Review

Potassium sodium niobate (KNN) lead-free piezoceramics: A review of phase boundary engineering based on KNN materials

  • Received: 19 July 2023 Revised: 22 August 2023 Accepted: 29 August 2023 Published: 18 September 2023
  • Lead zirconia titanate (PZT) is the most often used piezoelectric material in various electronic applications like energy harvesters, ultrasonic capacitors and motors. It is true that PZT has a lot of significant drawbacks due to its 60% lead content, despite its outstanding ferroelectric, dielectric and piezoelectric properties which influenced by PZT's morphotropic phase boundary. The recently found potassium sodium niobate (KNN) is one of the most promising candidates for a new lead-free piezoelectric material. For the purpose of providing a resource and shedding light on the future, this paper provides a summary of the historical development of different phase boundaries in KNN materials and provides some guidance on how to achieve piezoelectric activity on par with PZT through a thorough examination and critical analysis of relevant articles by providing insight and perspective of KNN, which consists of detailed evaluation of the design, construction of phase boundaries and engineering for applications.

    Citation: Hidayah Mohd Ali Piah, Mohd Warikh Abd Rashid, Umar Al-Amani Azlan, Maziati Akmal Mohd Hatta. Potassium sodium niobate (KNN) lead-free piezoceramics: A review of phase boundary engineering based on KNN materials[J]. AIMS Materials Science, 2023, 10(5): 835-861. doi: 10.3934/matersci.2023045

    Related Papers:

    [1] Valeriy Ishchuk, Danil Kuzenko, Vladimir Sobolev . Piezoelectric and functional properties of materials with coexisting ferroelectric and antiferroelectric phases. AIMS Materials Science, 2018, 5(4): 711-741. doi: 10.3934/matersci.2018.4.711
    [2] Wentao Qu, Qian Zhang, Guibian Li, Boyang Pan, Haiying Liu . Microstructure and thermophysical properties of NiTiZr phase change alloys for downhole tool heat storage. AIMS Materials Science, 2025, 12(1): 85-100. doi: 10.3934/matersci.2025007
    [3] Roman Koleňák, Igor Kostolný . Investigation of microstructure and wettability of selected lead-free solders for higher application temperatures. AIMS Materials Science, 2018, 5(5): 889-901. doi: 10.3934/matersci.2018.5.889
    [4] Frank Tietz . Phase relations of NASICON materials and compilation of the quaternary phase diagram Na2O-P2O5-SiO2-ZrO2. AIMS Materials Science, 2017, 4(6): 1305-1318. doi: 10.3934/matersci.2017.6.1305
    [5] Murat Aycibin, Naciye ECE . First-principles calculation of the electronic and optical properties of BiRhO3 compound. AIMS Materials Science, 2017, 4(4): 894-904. doi: 10.3934/matersci.2017.4.894
    [6] Ruby Maria Syriac, A.B. Bhasi, Y.V.K.S Rao . A review on characteristics and recent advances in piezoelectric thermoset composites. AIMS Materials Science, 2020, 7(6): 772-787. doi: 10.3934/matersci.2020.6.772
    [7] Sergey V. Belim . Study of ordering in 2D ferromagnetic nanoparticles arrays: Computer simulation. AIMS Materials Science, 2023, 10(6): 948-964. doi: 10.3934/matersci.2023051
    [8] Fahmi Astuti, Rima Feisy Azmi, Mohammad Arrafi Azhar, Fani Rahayu Hidayah Rayanisaputri, Muhammad Redo Ramadhan, Malik Anjelh Baqiya, Darminto . Employing Na2CO3 and NaCl as sources of sodium in NaFePO4 cathode: A comparative study on structure and electrochemical properties. AIMS Materials Science, 2024, 11(1): 102-113. doi: 10.3934/matersci.2024006
    [9] Faiza Lourdjane, Mouhyddine Kadi-Hanifi, Azzeddine Abderrahmane Raho . Precipitation of the metastable phases in a tin microalloyed Al-10at%Ag alloy. AIMS Materials Science, 2017, 4(1): 1-15. doi: 10.3934/matersci.2017.1.1
    [10] Mariya Aleksandrova . Spray deposition of piezoelectric polymer on plastic substrate for vibrational harvesting and force sensing applications. AIMS Materials Science, 2018, 5(6): 1214-1222. doi: 10.3934/matersci.2018.6.1214
  • Lead zirconia titanate (PZT) is the most often used piezoelectric material in various electronic applications like energy harvesters, ultrasonic capacitors and motors. It is true that PZT has a lot of significant drawbacks due to its 60% lead content, despite its outstanding ferroelectric, dielectric and piezoelectric properties which influenced by PZT's morphotropic phase boundary. The recently found potassium sodium niobate (KNN) is one of the most promising candidates for a new lead-free piezoelectric material. For the purpose of providing a resource and shedding light on the future, this paper provides a summary of the historical development of different phase boundaries in KNN materials and provides some guidance on how to achieve piezoelectric activity on par with PZT through a thorough examination and critical analysis of relevant articles by providing insight and perspective of KNN, which consists of detailed evaluation of the design, construction of phase boundaries and engineering for applications.



