colossal magnetoresistance thin films. Appl. Phys. Lett. 2000, 77, 2903–2905. [Google Scholar] [CrossRef]
  34. Hrouda, F. Theoretical models of magnetic anisotropy to strain relationship revisited. Phys. Earth Planet. Inter. 1993, 77, 237–249. [Google Scholar] [CrossRef]
  35. Suzuki, Y.; Hwang, H.Y.; Cheong, S.-W.; van Dover, R.B. The role of strain in magnetic anisotropy of manganite thin films. Appl. Phys. Lett. 1997, 71, 140–142. [Google Scholar] [CrossRef]
  36. Nath, T.K.; Rao, R.A.; Lavric, D.; Eom, C.B.; Wu, L.; Tsui, F. Effect of three-dimensional strain states on magnetic anisotropy of La0.8Ca0.2MnO3 epitaxial thin films. Appl. Phys. Lett. 1999, 74, 1615–1617. [Google Scholar] [CrossRef]
  37. Huang, W.; Zhu, J.; Zeng, H.Z.; Wei, X.H.; Zhang, Y.; Li, Y.R. Strain induced magnetic anisotropy in highly epitaxial CoFe2O4 thin films. Appl. Phys. Lett. 2006, 89, 262506. [Google Scholar] [CrossRef]
  38. 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]
  39. Wang, H.S.; Wertz, E.; Hu, Y.F.; Li, Q.; Schlom, D.G. Role of strain in magnetotransport properties of Pr0.67Sr0.33MnO3 thin films. J. Appl. Phys. 2000, 87, 7409–7414. [Google Scholar] [CrossRef]
  40. Boikov, Y.A.; Gunnarsson, R.; Claeson, T. Strain-enhanced phase separation affecting electro-and magnetotransport in La0.67Ca0.33MnO3 films. J. Appl. Phys. 2004, 96, 435–442. [Google Scholar] [CrossRef]
  41. Yang, F.; Kemik, N.; Biegalski, M.D.; Christen, H.M.; Arenholz, E.; Takamura, Y. Strain engineering to control the magnetic and magnetotransport properties of La0.67Sr0.33MnO3 thin films. Appl. Phys. Lett. 2010, 97, 092503. [Google Scholar] [CrossRef] [Green Version]
  42. Streiffer, S.K.; Eastman, J.A.; Fong, D.D.; Thompson, C.; Munkholm, A.; Murty, M.R.; Auciello, O.; Bai, G.R.; Stephenson, G.B. Observation of nanoscale 180° stripe domains in ferroelectric PbTiO3 thin films. Phys. Rev. Lett. 2002, 89, 067601. [Google Scholar] [CrossRef]
  43. Yanase, N.Y.N.; Abe, K.A.K.; Fukushima, N.F.N.; Kawakubo, T.K.T. Thickness dependence of ferroelectricity in heteroepitaxial BaTiO3 thin film capacitors. Jpn. J. Appl. Phys. 1999, 38, 5305. [Google Scholar] [CrossRef]
  44. Abe, K.; Yanase, N.; Sano, K.; Izuha, M.; Fukushima, N.; Kawakubo, T. Modification of ferroelectricity in heteroepitaxial (Ba,Sr)TiO3 films for non-volatile memory applications. Integr. Ferroelectr. 1998, 21, 197–206. [Google Scholar] [CrossRef]
  45. Christen, H.M.; Specht, E.D.; Silliman, S.S.; Harshavardhan, K.S. Ferroelectric and antiferroelectric coupling in superlattices of paraelectric perovskites at room temperature. Phys. Rev. B 2003, 68, 020101. [Google Scholar] [CrossRef] [Green Version]
  46. Enriquez, E.; Li, Q.; Bowlan, P.; Lu, P.; Zhang, B.; Li, L.; Wang, H.; Taylor, A.J.; Yarotski, D.; Prasankumar, R.P.; et al. Induced ferroelectric phases in SrTiO3 by a nanocomposite approach. Nanoscale 2020, 12, 18193–18199. [Google Scholar] [CrossRef] [PubMed]
  47. Sangle, A.L.; Lee, O.J.; Kursumovic, A.; Zhang, W.; Chen, A.; Wang, H.; MacManus-Driscoll, J.L. Very high commutation quality factor and dielectric tunability in nanocomposite SrTiO3 thin films with Tc enhanced to >300 °C. Nanoscale 2018, 10, 3460–3468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Attfield, J.P. A simple approach to lattice effects in conducting perovskite-type oxides. Chem. Mater. 1998, 10, 3239–3248. [Google Scholar] [CrossRef]
  49. Dawson, J.A.; Sinclair, D.C.; Harding, J.H.; Freeman, C.L. A-site strain and displacement in Ba1– xCaxTiO3 and Ba1–xSrxTiO3 and the consequences for the curie temperature. Chem. Mater. 2014, 26, 6104–6112. [Google Scholar] [CrossRef] [Green Version]
  50. Azad, A.K.; Eriksson, S.-G.; Ivanov, S.A.; Mathieu, R.; Svedlindh, P. Influence of a-site cations on structural and magnetic properties in the double perovskites Ca2−xSrxMnWO6 and Sr2−xBaxMnWO6 (0 ≤ x ≤ 2.0). Ferroelectrics 2004, 302, 181–185. [Google Scholar] [CrossRef]
  51. Rondinelli, J.M.; May, S.J.; Freeland, J.W. Control of octahedral connectivity in perovskite oxide heterostructures: An emerging route to multifunctional materials discovery. MRS Bull. 2012, 37, 261–270. [Google Scholar] [CrossRef] [Green Version]
  52. Urushibara, A.; Moritomo, Y.; Arima, T.; Asamitsu, A.; Kido, G.; Tokura, Y. Insulator-metal transition and giant magnetoresistance in La1−xSrxMnO3. Phys. Rev. B 1995, 51, 14103–14109. [Google Scholar] [CrossRef] [PubMed]
  53. Uehara, M.; Mori, S.; Chen, C.H.; Cheong, S.-W. Percolative phase separation underlies colossal magnetoresistance in mixed-valent manganites. Nature 1999, 399, 560–563. [Google Scholar] [CrossRef]
  54. Raveau, B.; Maignan, A.; Martin, C.; Hervieu, M. Colossal magnetoresistance manganite perovskites:  Relations between crystal chemistry and properties. Chem. Mater. 1998, 10, 2641–2652. [Google Scholar] [CrossRef]
  55. Rao, C.N.R.; Arulraj, A.; Cheetham, A.K.; Raveau, B. Charge ordering in the rare earth manganates: The experimental situation. J. Phys. Condens. Matter 2000, 12, R83–R106. [Google Scholar] [CrossRef]
  56. Raveau, B. Impact of crystal chemistry upon the physics of strongly correlated electrons in oxides. Angew. Chem. Int. Ed. 2013, 52, 167–175. [Google Scholar] [CrossRef] [PubMed]
  57. Sinclair, D.C.; Attfield, J.P. The influence of A-cation disorder on the Curie temperature of ferroelectric ATiO3 perovskites. Chem. Commun. 1999, 16, 1497–1498. [Google Scholar] [CrossRef]
  58. Hubert, C.; Levy, J.; Carter, A.C.; Chang, W.; Kiechoefer, S.W.; Horwitz, J.S.; Chrisey, D.B. Confocal scanning optical microscopy of BaxSr1−xTiO3 thin films. Appl. Phys. Lett. 1997, 71, 3353–3355. [Google Scholar] [CrossRef]
  59. Gim, Y.; Hudson, T.; Fan, Y.; Kwon, C.; Findikoglu, A.T.; Gibbons, B.J.; Park, B.H.; Jia, Q.X. Microstructure and dielectric properties of Ba1−xSrxTiO3 films grown on LaAlO3 substrates. Appl. Phys. Lett. 2000, 77, 1200–1202. [Google Scholar] [CrossRef]
  60. Park, B.H.; Li, Y.R.; Xiong, J.; Jia, Q.X. Dielectric properties of epitaxial Ba1-XSrxTiO3 films on MgO substrates. Funct. Mater. Lett. 2011, 4, 41–44. [Google Scholar] [CrossRef]
  61. Islam, M.A.; Rondinelli, J.M.; Spanier, J.E. Normal mode determination of perovskite crystal structures with octahedral rotations: Theory and applications. J. Phys. Condens. Matter 2013, 25, 175902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. People, R.; Bean, J.C. Calculation of critical layer thickness versus lattice mismatch for GexSi1- x/Si strained-layer heterostructures. Appl. Phys. Lett. 1985, 47, 322–324. [Google Scholar] [CrossRef]
  63. Chen, Y.B.; Sun, H.P.; Katz, M.B.; Pan, X.Q.; Choi, K.J.; Jang, H.W.; Eom, C.B. Interface structure and strain relaxation in BaTiO3 thin films grown on GdScO3 and DyScO3 substrates with buried coherent SrRuO3 layer. Appl. Phys. Lett. 2007, 91, 252906. [Google Scholar] [CrossRef] [Green Version]
  64. Dimakis, E.; Iliopoulos, E.; Tsagaraki, K.; Adikimenakis, A.; Georgakilas, A. Biaxial strain and lattice constants of InN (0001) films grown by plasma-assisted molecular beam epitaxy. Appl. Phys. Lett. 2006, 88, 191918. [Google Scholar] [CrossRef]
  65. Janolin, P.E.; Anokhin, A.S.; Gui, Z.; Mukhortov, V.M.; Golovko, Y.I.; Guiblin, N.; Ravy, S.; El Marssi, M.; Yuzyuk, Y.I.; Bellaiche, L. Strain engineering of perovskite thin films using a single substrate. J. Phys. Condens. Matter 2014, 26, 292201. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, L.; Yuan, Y.; Lapano, J.; Brahlek, M.; Lei, S.; Kabius, B.; Gopalan, V.; Engel-Herbert, R. Continuously tuning epitaxial strains by thermal mismatch. ACS Nano 2018, 12, 1306–1312. [Google Scholar] [CrossRef] [PubMed]
  67. Rasic, D.; Sachan, R.; Prater, J.; Narayan, J. Structure-property correlations in thermally processed epitaxial LSMO films. Acta Mater. 2019, 163, 189–198. [Google Scholar] [CrossRef]
  68. Vailionis, A. The Effects of Strain on Crystal Structure and Properties during Epitaxial Growth of Oxides. In Epitaxial Growth of Complex Metal Oxides; Elsevier: Amsterdam, The Netherlands, 2015; pp. 175–207. ISBN 978-1-78242-245-7. [Google Scholar]
  69. Guzman, R.; Gazquez, J.; Rouco, V.; Palau, A.; Magen, C.; Varela, M.; Arbiol, J.; Obradors, X.; Puig, T. Strain-driven broken twin boundary coherence in YBa2Cu3O7−δ nanocomposite thin films. Appl. Phys. Lett. 2013, 102, 081906. [Google Scholar] [CrossRef] [Green Version]
  70. Arpaia, R.; Golubev, D.; Baghdadi, R.; Ciancio, R.; Dražić, G.; Orgiani, P.; Montemurro, D.; Bauch, T.; Lombardi, F. Transport properties of ultrathin YBa2Cu3O7 nanowires: A route to single-photon detection. Phys. Rev. B 2017, 96, 064525. [Google Scholar] [CrossRef] [Green Version]
  71. Jia, C.-L.; Mi, S.-B.; Jin, L. Quantitative HRTEM and its application in the study of oxide materials. Chin. Phys. B 2018, 27, 056803. [Google Scholar] [CrossRef]
  72. Liao, Z.; Huijben, M.; Zhong, Z.; Gauquelin, N.; Macke, S.; Green, R.J.; van Aert, S.; Verbeeck, J.; van Tendeloo, G.; Held, K.; et al. Controlled lateral anisotropy in correlated manganite heterostructures by interface-engineered oxygen octahedral coupling. Nat. Mater. 2016, 15, 425–431. [Google Scholar] [CrossRef] [PubMed]
  73. Kan, D.; Aso, R.; Sato, R.; Haruta, M.; Kurata, H.; Shimakawa, Y. Tuning magnetic anisotropy by interfacially engineering the oxygen coordination environment in a transition metal oxide. Nat. Mater. 2016, 15, 432–437. [Google Scholar] [CrossRef] [PubMed]
  74. Fang, Y.; Wei, H.; Dong, Q.; Huang, J. Quantification of re-absorption and re-emission processes to determine photon recycling efficiency in perovskite single crystals. Nat. Commun. 2017, 8, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. He, L.; Vanderbilt, D. First-principles study of oxygen-vacancy pinning of domain walls in PbTiO3. Phys. Rev. B 2003, 68, 134103. [Google Scholar] [CrossRef] [Green Version]
  76. Eichel, R.-A. Structural and dynamic properties of oxygen vacancies in perovskite oxides—Analysis of defect chemistry by modern multi-frequency and pulsed EPR techniques. Phys. Chem. Chem. Phys. 2010, 13, 368–384. [Google Scholar] [CrossRef] [PubMed]
  77. Herklotz, A.; Lee, D.; Guo, E.-J.; Meyer, T.L.; Petrie, J.R.; Lee, H.N. Strain coupling of oxygen non-stoichiometry in perovskite thin films. J. Phys. Condens. Matter 2017, 29, 493001. [Google Scholar] [CrossRef] [PubMed]