    Piezoelectric material lead zirconia titanate (PZT) is widely used in a variety of electronic devices, including energy harvesters, ultrasonic capacitors and motors, because of its excellent ferroelectric, dielectric and piezoelectric properties. Even if there are a wide variety of applications and their market size is growing, the quest for new applications continues in the industries that are related with it. In part, this is owing to the rapid advancement of alternative technologies, which puts the existing market at risk of being overtaken at any time. It is true that despite its impressive electrical properties, PZT has several severe disadvantages. It contains over 60% lead [1]. According to the Comprehensive Environmental Response, Compensation, and Liability Act of 1980, lead is one of the most toxic substances that might potentially cause damage to the neurological system as well as the development of spermatozoid. Lead is on the list of most dangerous compounds because of its cancer-causing properties.

    This paper will give a rundown of some of the more recent applications of lead-based and lead-free piezoceramics, as well as look ahead to possible future uses for these materials. For example, a great deal of work has been done to discover new types of piezoelectric materials. Potassium sodium niobate (KNN) is one of the most promising prospects for a new lead-free piezoelectric material that has been discovered as a result of these efforts. As recently as 2004, when Saito et al. reported on the development of KNN-based piezoceramics, the aim seemed to be within reach [2]. KNbO3 (KN) and NaNbO3 perovskites were combined to generate KNN, a new perovskite material (NN). In the A site, potassium and sodium ions are present, while niobium ions are present in the B site. Curie temperature was 420 ℃, dielectric constant was 700, remanent polarisation was 14 μC/cm2, low coercive field was 140 kV/cm2 and high piezoelectric constant was 1.6 volts/cm [1,3,4].

    In comparison to PZT, the piezoelectric characteristics of pure KNN are considerably lacking. Notably, this current work shows an unwavering commitment to the KNN-based system, whereas multiple outstanding evaluations have compared alternative lead-free piezoceramic in comparison [5,6,7,8,9]. There have been hundreds of articles in the previous decades about the development of KNN ceramics, indicating undeniable interest in this system as the most promising possibility for commercially viable lead-free piezoceramics candidates. A great deal of research has been devoted to the creation of KNN-based piezoelectrics and exceptionally promising piezoelectric qualities have been reported [10,11,12]. Li and co-workers wrote and published a comprehensive study of KNN-based materials focusing on their fundamental properties, processing processes, property change and beneficial applications. A comprehensive evaluation of the design, construction of phase boundaries and improvement of piezoelectric characteristics in KNN-based materials was published by Wu and colleagues [13]. Phase barriers in KNN-based materials have been studied extensively by the authors and they gave viable ideas for increasing piezoelectric activity in KNN-based materials [14]. KNN utilising doped KNN to improve its physical and electrical performance is also being studied by the researchers [15]. There has been a recent pattern of progressive growth in the number of papers discussing materials based on alkali niobium. Among these articles, the building of phase boundaries has emerged as a key method for the enhancement of the electrical properties of these materials. As a consequence of this, phase boundaries play a crucial part in the overall development of alkali niobium lead-free piezoelectrics [13]. According to the findings in this paper, the current state of KNN-based ceramics has been thoroughly assessed by a thorough examination and a critical analysis of relevant articles. As a result, we would be thrilled if this study might serve as a useful resource and shed light on the future development of KNN-based materials.

    In a macroscopically nonpolar ceramic, the employment of ferroelectric materials enables one to generate a unique polar axis by a so-called poling process, as explained by Jaffe et al. (1958). Inducing a piezoelectric effect in this ferroelectric material was revealed to be far more powerful than in previously studied single crystal piezoelectrics. Consequently, the success of developing the first polycrystalline piezoelectrics prompted a thorough search for improved piezoceramics, which ultimately resulted in the development of PZT ceramic solid solutions in the 1950s. During the poling process, it is generally agreed upon that a high degree of alignments of ferroelectric dipoles can be driven by a large amount of thermodynamically equivalent states when the material is subjected to a driving electric field, which makes it simple to generate enhanced electrical properties [16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39]. In the field of piezoelectric materials, PZT's morphotropic phase boundary (MPB) construction is regarded a typical case where the crystal structure undergoes a sudden transition and the piezoelectricity of the mixture is optimised at MPB [17,22,33,40].

    The discovery of MPB-induced property enhancement has set a new standard for the future discovery of piezoceramics. The results of these studies have been published in a number of academic references. The change of phase boundaries is a potent instrument for the promotion of electrical properties of piezoelectric materials, the involved phase boundaries and their corresponding types are substantially responsible for the augmentation in piezoelectric activity, regardless of whether a material is lead based or lead free [10,15,16,17,27,28,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72] is because the involved phase boundaries and their corresponding types.

    There are three different types of phase transitions known to exist in KNN materials, corresponding to the temperatures at which the material changes from rhombohedral to orthorhombic (TR-O), orthorhombic to tetragonal (TO-T) and rhombohedral to tetragonal (TR-T), respectively [6,73]. The temperature dependency of the individual dielectric peaks provides a clear illustration of TR-O, TO-T and TC. The R-O or O-T phase boundary in KNN may be created at room temperature if the TR-O or TO-T temperature was shifted to room temperature by the use of additives. On the other hand, the phase boundary that corresponds to R-O or O-T possesses the polymorphic phase transition (PPT) characteristic [74,75,76,77] and it is dependent not only on the compositions but also on the temperatures, as shown in Figure 1. This is the case when R-O or O-T corresponds to a transition between two different types of phases. When the compositions of a piezomaterial are located at the phase boundaries, the polarisation of the material can be rotated more easily between different symmetries, which can result in an improvement in the material's dielectric and piezoelectric properties. This is true regardless of whether the material possesses MPB or PPT characteristics.

    Figure 1.  (a) Polymorphic phase transition (b) morphotropic phase boundary [13].

    However, recent developments suggest that there is only a small improvement in the d33 values for KNN materials having O-T or R-O phase boundaries [12,37,38,76,77,78], which is inferior to the vast majority of PZT ceramics [3] or even textured KNN [10]. Since polarisation axis rotations are easily produced in compositions close to an R-T boundary, such boundaries may be required to further boost the piezoelectric activity of KNN [79,80]. This assumption has also been recently confirmed in potassium-sodium niobate piezoceramics by experimental methods [81,82,83,84,85,86] and a greater d33 value has been reached by creating R-T phase boundaries. In order to form R-T phase barrier, two or more additives such as Ta and Sb and few other additives must be used to simultaneously lower their TR-O and raise their TO-T values toward room temperature [71,87,88,89]. New concepts can thus be applied to the development of high-performance KNN-based materials.

    A material's phase boundaries can be influenced by its chemical composition or other parameters [90,91,92,93,94] that define its electrical properties. Materials with compositions around the phase boundaries (e.g., R-O, R-T, or O-T) include Bi0.5Na0.5TiO3, and (K, Na)NbO3. Higher d33 values are often achievable with BaTiO3-based materials, but their low TC makes them unsuitable for use in high-temperature environments. However, these materials can be used to make actuators that can function at temperatures below 100 ℃. The low "depoling temperature" of Bi0.5Na0.5TiO3-based materials, like that of BaTiO3-based materials, limits their practical applications [95,96,97,98,99]. However, Jo et al. (2011), have created a huge strain that is equivalent or even superior to PZT. Figure 2a, b shows that the phase boundary types and the composition itself determine the width of the d33 and TC distributions in KNN-based materials (e.g., d33 = 171–490 pC/N, TC = 178–475 ℃) [96]. In particular, by creating new R-T phase barriers, we were able to make significant progress in the piezoelectricity of KNN-based ceramics. The gap between lead-free and lead-based materials has been narrowed thanks to the development of novel phase boundaries that can be used to improve the piezoelectricity of KNN materials with all information taken directly from the sources cited. As can be seen in Figure 2a, PNN-based and relaxor PT single crystals with an MPB [93,94] exhibit a massive d33 value, whereas PZT-based materials exhibit a broad d33 distribution that is sensitive to the doping components. Recent research on PNN-based single crystal relaxors has shown a piezoelectric coefficient of 3500 pC/N [100]. Doping a relaxor ferroelectric crystal with rare-earth elements allowed for the measurement of this value. This study is the latest one on the subject, having been completed in 2022. On the other hand, it is likely that fresh research has been carried out since then, which might provide information that is more up to date about the piezoelectric coefficient for PNN-based single crystal relaxor.

    Figure 2.  Value of d33 vs Tc of (a) lead-based (b) lead-free ceramics from number of researches [13].

    The phase boundary of KNN is the composition and temperature transition between distinct crystal structures. It is essential to have an understanding of the phase boundary in order to modify the properties of KNN-based materials to accommodate a variety of applications. Structure analysis methods including X-ray diffraction (XRD) and neutron diffraction are commonly used to determine where the phase boundary lies in KNN. Diffraction patterns can reveal the crystal structure of a material, allowing for the determination of its various phases. For instance, XRD was used in a study to determine the crystal structure and phase transitions of KNN perovskites [101]. For instance, the crystal structure of A-site substituted perovskite KNN compounds was determined using high-resolution XRD and neutron diffraction. Finding a phase-coexistence zone at the monoclinic-tetragonal phase transition provides evidence of a first-order phase barrier, as reported in previous research [102]. Researchers showed that XRD measurements may be used to distinguish between the various phases of KNN perovskites and to identify the region of transition between them TO-T and TT-C were detected at 236 and 440 ℃, indicating a trend towards lower temperature side for phase transition temperatures [103].