  78. Yang, Q.; Cao, J.X.; Zhou, Y.C.; Zhang, Y.; Ma, Y.; Lou, X.J. Tunable oxygen vacancy configuration by strain engineering in perovskite ferroelectrics from first-principles study. Appl. Phys. Lett. 2013, 103, 142911. [Google Scholar] [CrossRef]
  79. Al-Hamadany, R.; Goss, J.P.; Briddon, P.R.; Mojarad, S.A.; Al-Hadidi, M.; O’Neill, A.G.; Rayson, M.J. Oxygen vacancy migration in compressively strained SrTiO3. J. Appl. Phys. 2013, 113, 024108. [Google Scholar] [CrossRef] [Green Version]
  80. Ma, J.; Zhang, Y.; Wu, L.; Song, C.; Zhang, Q.; Zhang, J.; Nan, C.-W.; Ma, J. Strain-induced modulation of oxygen vacancies and magnetic properties in La0.5Sr0.5MnO3 thin films. MRS Commun. 2016, 6, 354–359. [Google Scholar] [CrossRef]
  81. Rajak, P.; Knez, D.; Chaluvadi, S.K.; Orgiani, P.; Rossi, G.; Méchin, L.; Ciancio, R. Evidence of Mn-ion structural displacements correlated with oxygen vacancies in La0.7Sr0.3MnO3 interfacial dead layers. ACS Appl. Mater. Interfaces 2021, 13, 55666–55675. [Google Scholar] [CrossRef] [PubMed]
  82. 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]
  83. Vasudevarao, A.; Kumar, A.; Tian, L.; Haeni, J.H.; Li, Y.L.; Eklund, C.-J.; Jia, Q.X.; Uecker, R.; Reiche, P.; Rabe, K.M.; et al. Multiferroic domain dynamics in strained strontium titanate. Phys. Rev. Lett. 2006, 97, 257602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Pertsev, N.A.; Tagantsev, A.K.; Setter, N. Phase transitions and strain-induced ferroelectricity in SrTiO3 epitaxial thin films. Phys. Rev. B 2000, 61, R825–R829. [Google Scholar] [CrossRef]
  85. Grünebohm, A.; Marathe, M.; Ederer, C. Ab initio phase diagram of BaTiO3 under epitaxial strain revisited. Appl. Phys. Lett. 2015, 107, 102901. [Google Scholar] [CrossRef]
  86. Pertsev, N.A.; Koukhar, V.G.; Waser, R.; Hoffmann, S. Effects of domain formation on the dielectric properties of ferroelectric thin films. Integr. Ferroelectr. 2001, 32, 235–249. [Google Scholar] [CrossRef]
  87. Li, Y.L.; Chen, L.Q. Temperature-strain phase diagram for BaTiO3 thin films. Appl. Phys. Lett. 2006, 88, 072905. [Google Scholar] [CrossRef] [Green Version]
  88. Pertsev, N.A.; Koukhar, V.G. Polarization instability in polydomain ferroelectric epitaxial thin films and the formation of heterophase structures. Phys. Rev. Lett. 2000, 84, 3722–3725. [Google Scholar] [CrossRef] [PubMed]
  89. Koukhar, V.G.; Pertsev, N.A.; Waser, R. Thermodynamic theory of epitaxial ferroelectric thin films with dense domain structures. Phys. Rev. B 2001, 64, 214103. [Google Scholar] [CrossRef] [Green Version]
  90. Fong, D.D.; Stephenson, G.B.; Streiffer, S.K.; Eastman, J.A.; Auciello, O.; Fuoss, P.H.; Thompson, C. Ferroelectricity in ultrathin perovskite films. Science 2004, 304, 1650–1653. [Google Scholar] [CrossRef] [PubMed]
  91. Junquera, J.; Ghosez, P. Critical thickness for ferroelectricity in perovskite ultrathin films. Nature 2003, 422, 506–509. [Google Scholar] [CrossRef] [PubMed]
  92. Pertsev, N.A.; Zembilgotov, A.G.; Tagantsev, A.K. Effect of mechanical boundary conditions on phase diagrams of epitaxial ferroelectric thin films. Phys. Rev. Lett. 1998, 80, 1988–1991. [Google Scholar] [CrossRef]
  93. Sheng, G.; Zhang, J.X.; Li, Y.L.; Choudhury, S.; Jia, Q.X.; Liu, Z.K.; Chen, L.Q. Domain stability of PbTiO3 thin films under anisotropic misfit strains: Phase-field simulations. J. Appl. Phys. 2008, 104, 054105. [Google Scholar] [CrossRef]
  94. Lee, H.N.; Christen, H.M.; Chisholm, M.F.; Rouleau, C.M.; Lowndes, D.H. Strong polarization enhancement in asymmetric three-component ferroelectric superlattices. Nature 2005, 433, 395–399. [Google Scholar] [CrossRef] [PubMed]
  95. Ogawa, Y.; Yamada, H.; Ogasawara, T.; Arima, T.; Okamoto, H.; Kawasaki, M.; Tokura, Y. Nonlinear magneto-optical Kerr rotation of an oxide superlattice with artificially broken symmetry. Phys. Rev. Lett. 2003, 90, 217403. [Google Scholar] [CrossRef] [PubMed]
  96. Park, B.H.; Peterson, E.J.; Jia, Q.X.; Lee, J.; Zeng, X.; Si, W.; Xi, X.X. Effects of very thin strain layers on dielectric properties of epitaxial Ba0.6Sr0.4TiO3 films. Appl. Phys. Lett. 2001, 78, 533–535. [Google Scholar] [CrossRef]
  97. Tenne, D.A.; Bruchhausen, A.; Lanzillotti-Kimura, N.D.; Fainstein, A.; Katiyar, R.S.; Cantarero, A.; Soukiassian, A.; Vaithyanathan, V.; Haeni, J.H.; Tian, W.; et al. Probing nanoscale ferroelectricity by ultraviolet Raman spectroscopy. Science 2006, 313, 1614–1616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Lee, J.H.; Fang, L.; Vlahos, E.; Ke, X.; Jung, Y.W.; Kourkoutis, L.F.; Kim, J.-W.; Ryan, P.J.; Heeg, T.; Roeckerath, M.; et al. A strong ferroelectric ferromagnet created by means of spin–lattice coupling. Nature 2010, 466, 954–958. [Google Scholar] [CrossRef]
  99. Lin, Y.; Dai, C.; Li, Y.R.; Chen, X.; Chen, C.L.; Bhalla, A.; Jia, Q.X. Strain relaxation in epitaxial (Pb,Sr)TiO3 thin films on NdGaO3 substrates. Appl. Phys. Lett. 2010, 96, 102901. [Google Scholar] [CrossRef]
  100. Harrington, S.A.; Zhai, J.; Denev, S.; Gopalan, V.; Wang, H.; Bi, Z.; Redfern, S.A.T.; Baek, S.-H.; Bark, C.W.; Eom, C.-B.; et al. Thick lead-free ferroelectric films with high curie temperatures through nanocomposite-induced strain. Nat. Nanotechnol. 2011, 6, 491–495. [Google Scholar] [CrossRef] [PubMed]