    However, identifying the temperature and composition regions where the phase transition occurs is essential for characterising the phase boundary. This data is essential for figuring out how KNN-based materials react in various environments. The phase boundary of KNN has been the subject of several studies that have utilised a variety of methods to characterise it. Examining KNN-based ceramic's piezoelectric and strain characteristics at varying compositions and temperatures is one of the methods. Zheng et al. (2015) looked at the conflicts between high strain and high piezoelectricity in KNN materials and proposed the creation of new phase boundaries to resolve the conflicts [104]. Using temperature-dependent XRD patterns, a study describes the structural evolution of the rhombohedral-tetragonal (R-T) phase boundary in KNN-based ceramics. Rhombohedral, rhombohedral-orthorhombic and rhombohedral-tetragonal phases were all identified in the study [105]. Adding dopants or substitute components to KNN is another way to shift the phase boundary. KNN's orthorhombic-rhombohedral phase transition temperature was shown to be controllable via dopant concentration when barium zirconate was included into the material. This research showed that the compositional adjustments can be used to manipulate the phase boundary of KNN. In order to fully appreciate its characteristics and modify its behaviour for specific applications, it is crucial to locate and characterise the phase boundary of KNN. Crystal structure and symmetry of KNN at various compositions and temperatures are typically determined by structural analysis methods including XRD and neutron diffraction. The phase boundary can be characterised by analysing the effects of dopants or substitute elements on the phase transition and taking measurements of the piezoelectric and strain properties of KNN-based ceramics. These investigations aid in the creation of high-quality KNN-based materials.

    In KNN-based piezoceramics, the polymorphic phase boundary (PPB) has been the primary focus of research, but there have been some attempts to actualize the MPB, or rhombohedral vs. tetragonal phase boundary, found in PZT. The primary motivation for doing so is to realise the lowest temperature rhombohedral phase in KNN under ambient conditions. KNN is an isotropic solid solution between the orthorhombic symmetry ferroelectric KN and the antiferroelectric NN [13,106,107]. KNN is quite different from PZT in that its phase diagram is extremely intricate, with three vertical MPBs between two distinct ferroelectric orthorhombic phases [108]. The one with a K content below 50% is particularly relevant. KNN has a high Curie temperature (420 ℃), however its piezoelectric characteristics are subpar while it is at ambient temperature. Hot-pressed dense samples have been measured to have a maximum d33 of 160 pC/N and a k2 of 0.3 [43]. As a result, the majority of previous research has concentrated on improving these functional qualities through chemical modifications. In this regard, BaZrO3 was shown to be effective by Wang et al. [109]. There has been further research in this area as a result of this, leading to the empirical conclusion that Zr4 appears to play a significant role [12,110]. These studies confirmed that KNN-based materials are capable of being fabricated into MPB similar to PZT. However, the induced MPB is likely to be skewed in the phase diagram of temperature versus composition. This MPB contradicts the group subgroup relationship since it is more akin to the PPB between rhombohedral and tetragonal. To improve the formulation of truly PZT-like MPB in KNN, that is, a vertical MPB between rhombohedral and tetragonal, further studies have been conducted on the idea that increasing the rhombohedral field in the PZT-like MPB compositions would make the existing rhombohedral symmetry more thermally-stable [110]. Since MPB existing at room temperature endures up to 250 ℃ with fairly strong functional qualities, it suggests that the notion is working despite its apparent simplicity. However, testing of this MPB's range needs to be done at temperatures at least 40 ℃ below room temperature. Here, we primarily examine how various phase boundaries in KNN materials have evolved over time and offer some guidance on how to achieve piezoelectric activity on par with PZT. Additionally, this article summarises the extensive discussions on how to build phase boundaries by chemical modifications.

    Since a significant advancement in KNN-based materials, the O-T phase boundary has become the most well-known. One of the most contentious aims of research into KNN based materials is the establishment of the O-T phase border [111,112,113]. The original breakthrough in its d33 value occurred in 2004 when O-T phase coexistence was brought down to or near room temperature through the use of additives. In KNN-based textured ceramics, for instance, the coexistence of O and T phases at room temperature and the use of a new preparation technique led to a high d33 value of 416 pC/N, while LiNbO3 or LiTaO3-modiffed KNN ceramics achieved enhanced piezoelectric activity by constructing the same phase boundary using conventional solid-state methods. These pioneering works have stimulated a surge in interest in lead-free piezoelectrics around the world and as a result, scientists have focused considerable effort on fostering piezoelectric activity by erecting O-T phase boundaries. Several previous attempts to construct O-T phase boundaries in KNN-based ceramics in order to induce enhancement of its piezoelectric activity have met with success. These include ion substitutions, solid solutions with other ABO3-type perovskites [98,99] or ABO3 multicomponents [114,115,116,117].

    Perovskite ABO3 materials, such as Bi0.5A0.5TiO3 (A = Na+ [118,119], K+ [120], Li+ [121]), BTiO3 (B = Ba2+, Sr2+ [122], Ca2+ [123], (Ba0.95Sr0.05)2+ [124], [(Bi0.5Na0.5)0.94Ba0.06]2+ [125] or BiMO3 (M = Sc3+ [126], Al3+ [127], Fe3+ [114], Co3+ [115]) and others could lowering TO-T values. ABO3 doping can efficiently stimulate the establishment of O-T phase barriers resulting in increased of piezoelectricity [124]. Even while the addition of ABO3-type additives can improve the d33 values of KNN-based materials [115], the resulting values are still lower than those of KNN ceramics including Li+, Ta5+ and Sb5+ [68,128,129]. However, ABO3's low price makes it a possible choice for real-world applications. KNN ceramics doped with ABO3 multicomponents commonly possess improved piezoelectric activity (d33 = 220−305 pC/N) combined with high TC values of > 300 ℃, as indicated in Table 1. The results reveal that O-T phase barriers can occur when the TO-T transition temperature of a material is shifted to room temperature via doping with most ABO3 multicomponents.

    Table 1.  Material system of KNN.
    KNN material system d33 (pC/N) kp Tc (℃) Ref.
    0.97(K0.5Na0.5)NbO3-0.03(Bi0.5Na0.5)TiO3 195 0.43 375 [118]
    (1−x)K0.5Na0.5NbO3-0.06BiFeO3 146 0.51 405 [130]
    K0.5Na0.5NbO3-LiTaO3-0.1MnO2 251 29.5 330–350 [131]
    0.96(K0.5Na0.5)0.95Li0.05Nb0.93Sb0.07O3-0.04CaZrO3 420 - - [132]
    0.993(K, Na)NbO3-0.007BiFeO3 220 0.46 - [132]
    (K0.5Na0.5)NbO3-0.01Bi(Ni2/3Nb1/3)O3 150 - 300–400 [133]
    (0.97−x)K0.40Na0.60Nb0.95Sb0.05O3-0.03Bi0.5K0.5HfO3-0.01SrZrO3 280 0.364 - [134]

     | Show Table
    DownLoad: CSV

    Doping with BiFeO3 or BiScO3 is particularly promising for enhancing d33 in BiMeO3 (Me = Fe, Sc, Al, etc.) additives and d33 values of > 250 pC/N are frequently produced in such ternary systems via the creation of O-T phase barriers [113,135]. Variations in the magnitude of the increase in d33 are largely attributable to the rates at which T-O and T are being shifted. For example, Jiang et al. studied KNN-LiSbO3-BiFeO3 ternary systems in order to shift the TO-T close to room temperature, finding that low quantities of BiFeO3 (0.4 mol%) enhance the d33 values from 202 to 260 pC/N [135]. This means that BiFeO3, as an ABO3 component, can effectively increase the piezoelectric activity of KNNLiSbO3 ceramics by generating O-T phase barriers at or near room temperature. Pure tetragonal structures were obtained in the ternary system K0.5Na0.5NbO3LiSbO3BiScO3, demonstrating a d33 value of 305 pC/N which the researchers reasoned that the phase boundary could not have played a significant role in the enhanced piezoelectric capabilities [136]. In contrast, we discovered that BiScO3 addition increases TR-O values and decreases TO-T values of KNN [81,82], allowing for the formation of the new R-T phase boundary at ambient temperature. Furthermore, at room temperature or lower, the dielectric constant of this material gradually increases, which may indicate the presence of phase barriers. Therefore, more research into this material system is required to determine where the increased piezoelectricity comes from [136].

    For Bi0.5K0.5Zr0.9Hf0.103 (BKZH) doped KNN based piezoceramics, Li and co-workers analysed the effects of BKZH doping in 2017 [137]. As the BKZH concentration increased, they found that the TO-T decreased and the TR-O increased. Therefore, BKZH doping aids in achieving room-temperature coexistence of various phases. This aids in increasing the piezoelectric coefficient (d33), to the value of 451 pC/N. There are three different phases that can occur, such as rhombohedral, orthorhombic and tetragonal. They also discovered that adding 1 mol% BKZH lowered the TC by 20 ℃. Another work by Li et al. [138] looked at the impacts of ZrO3 (BLKZ) as a dopant on (Bi0.45La0.05K0.5) and observed comparable effects on phase transition temperatures. The incorporation of BLKZ resulted in the formation of a new phase boundary at room temperature between the rhombohedral, orthorhombic and tetragonal phases. However, the KNN-BLKZ phase structure is only tetragonal when BLKZ is equal to or more than 5 mol%. At d33 = 385 pC/N and TC = 245 ℃, KNN-0.035BLKZ performs best. BiNaZrAO3 type dopants are used by scientists to lower TO-T and raise TR-O, but they also lower TC. In 2020, Qiao et al. [139] employed BiNaZrHfO3 and found that the orthorhombic phase boundary could be reached at TC = 235 ℃. Shi et al. [140] looked into BiSmNaZrO3 (BSNZ) with the aim of achieving high piezo characteristics with an R-O-T phase structure. The group managed to get d33 = 508 pC/N and TC = 268 ℃. In addition, CaBiNaZrO3 was utilised as a dopant by Pan et al. [141]. They discovered that the dopant does not lower the curie temperature in CxBNZ when the Ca content is less than or equivalent to 0.4. Dopants such as Ba0.5Ca0.5Zr0.5Hf0.5O3 (BCZH) [142] and Ba(Li0.4W0.6)O3 (BLW) [143] reduce TO-T but also reduce TC.