  101. Sun, S.; Yang, K.-Y.; Wang, C.-M.; Juan, T.-K.; Chen, W.T.; Liao, C.Y.; He, Q.; Xiao, S.; Kung, W.-T.; Guo, G.-Y.; et al. High-efficiency broadband anomalous reflection by gradient meta-surfaces. Nano Lett. 2012, 12, 6223–6229. [Google Scholar] [CrossRef] [PubMed]
  102. Chen, A.; Jia, Q.X. A pathway to desired functionalities in vertically aligned nanocomposites and related architectures. MRS Bull. 2021, 46, 115–122. [Google Scholar] [CrossRef]
  103. Ni, Y.; Khachaturyan, A.G. From chessboard tweed to chessboard nanowire structure during pseudospinodal decomposition. Nat. Mater. 2009, 8, 410–414. [Google Scholar] [CrossRef] [PubMed]
  104. Ni, Y.; Rao, W.; Khachaturyan, A.G. Pseudospinodal mode of decomposition in films and formation of chessboard-like nanostructure. Nano Lett. 2009, 9, 3275–3281. [Google Scholar] [CrossRef] [PubMed]
  105. Moshnyaga, V.; Damaschke, B.; Shapoval, O.; Belenchuk, A.; Faupel, J.; Lebedev, O.I.; Verbeeck, J.; van Tendeloo, G.; Mücksch, M.; Tsurkan, V.; et al. Structural phase transition at the percolation threshold in epitaxial (La0.7Ca0.3MnO3)1–x:(MgO)x nanocomposite films. Nat. Mater. 2003, 2, 247–252. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, W.; Fan, M.; Li, L.; Chen, A.; Su, Q.; Jia, Q.X.; MacManus-Driscoll, J.L.; Wang, H. Heterointerface design and strain tuning in epitaxial BiFeO3:CoFe2O4 nanocomposite films. Appl. Phys. Lett. 2015, 107, 212901. [Google Scholar] [CrossRef]
  107. Zheng, H.; Kreisel, J.; Chu, Y.-H.; Ramesh, R.; Salamanca-Riba, L. Heteroepitaxially enhanced magnetic anisotropy in BaTiO3–CoFe2O4 nanostructures. Appl. Phys. Lett. 2007, 90, 113113. [Google Scholar] [CrossRef]
  108. Zhang, W.; Chen, A.; Jian, J.; Zhu, Y.; Chen, L.; Lu, P.; Jia, Q.X.; MacManus-Driscoll, J.L.; Zhang, X.; Wang, H. Strong perpendicular exchange bias in epitaxial La0.7Sr0.3MnO3:BiFeO3 nanocomposite films through vertical interfacial coupling. Nanoscale 2015, 7, 13808–13815. [Google Scholar] [CrossRef]
  109. Zhao, R.; Li, W.; Lee, J.H.; Choi, E.M.; Liang, Y.; Zhang, W.; Tang, R.; Wang, H.; Jia, Q.X.; MacManus-Driscoll, J.L.; et al. Precise tuning of (YBa2Cu3O7-δ)1−x:(BaZrO3)x thin film nanocomposite structures. Adv. Funct. Mater. 2014, 24, 5240–5245. [Google Scholar] [CrossRef]
  110. MacManus-Driscoll, J.L.; Zerrer, P.; Wang, H.; Yang, H.; Yoon, J.; Fouchet, A.; Yu, R.; Blamire, M.G.; Jia, Q.X. Strain control and spontaneous phase ordering in vertical nanocomposite heteroepitaxial thin films. Nat. Mater. 2008, 7, 314–320. [Google Scholar] [CrossRef] [PubMed]
  111. Chen, A.; Zhang, W.; Khatkhatay, F.; Su, Q.; Tsai, C.-F.; Chen, L.; Jia, Q.X.; MacManus-Driscoll, J.L.; Wang, H. Magnetotransport properties of quasi-one-dimensionally channeled vertically aligned heteroepitaxial nanomazes. Appl. Phys. Lett. 2013, 102, 093114. [Google Scholar] [CrossRef]
  112. Chen, A.; Zhang, W.; Jian, J.; Wang, H.; Tsai, C.-F.; Su, Q.; Jia, Q.X.; MacManus-Driscoll, J.L. Role of boundaries on low-field magnetotransport properties of La0.7Sr0.3MnO3-Based nanocomposite thin films. J. Mater. Res. 2013, 28, 1707–1715. [Google Scholar] [CrossRef]
  113. MacManus-Driscoll, J.L.; Suwardi, A.; Wang, H. Composite epitaxial thin films: A new platform for tuning, probing, and exploiting mesoscale oxides. MRS Bull. 2015, 40, 933–942. [Google Scholar] [CrossRef]
  114. Chen, A.; Bi, Z.; Hazariwala, H.; Zhang, X.; Su, Q.; Chen, L.; Jia, Q.X.; MacManus-Driscoll, J.L.; Wang, H. Microstructure, magnetic, and low-field magnetotransport properties of self-assembled (La0.7Sr0.3MnO3)0.5:(CeO2)0.5 vertically aligned nanocomposite thin films. Nanotechnology 2011, 22, 315712. [Google Scholar] [CrossRef]
  115. Lee, S.; Zhang, W.; Khatkhatay, F.; Jia, Q.X.; Wang, H.; MacManus-Driscoll, J.L. Strain tuning and strong enhancement of ionic conductivity in SrZrO3–RE2O3 (RE = Sm, Eu, Gd, Dy, and Er) nanocomposite films. Adv. Funct. Mater. 2015, 25, 4328–4333. [Google Scholar] [CrossRef]
  116. Li, W.; Zhang, W.; Wang, L.; Gu, J.; Chen, A.; Zhao, R.; Liang, Y.; Guo, H.; Tang, R.; Wang, C.; et al. Vertical interface induced dielectric relaxation in nanocomposite (BaTiO3)1−x:(Sm2O3)x thin films. Sci. Rep. 2015, 5, 11335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Lee, S.; Sangle, A.; Lu, P.; Chen, A.; Zhang, W.; Lee, J.S.; Wang, H.; Jia, Q.X.; MacManus-Driscoll, J.L. Novel electroforming-free nanoscaffold memristor with very high uniformity, tunability, and density. Adv. Mater. 2014, 26, 6284–6289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Zheng, H.; Wang, J.; Lofland, S.E.; Ma, Z.; Mohaddes-Ardabili, L.; Zhao, T.; Salamanca-Riba, L.; Shinde, S.R.; Ogale, S.B.; Bai, F.; et al. Multiferroic BaTiO3-CoFe2O4 nanostructures. Science 2004, 303, 661–663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Oses, C.; Toher, C.; Curtarolo, S. High-entropy ceramics. Nat. Rev. Mater. 2020, 5, 295–309. [Google Scholar] [CrossRef]
  120. Enriquez, E.; Chen, A.; Harrell, Z.; Dowden, P.; Koskelo, N.; Roback, J.; Janoschek, M.; Chen, C.; Jia, Q.X. Oxygen vacancy-tuned physical properties in perovskite thin films with multiple B-site valance states. Sci. Rep. 2017, 7, 46184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Burschka, J.; Pellet, N.; Moon, S.-J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M.K.; Grätzel, M. Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature 2013, 499, 316–319. [Google Scholar] [CrossRef]