    Previous studies using XRD patterns discovered that the KNL-(N0.9−xTSx) ceramic system has both orthorhombic symmetry (O) and tetragonal symmetry (T). The tetragonal symmetry-related peaks become more prominent in high-Sb content samples. A larger concentration of Sb5+ creates a structural alteration that modifies the T/O content in the structure, while on the one hand, the substitutional addition of Sb5+ should lower the cell (owing to the smaller ionic radius) shifting the peaks towards higher angles. It has been shown by other researchers that small cations in B-sites allow for the coexistence of different phases (from rhombohedral-tetragonal to orthorhombic-tetragonal) at room temperature, thereby lowering the TO-T towards room temperature [144,145].

    Meanwhile, field emission scanning electron microscopy (FESEM) images show that the average grain size drops from 1.9 µm at x = 0.00 to 1.3 µm at x = 0.04, as Sb addition increases. Moreover, a nonhomogeneous microstructure is observed for the lowest Sb concentration (x = 0.02). As a matter of fact, the poor additive distribution over the entire sample causes grains with antimony to exhibit a restriction in their normal grain growth, while grains without or with low antimony concentration display grain sizes similar to the pure sample. Finally, at the greatest antimony content (x = 0.04), the improved additive distribution throughout the entire sample results in a more uniform microstructure with a smaller average grain size. Several parameters, including grain size, final microstructure, dielectric permittivity and density values, may have contributed to the identical values recorded for samples with x = 0.02 and 0.04 [146].

    Improving piezoelectric characteristics has been viewed as a functional benefit of stabilising an O-T coexistence phase boundary around room temperature. In this case, the KNL-(N0.9−xTSx) ceramic's ferroelectric properties will shed light on the implications of the Sb5+ substitution. Previous studies established that as the Sb5+ dopant concentration increased, the orthorhombic and tetragonal phases stabilised close to R-T (making a threshold of multiphasic coexistence where the tetragonal phase is predominant), leading to enhanced ferroelectric properties at x = 0.02. It is well established that various phases exhibit polarisation in distinct directions. Since there are six possible polarisation directions in the tetragonal phase and twelve in the orthorhombic phase, when their coexistence is stabilised, a total of 18 polarisation directions are viable. Ferroelectric behaviour due to spontaneous polarisation was consistently observed in all samples when tested in a strong external electric field [147].

    Therefore, the study showed that the O-T phase coexistence phase boundary could be stabilised by substituting Sb ions in the perovskite structure, which also causes an increased degree of polarisation directions near room temperature, easing the polarisation process of the system, and where the average grain size also increases, encouraging the increased domain size and favouring more domain walls.

    The Curie temperature of undoped KNN is more than 400 ℃ (TC). Both the rhombohedral-to-orthorhombic (TR-O) and the orthorhombic-to-tetragonal (O-T) transition temperatures are present (TO-T). On the other hand, there are two major drawbacks of KNN-based lead-free ceramics. The piezoelectricity is low, only 80–180 pC/N [2,142,143,148,149,150,151]. Low sintered densification is another issue that makes manufacturing challenging since the R-O phase boundary of KNN occurs at relatively cold temperatures (about −123 ℃) [128,129] and yields a subpar d33 which is frequently disregarded. The TR-O values of KNN have recently been boosted with the addition of additives such as Sb5+ [75,151,152], Ta5+ [153], AZrO3 (A = Ba2+, Sr2+, Ca2+) [138,139], BiScO3 [154], etc. For example, the coexistence of R-O ferroelectric mixed phases has been shown to improve d33 (230 pC/N) and the TR-O value of (K0.48Na0.52)(Nb, Sb)O3 can be controlled to near room temperature by refining the Sb concentration [75]. In addition, it was discovered that doping with Ta5+ may reduce TC and TO-T while simultaneously increasing TR-O of KNN [153], in a manner that was comparable to that of Sb5+ [75]. When AZrO3 (A = Ba2+, Sr2+, or Ca2+) is added to a pure KNN ceramic, the TR-O temperature is lowered to room temperature [109]. The R-O phase boundary and the types of chemical modifications also influence their d33 values, but increasing the d33 typically necessitates reducing the TC [12,75,154]. It may be possible to build new phase boundaries (e.g., R-T) by simultaneously shifting TO-T and TR-O close to room temperature, despite the fact that R-O KNN materials have poor overall performance of d33 and TC.

    In a particular research paper, CaZrO3 was used as a modifier to develop (K0.48Na0.48Li0.04)(Nb0.95Sb0.05)O3 lead free ceramics. The CZ-KNLNS powders have a cubic structure and their particles average between 300 and 700 nm in size. It is clear that the addition of CaZrO3 slows down particle growth because the particle size reduces when the amount of CaZrO3 is increased. As the ceramic part becomes denser, the grain size decreases and the grain boundaries become undetectable. CaZrO3 may have accumulated at the grain borders, inhibiting growth and causing the decrease in grain size. Research demonstrated that by adjusting the CaZrO3 concentration, the phase constitution and domain structure of CZ-KNLNS ceramics may be modified. The presence of rhombohedral (R) phase characteristics at x = 0.06 indicates a steady decrease in the quantity of orthorhombic phase and an increase in the amount of R phase. This region is where the orthorhombic and rhombohedral (O-R) phases can coexist. Doping ions of different radii are introduced into the KNLNS ceramic lattice, with the smaller ion replacing the A site and the bigger ion replacing the B site. This leads to a smaller average ionic radius for the A-site and a larger average ionic radius for the B-site, leading the R phase to develop in the KNLNS at temperatures close to room temperature. Multi-phase coexistence is thought to play a significant role in the high quality of KNLNS based ceramics [155]. The findings showed the coexistence zone of orthorhombic-rhombohedral (O-R) phase can be thought of as occurring between x = 0.03 and x = 0.06, as the phase structure changed from orthorhombic to rhombohedral with increasing CaZrO3 content. The ceramics exhibit a coexisting rhombohedral and orthorhombic phase structure at x = 0.04. The piezoelectricity of a material can be enhanced by applying a polarised electric field, as the electric domain of a multi-phase coexisting structure can be switched more easily than in a single-phase structure [155,156]. The enhanced piezoelectric characteristics are a result of the combination of a dense homogenous microstructure and a stable domain configuration at x = 0.04.

    Considering that the R-T phase boundary of PZT exhibits large d33 and a weak temperature dependency of piezoelectric capabilities [3], we wondered if a similar R-T phase boundary might be produced in KNN materials to further enhance their piezoelectric properties and temperature stability. Several studies suggest that it is possible to engineer a R-T phase boundary in KNN by controlling the establishment of R-O and O-T phase borders via doping in additives [12,32,75,154,157,158,159,160,161,162]. It is possible to raise TR-O and decrease TO-T of KNN materials by adding ions or ABO3 materials (e.g., Sb5+ [75], Ta5+ [153], BaZrO3 [160], CaZrO3 [161], SrZrO3 [161]). As a result, it is probable that the R-T phase barrier should be produced in KNN materials by changing the composition ratios of the doped additives. Doping with various additives in earlier investigations led to the eventual induction of the R-T phase boundary in KNN. Since then, a number of material systems based on KNN have been produced that have R-T phase borders [80,81,83,139,140]. By fabricating the R-T phase boundary we found that the high d33 value in 0.96(K0.5Na0.5)0.95Li0.05Nb1−xSbxO3-0.04BaZrO3 ceramics reported by researchers in 2013 (425 pC/N) matched the results reported for textured KNN-based ceramics (416 pC/N) [10,163,164,165]. In this work, Sb5+ and BaZrO3 are employed together to improve the TR-O and reduce the TO-T of KNN materials [75,165]. Despite the huge d33 (344–425 pC/N), ceramics typically have TC values of 200 ℃ [71,89,104]. This means that for efficient phase boundary design, both a larger TC and a larger d33 in KNN are necessary [83,85,166].

    Doping CZrO3 (C = Ba2+, Sr2+, Ca2+) has previously been reported to create three distinct phase transition temperatures (i.e., TC, TR-O, and TO-T) in KNN [145]. Furthermore, Zr4+ plays a significant influence in the phase transition temperatures (particularly TR-O). Ramajo et al. (2013) observed that adding Zr4+ softened the material's ferroelectric characteristics and improved its density, even though effects of CZrO3 on TC, TR-O and TO-T are mostly independent of A site type [82,104,151,165,167]. It has been reported before that KNN materials' TO-T can be changed by adding Bi3+. Therefore, the R-T phase boundary can be formed by changing the TR-O and TO-T of KNN concurrently using additives containing two such elements, such as Bi3+ and Zr4+. Furthermore, it is crucial to remember that the R-T phase boundary can be produced by employing the aforementioned chemicals, resulting in a significant d33 of 400–490 pC/N [92,145,146]. In comparison to other phase boundaries (such as O-T and R-O), the R-T phase boundary is a more effective technique to boost the piezoelectric activity of KNN based ceramics. Furthermore, d33 and TC outperform BT and BNT-based materials across the board in KNN. Our discovery may lead the way for lead-free ceramics to be used in the real world, as the R-T phase barrier considerably enhances the piezoelectric activity of KNN-based ceramics [92,145,146,147]. Our research has led us to the idea that material systems exhibiting R-T phase boundaries are excellent potential candidates for potassium sodium niobate piezoceramics.

    It is well-established that KNN's piezoelectric capabilities can be considerably enhanced by establishing phase boundaries using various additives, that is, the development of phase boundaries driven by additives plays a significant role in their enhanced piezoelectric properties. KNN-based materials may have their electrical characteristics altered by other influences. Here, we will provide a high-level overview of the KNN-based piezoceramic engineering.