  122. Wang, Y.; Sato, N.; Fujino, T. Synthesis of BaZrS3 by short time reaction at lower temperatures. J. Alloys Compd. 2001, 327, 104–112. [Google Scholar] [CrossRef]
  123. Sarmiento-Pérez, R.; Cerqueira, T.F.T.; Körbel, S.; Botti, S.; Marques, M.A.L. Prediction of stable nitride perovskites. Chem. Mater. 2015, 27, 5957–5963. [Google Scholar] [CrossRef]
  124. Leppert, L.; Reyes-Lillo, S.E.; Neaton, J.B. Electric field- and strain-induced Rashba effect in hybrid halide perovskites. J. Phys. Chem. Lett. 2016, 7, 3683–3689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Chen, Y.; Lei, Y.; Li, Y.; Yu, Y.; Cai, J.; Chiu, M.-H.; Rao, R.; Gu, Y.; Wang, C.; Choi, W.; et al. Strain engineering and epitaxial stabilization of halide perovskites. Nature 2020, 577, 209–215. [Google Scholar] [CrossRef] [PubMed]
  126. Moloney, E.G.; Yeddu, V.; Saidaminov, M.I. Strain engineering in halide perovskites. ACS Mater. Lett. 2020, 2, 1495–1508. [Google Scholar] [CrossRef]
  127. Kim, H.-S.; Park, N.-G. Importance of tailoring lattice strain in halide perovskite crystals. NPG Asia Mater. 2020, 12, 1–14. [Google Scholar] [CrossRef]
  128. Jones, T.W.; Osherov, A.; Alsari, M.; Sponseller, M.; Duck, B.C.; Jung, Y.-K.; Settens, C.; Niroui, F.; Brenes, R.; Stan, C.V. Lattice strain causes non-radiative losses in halide perovskites. Energy Environ. Sci. 2019, 12, 596–606. [Google Scholar] [CrossRef] [Green Version]
  129. Liu, D.; Luo, D.; Iqbal, A.N.; Orr, K.W.P.; Doherty, T.A.S.; Lu, Z.-H.; Stranks, S.D.; Zhang, W. Strain analysis and engineering in halide perovskite photovoltaics. Nat. Mater. 2021, 20, 1337–1346. [Google Scholar] [CrossRef]
Figure 1. (a) The ideal ABO3 perovskite crystal structure showing tilt in all three directions. (b) Distortion of [BO6] octahedra along various directions, lowering the symmetry of the cubic structure and forming other crystal structures. The positive sign indicates in-phase rotation (c+) and the negative sign shows out-of-phase rotation (c), both about z-axis. Reprinted with permission from ref. [26]. Copyright 2011, Wiley.
Figure 1. (a) The ideal ABO3 perovskite crystal structure showing tilt in all three directions. (b) Distortion of [BO6] octahedra along various directions, lowering the symmetry of the cubic structure and forming other crystal structures. The positive sign indicates in-phase rotation (c+) and the negative sign shows out-of-phase rotation (c), both about z-axis. Reprinted with permission from ref. [26]. Copyright 2011, Wiley.
Nanomaterials 12 00835 g001
Figure 2. (a) The isovalent phase diagrams of rare-earth manganites with A-site composition exhibit changes in the octahedral connectivity (B-O-B bond angles, upper panels) and symmetry. Changes to the octahedral connectivity (horizontal arrows) affect the temperatures associated with magnetic and charge-ordering phase transitions, whereas periodic orderings of the octahedral shape and size (vertical arrows) occur at the orbital-ordering transition. Reprinted with permission from ref. [51]. Copyright © 2022, The Materials Research Society. (b) Lattice parameters for La1−xSrxMnO3 crystals at room temperature. (c) Electronic phase diagram of La1−xSrxMnO3. Open circles and filled triangles are the Neel (TN) and Curie (TC) temperatures, respectively. The abbreviations mean paramagnetic insulator (P.I.), paramagnetic metal (P.M.), spin-canted insulator (CN.I.), ferromagnetic insulator (FI), and ferromagnetic metal (FM). Reprinted with permission from ref. [52]. Copyright 1995, American Institute of Physics.
Figure 2. (a) The isovalent phase diagrams of rare-earth manganites with A-site composition exhibit changes in the octahedral connectivity (B-O-B bond angles, upper panels) and symmetry. Changes to the octahedral connectivity (horizontal arrows) affect the temperatures associated with magnetic and charge-ordering phase transitions, whereas periodic orderings of the octahedral shape and size (vertical arrows) occur at the orbital-ordering transition. Reprinted with permission from ref. [51]. Copyright © 2022, The Materials Research Society. (b) Lattice parameters for La1−xSrxMnO3 crystals at room temperature. (c) Electronic phase diagram of La1−xSrxMnO3. Open circles and filled triangles are the Neel (TN) and Curie (TC) temperatures, respectively. The abbreviations mean paramagnetic insulator (P.I.), paramagnetic metal (P.M.), spin-canted insulator (CN.I.), ferromagnetic insulator (FI), and ferromagnetic metal (FM). Reprinted with permission from ref. [52]. Copyright 1995, American Institute of Physics.
Nanomaterials 12 00835 g002
Figure 3. (a) Crystal structure of an unstrained perovskite. (b,c) Schematic illustration of an epitaxial perovskite film grown on a perovskite single-crystal substrate showing: (b) biaxiall compression, (c) biaxial tension, (d) the a-axis lattice constant in angstroms of some perovskites and perovskite-related materials of interest. The substrates can be cubic, pseudotetragonal or pseudocubic. Reprinted with permission from ref. [9]. Copyright 2014, Cambridge University Press.
Figure 3. (a) Crystal structure of an unstrained perovskite. (b,c) Schematic illustration of an epitaxial perovskite film grown on a perovskite single-crystal substrate showing: (b) biaxiall compression, (c) biaxial tension, (d) the a-axis lattice constant in angstroms of some perovskites and perovskite-related materials of interest. The substrates can be cubic, pseudotetragonal or pseudocubic. Reprinted with permission from ref. [9]. Copyright 2014, Cambridge University Press.
Nanomaterials 12 00835 g003
Figure 4. Schematic representation of octahedral rotational patterns in La0.67Sr0.33MnO3. (a,b) Under compressive stress. (c,d) Under tensile stress. (a,c) are the side view and (b,d) are the top view of the unit cells. Reprinted with permission from ref. [68]. Copyright 2015, Elsevier Ltd. Lattice modulations: (e) Observed (blue) and simulated (red) XRD profiles around LSMO (220) reflections for films grown on (La, Sr)(Al, Ta)O3 (LSAT) and NdGaO3 (NGO) substrates. (f) Schematic drawing of the lattice modulations used in calculations. (g) Schematic picture of a reciprocal space showing the substrate peak (red) together with the LSMO (220) peak and its first-order satellites (blue). Here, Q = 4πsin(θ/λ), where θ is the Bragg angle and λ = 1.540598 Å. Reprinted with permission from ref. [38]. Copyright 2011, American Physical Society.
Figure 4. Schematic representation of octahedral rotational patterns in La0.67Sr0.33MnO3. (a,b) Under compressive stress. (c,d) Under tensile stress. (a,c) are the side view and (b,d) are the top view of the unit cells. Reprinted with permission from ref. [68]. Copyright 2015, Elsevier Ltd. Lattice modulations: (e) Observed (blue) and simulated (red) XRD profiles around LSMO (220) reflections for films grown on (La, Sr)(Al, Ta)O3 (LSAT) and NdGaO3 (NGO) substrates. (f) Schematic drawing of the lattice modulations used in calculations. (g) Schematic picture of a reciprocal space showing the substrate peak (red) together with the LSMO (220) peak and its first-order satellites (blue). Here, Q = 4πsin(θ/λ), where θ is the Bragg angle and λ = 1.540598 Å. Reprinted with permission from ref. [38]. Copyright 2011, American Physical Society.
Nanomaterials 12 00835 g004
Figure 5. (a) Resistivity vs. temperature behavior, at zero applied magnetic field, of 22 nm thick La0.7Sr0.3MnO3 films on different substrates (10 nm thick on NdScO3). (b) The temperature dependence of the magnetization normalized at 10 K of samples cooled in a 1000 Oe field of H = 1000 Oe. Circles indicate the Curie temperature, TC. (c) The Curie temperature TC vs. the εB and ε* strains. The best fit plane to the data is also shown. Reprinted with permission from ref. [30]. Copyright 2009, American Institute of Physics.
Figure 5. (a) Resistivity vs. temperature behavior, at zero applied magnetic field, of 22 nm thick La0.7Sr0.3MnO3 films on different substrates (10 nm thick on NdScO3). (b) The temperature dependence of the magnetization normalized at 10 K of samples cooled in a 1000 Oe field of H = 1000 Oe. Circles indicate the Curie temperature, TC. (c) The Curie temperature TC vs. the εB and ε* strains. The best fit plane to the data is also shown. Reprinted with permission from ref. [30]. Copyright 2009, American Institute of Physics.
Nanomaterials 12 00835 g005
Figure 6. (a) Formation energies (Ef) of different types of oxygen vacancy under different strains. The black, red, and blue symbols represent Ef of V c u d , V c s w , and V a b   , respectively. The formation of the “up-down” pattern along the x-axis causes the sudden decrease in energy of Vab when 2% tensile strain is applied. (b) The relative energy stabilities and the polarization patterns of different oxygen vacancy configurations in 3% compressive-strained PbTiO3. Reprinted with permission from ref. [65]. Copyright 2013, American Institute of Physics. (c) Thickness dependence of HC (right axis) and HEB (left axis). For comparison, inset shows HC and HEB of La0.5Sr0.5MnO3 films grown on the LaAlO3 substrate with the thickness of 8 and 40 nm (d) The thickness dependence of Mn4+/(Mn3+ + Mn4+) ratio in La0.5Sr0.5MnO3 films grown on the SrTiO3 substrate. The red dots correspond to the data of La0.5Sr0.5MnO3 films grown on the LaAlO3 substrate for comparison. Reprinted with permission from ref. [67]. Copyright 2016, Materials Research Society.
Figure 6. (a) Formation energies (Ef) of different types of oxygen vacancy under different strains. The black, red, and blue symbols represent Ef of V c u d , V c s w , and V a b   , respectively. The formation of the “up-down” pattern along the x-axis causes the sudden decrease in energy of Vab when 2% tensile strain is applied. (b) The relative energy stabilities and the polarization patterns of different oxygen vacancy configurations in 3% compressive-strained PbTiO3. Reprinted with permission from ref. [65]. Copyright 2013, American Institute of Physics. (c) Thickness dependence of HC (right axis) and HEB (left axis). For comparison, inset shows HC and HEB of La0.5Sr0.5MnO3 films grown on the LaAlO3 substrate with the thickness of 8 and 40 nm (d) The thickness dependence of Mn4+/(Mn3+ + Mn4+) ratio in La0.5Sr0.5MnO3 films grown on the SrTiO3 substrate. The red dots correspond to the data of La0.5Sr0.5MnO3 films grown on the LaAlO3 substrate for comparison. Reprinted with permission from ref. [67]. Copyright 2016, Materials Research Society.
Nanomaterials 12 00835 g006
Figure 7. (a) Expected shift in TC of (100) SrTiO3 with biaxial in-plane strain, based on thermodynamic analysis. The arrows indicate the predicted direction of the polarization for strained SrTiO3: in-plane for biaxial tensile strain and out-of-plane for biaxial compressive strain. The εs values for SrTiO3 fully constrained (commensurate) to the lattice constants of LSAT and (110) DyScO3 substrates are indicated by the positions of the corresponding arrows. The cross shows the observed TC shift of a 500 Å thick SrTiO3 film epitaxially grown on (110) DyScO3. Adapted with permission from ref. [13]. Copyright 2004, Springer Nature. (b) Expected TC of (001) BaTiO3 under biaxial in-plane strain (εs), based on thermodynamic analysis. The green region represents the range (error bars) in the predicted TC resulting from the spread in reported property coefficients for BaTiO3 that enter into the thermodynamic analysis. The data points show the observed εs and TC values of coherent BaTiO3 films grown by MBE on GdScO3 (blue circle) and DyScO3 (red circle) substrates and by PLD on GdScO3 (blue square) and DyScO3 (red square) substrates. Adapted with permission from ref. [11]. Copyright 2004, AAAS.
Figure 7. (a) Expected shift in TC of (100) SrTiO3 with biaxial in-plane strain, based on thermodynamic analysis. The arrows indicate the predicted direction of the polarization for strained SrTiO3: in-plane for biaxial tensile strain and out-of-plane for biaxial compressive strain. The εs values for SrTiO3 fully constrained (commensurate) to the lattice constants of LSAT and (110) DyScO3 substrates are indicated by the positions of the corresponding arrows. The cross shows the observed TC shift of a 500 Å thick SrTiO3 film epitaxially grown on (110) DyScO3. Adapted with permission from ref. [13]. Copyright 2004, Springer Nature. (b) Expected TC of (001) BaTiO3 under biaxial in-plane strain (εs), based on thermodynamic analysis. The green region represents the range (error bars) in the predicted TC resulting from the spread in reported property coefficients for BaTiO3 that enter into the thermodynamic analysis. The data points show the observed εs and TC values of coherent BaTiO3 films grown by MBE on GdScO3 (blue circle) and DyScO3 (red circle) substrates and by PLD on GdScO3 (blue square) and DyScO3 (red square) substrates. Adapted with permission from ref. [11]. Copyright 2004, AAAS.
Nanomaterials 12 00835 g007
Figure 8. Four strain-phase diagrams of (001) pseudocubic-oriented PbTiO3 calculated using thermodynamic analysis or phase-field simulations and different assumptions of the ferroelectric domain states. (a) Single domain for all ferroelectric states. Adapted with permission from ref. [92]. Copyright 1998, American Physical Society. (b) Either single- or double-domain states with domain–wall orientations restricted to be 45° from the film/substrate interface. Adapted with permission from ref. [88]. Copyright 2000, American Physical Society. (c) Single- or double-domain states with domain–wall orientations restricted to be either 45° or 90° from the film/substrate interface. Adapted with permission from ref. [89]. Copyright 2001, American Physical Society. First-order phase transitions are shown by thick lines in (ac). (d) From three-dimensional phase-field simulations that automatically predict the possible multidomain states without assuming the domain–wall orientations. Adapted with permission from ref. [93]. Copyright 2008, American Institute of Physics.
Figure 8. Four strain-phase diagrams of (001) pseudocubic-oriented PbTiO3 calculated using thermodynamic analysis or phase-field simulations and different assumptions of the ferroelectric domain states. (a) Single domain for all ferroelectric states. Adapted with permission from ref. [92]. Copyright 1998, American Physical Society. (b) Either single- or double-domain states with domain–wall orientations restricted to be 45° from the film/substrate interface. Adapted with permission from ref. [88]. Copyright 2000, American Physical Society. (c) Single- or double-domain states with domain–wall orientations restricted to be either 45° or 90° from the film/substrate interface. Adapted with permission from ref. [89]. Copyright 2001, American Physical Society. First-order phase transitions are shown by thick lines in (ac). (d) From three-dimensional phase-field simulations that automatically predict the possible multidomain states without assuming the domain–wall orientations. Adapted with permission from ref. [93]. Copyright 2008, American Institute of Physics.
Nanomaterials 12 00835 g008
Figure 9. (a) P(E) curves of S2B4C2 (Pr ≈ 16.5 µC cm−2) and S10B10C10 (Pr ≈ 3.5 µC cm−2) at 1000 kV cm−1. (Note that this electric field is higher than that applied to measure the polarization of the BaTiO3 film, whose polarization is near saturation at 400 kV cm−1, whereas the S2B4C2 requires a higher electric field (∼700 kV cm−1)). (b), ε(E) curves of S2B4C2 (black line), S2B6C2 (red line), S4B2C2 (green line) and S10B10C10 (light blue line) showing differing degrees of asymmetry. (c) In-plane (open squares) and out-of-plane (filled squares) lattice parameters of various superlattices. c′ corresponds to the supercell c divided by the number of constituent perovskites in a supercell. (d), Pr values from E = ±750 kV cm−1 loops. The partially relaxed S2B8C2 structure was measured with E = ±650 kV cm−1 because of its lower breakdown strength. (e) Diagrams of supercells showing the different local environments possible for the TiO6-octahedra (bound by the same or different A-site cations). Heterointerfacial TiO6 octahedra are shaded in grey and indicated by solid black arrows. Reprinted with permission from ref. [94]. Copyright 2010, Springer Nature.
Figure 9. (a) P(E) curves of S2B4C2 (Pr ≈ 16.5 µC cm−2) and S10B10C10 (Pr ≈ 3.5 µC cm−2) at 1000 kV cm−1. (Note that this electric field is higher than that applied to measure the polarization of the BaTiO3 film, whose polarization is near saturation at 400 kV cm−1, whereas the S2B4C2 requires a higher electric field (∼700 kV cm−1)). (b), ε(E) curves of S2B4C2 (black line), S2B6C2 (red line), S4B2C2 (green line) and S10B10C10 (light blue line) showing differing degrees of asymmetry. (c) In-plane (open squares) and out-of-plane (filled squares) lattice parameters of various superlattices. c′ corresponds to the supercell c divided by the number of constituent perovskites in a supercell. (d), Pr values from E = ±750 kV cm−1 loops. The partially relaxed S2B8C2 structure was measured with E = ±650 kV cm−1 because of its lower breakdown strength. (e) Diagrams of supercells showing the different local environments possible for the TiO6-octahedra (bound by the same or different A-site cations). Heterointerfacial TiO6 octahedra are shaded in grey and indicated by solid black arrows. Reprinted with permission from ref. [94]. Copyright 2010, Springer Nature.
Nanomaterials 12 00835 g009
Figure 10. Schematic illustration (not to scale) of a vertically aligned epitaxial nanocomposite film on a substrate: (a) an array of vertically aligned nanopillar-like material A with a feature size d entrenched in a matrix of material B and (b) the lattice–strain interactions across the vertical interface. For simplicity, both phases are shown to be strained equally, but in different strain states (compression or tension).
Figure 10. Schematic illustration (not to scale) of a vertically aligned epitaxial nanocomposite film on a substrate: (a) an array of vertically aligned nanopillar-like material A with a feature size d entrenched in a matrix of material B and (b) the lattice–strain interactions across the vertical interface. For simplicity, both phases are shown to be strained equally, but in different strain states (compression or tension).
Nanomaterials 12 00835 g010
Figure 11. (a) The pillar-size-dependent vertical strain in the nanocomposite film with a fixed pillar volume. In this case, ε is proportional to 1/d. (b) Pillar-volume-dependent vertical strain in the nanocomposite film with a fixed pillar size. In this case, ε is proportional to the pillar volume V when V is below 20–30 percent. (c) Illustrations of growth temperature, pillar size and strain state in the nanocomposites. Reprinted with permission from ref. [102]. Copyright 2021, Springer Nature.
Figure 11. (a) The pillar-size-dependent vertical strain in the nanocomposite film with a fixed pillar volume. In this case, ε is proportional to 1/d. (b) Pillar-volume-dependent vertical strain in the nanocomposite film with a fixed pillar size. In this case, ε is proportional to the pillar volume V when V is below 20–30 percent. (c) Illustrations of growth temperature, pillar size and strain state in the nanocomposites. Reprinted with permission from ref. [102]. Copyright 2021, Springer Nature.
Nanomaterials 12 00835 g011
Figure 12. X-ray diffraction θ–2θ scans of SrTiO3:MgO vertically aligned nanocomposite films deposited on SrTiO3 substrate before and after etching off MgO nanopillars. Reprinted with permission from ref. [11]. Copyright 2019, Wiley.
Figure 12. X-ray diffraction θ–2θ scans of SrTiO3:MgO vertically aligned nanocomposite films deposited on SrTiO3 substrate before and after etching off MgO nanopillars. Reprinted with permission from ref. [11]. Copyright 2019, Wiley.
Nanomaterials 12 00835 g012
Figure 13. (a) A cross-sectional STEM image of an LSMO:MgO nanocomposite with phase-separated LSMO (white) and MgO (dark) phases. Inset: A top-view STEM image of the LSMO:MgO film with MgO nanopillars in the LSMO matrix. The MgO volume is 22%. (b) The strain distribution in the LSMO matrix and MgO nanopillars (MgO volume: 41%) calculated by phase-field simulation. (c) The second-phase volume-dependent vertical strain in the LSMO matrix. The red (dot) data represent the LSMO:MgO system, and the blue (square) data set represent the LSMO:ZnO system. (d) The second-phase volume-dependent resistance at 100 K (red dataset) and saturated magnetization (blue dataset) at 20 K for the LSMO:MgO (solid curve) and LSMO:ZnO (dashed curves) nanocomposite films. Reprinted with permission from ref. [32]. Copyright 2016, AAAS.
Figure 13. (a) A cross-sectional STEM image of an LSMO:MgO nanocomposite with phase-separated LSMO (white) and MgO (dark) phases. Inset: A top-view STEM image of the LSMO:MgO film with MgO nanopillars in the LSMO matrix. The MgO volume is 22%. (b) The strain distribution in the LSMO matrix and MgO nanopillars (MgO volume: 41%) calculated by phase-field simulation. (c) The second-phase volume-dependent vertical strain in the LSMO matrix. The red (dot) data represent the LSMO:MgO system, and the blue (square) data set represent the LSMO:ZnO system. (d) The second-phase volume-dependent resistance at 100 K (red dataset) and saturated magnetization (blue dataset) at 20 K for the LSMO:MgO (solid curve) and LSMO:ZnO (dashed curves) nanocomposite films. Reprinted with permission from ref. [32]. Copyright 2016, AAAS.
Nanomaterials 12 00835 g013
Table 1. Summary of vertical lattice strain reported in different material systems with direct lattice-matching between the constituents A and B. For the direct lattice-matching of strained lattices, m lattices of phase A match with K lattices of phase B (m and k are positive integer numbers). The data was taken with permission from [32]. Copyright 2016, American Association for the Advancement of Science.
Table 1. Summary of vertical lattice strain reported in different material systems with direct lattice-matching between the constituents A and B. For the direct lattice-matching of strained lattices, m lattices of phase A match with K lattices of phase B (m and k are positive integer numbers). The data was taken with permission from [32]. Copyright 2016, American Association for the Advancement of Science.
Systemabulk [Å]K|f| [%]zz| [%]Reference
LSMO:MgO3.87:4.2118.412.0[32]
LCMO:MgO3.86:4.2118.672.1[105]
BFO:CFO3.96:8.3925.761.0[106]
BTO:CFO4.04:8.3923.761.6[107]
BFO:LSMO3.96:3.8712.291.3[108]
BZO:YBCO11.679:4.19337.701.0[109]
Table 2. Summary of vertical lattice strain reported in different materials systems with domain-matching between the constituents A and B. The data was taken with permission from [32]. Copyright 2016, American Association for the Advancement of Science.
Table 2. Summary of vertical lattice strain reported in different materials systems with domain-matching between the constituents A and B. The data was taken with permission from [32]. Copyright 2016, American Association for the Advancement of Science.
Systemabulk [Å]mm:m + 1|f| [%]zz| [%]Reference
LSMO (001):ZnO (110)3.87:3.245.145:60.46≈0.5, 1.0[110,111]
LSMO (111):ZnO (0001)6.70:5.213.503:4 + 4:58.5 × 10−30[112,113]
CeO2:LSMO *5.41:3.872.512:3 + 3:40.13<0.1[114]
SrZrO3:Gd2O3 **4.09:2.701.942:30.980.9[115]
BTO:Sm2O34.04:2.732.082:31.172.35[116]
BFO:Sm2O33.96:2.732.222:3 + 2:3 + 3:41.71≈1.4[110]
STO:Sm2O33.905:2.732.322:3 + 2:3 + 3:40.31[117]
* The calculated m value is 2.51 for CeO2:LSMO; therefore, m can be valued as 2 or 3. Therefore, both 2:3 and 3:4 matchings exist and align alternatively (50%:50%). In the STO:Sm2O3 system, m is 2.32. m can be either 2 or 3. Both 2:3 and 3:4 matchings exist with a frequency of 66% 2:3 and 34% 3:4. In the SrZrO3:Gd2O3 system, the calculated m is 1.94; therefore, the m is set to 2 for domain-matching. ** Bulk lattice constant of Gd2O3 is 10.80 Å. The plane spacing for Gd2O3 (004) is 10.80 Å/4 = 2.7 Å.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dhole, S.; Chen, A.; Nie, W.; Park, B.; Jia, Q. Strain Engineering: A Pathway for Tunable Functionalities of Perovskite Metal Oxide Films. Nanomaterials 2022, 12, 835. https://doi.org/10.3390/nano12050835

AMA Style

Dhole S, Chen A, Nie W, Park B, Jia Q. Strain Engineering: A Pathway for Tunable Functionalities of Perovskite Metal Oxide Films. Nanomaterials. 2022; 12(5):835. https://doi.org/10.3390/nano12050835

Chicago/Turabian Style

Dhole, Samyak, Aiping Chen, Wanyi Nie, Baeho Park, and Quanxi Jia. 2022. "Strain Engineering: A Pathway for Tunable Functionalities of Perovskite Metal Oxide Films" Nanomaterials 12, no. 5: 835. https://doi.org/10.3390/nano12050835

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