    Many studies are currently investigating how dopants change the TR-O, TO-T and TC of KNN-based piezoceramics. The effects of several dopants have been studied in depth. Research has looked into a variety of ion dopants, including Li+, Sb5+, Ta5+, Zr4+, Hf4+ and Ag+. The formation of an MPB requires the inclusion of these ions [73,83,84,168]. Increasing the amounts of Sb, Ta, Zr and Hf reduces the TO-T and lowers the TR-O. The operating range of a piezo material is reduced, and the TC is lowered by Ta, Sb and Hf. With its ability to boost TC while lowering TO-T, Li is a popular dopant for achieving and maintaining a high Curie temperature. Similar to Yang et al. (2020), who employed Ag+ to obtain high Curie temperature, the TC = 250 ℃ for a doped KNN based piezo material with high piezoelectric characteristics was attained [169]. Wong et al. (2016) created a high-performance KNN-based piezo material with TC = 483 ℃ and d33 > 200 pC/N using Li as a dopant [170]. The increased piezoelectric activity of KNN may potentially be attributable to the presence of other variables. For instance, KNN's A and B sites share valence with Li, Ta and Sb. Therefore, using them can prevent the formation of certain flaws. Safari et al. [171] claimed that the greater electronegativities of Ta5+ and Sb5+ can raise the covalence of the bonds with respect to Nb5+, leading to increased d33 values in (Li, Ta, Sb) co-doped KNN ceramics. The improved piezoelectric characteristics of KNN-based ceramics were also attributed to the high electronegativity of Sb5+, according to a theory proposed by Gao et al. [129]. Consequently, a number of parameters (electron configuration, ligand field theory, covalence, etc.) are taken into account while attempting to enhance the electrical properties by doping chemicals. Therefore, the introduction of one or more components into KNN might affect distinct phase boundaries; the fundamental benefit of these elements is that they cause a simultaneous change in TR-O, TO-T, and TC.

    In order to solve the densification issue that is present in KNN-based piezoceramics, enhance the sintering temperature, and make the material more temperature stable, oxides are utilised in addition to dopants that are of the ion or perovskite type. The inclusion of ZnO has the potential to improve the thermal stability of KNN [148]. MnO2 is yet another oxide that sees widespread application in KNN-based piezoceramics. Deng et al. [131] conducted research on the effect of adding MnO2 to KNLNT. They discovered that adding MnO2 to the material raised the density, lowered the temperature at which it sintered and, most crucially, improved its piezoelectric capabilities. They were able to produce d33 = 251 pC/N with doped KNLNT containing 0.1 mol% MnO2, which is a greater value than undoped KNLNT. Condurache et al. conducted research in 2019 [172] on the additional step of MnO2 and showed that the phase structure can be altered by adding MnO2 either before or after the calcination process. If MnO2 is supplied before calcination, the ratio of orthorhombic to tetragonal phase is 1.9. Conversely, if MnO2 is added after calcination, the ratio is 0.7. In 2019, Chen et al. [173] used FeO and MgO oxide types in addition to BaZrO3 dopant and achieved a piezo material that has a large piezoelectric coefficient and coupling factor, where d33 = 324 pC/N and KP = 59% respectively. This was accomplished by creating a piezo material that was doped with BaZrO3. It also has a high Curie temperature, which comes in at 304 ℃. In order to achieve a good piezoelectric response for KNN-based materials, Zhao et al. [174] utilised Fe2O3 and CuO. The results of their investigation demonstrate that CuO-doped KNN has a better electro-strain response than Fe2O3 doped KNN does.

    Meanwhile, recent research has found that rare earth (RE) elements have unusually large outer electron shells and permit numerous orbital bonding paths with other ions. As a result, local structural heterogeneity is created by rare earth dopants and the interfacial energy barriers between phases/polarization states are lowered [175,176,177]. It has been reported that rare-earth elements can stabilise and reduce the dissipation factor in dielectric ceramics, both of which are considered to be desirable properties. In a study that showed the existence of a ferroelectric-relaxor transition (Tf-r), it was discovered that adding moderate rare earth ions lowered the Tf-r. Changing from a relaxor to a ferroelectric due to an applied electric field. Tf-r decreased and approached room temperature with the coexistence of nonergodic and ergodic relaxor states, which was responsible for the improved strain property [163,164]. The piezoelectric constant, d33, of BNKT ceramics was raised after optimal Sm3+ was added increasing x from 0 to 0.015 increases the value from 150 to 190 pC/N and held steady at 165–180 pC/N between 0.015 and 0.030. One of the promising dopants is yttrium (Y3+), a kind of rare-earth element. This is due to the fact that it improves fatigue endurance, remanent polarisation and leakage current. According to Akmal's findings, the addition of yttrium reduced the electrical resistivity to 2.153 × 106 Ω, but at dopant concentrations greater than 0.5 mol%, the resistivity began to increase [178,179]. The addition of Y dopant increases the resistance magnitude in doped KNN films. Since the A-site of has an inherent positive charge, charge imbalance in a doped KNN lattice is compensated by cation vacancies [21,22,180]. Therefore, an increase in Y concentration results in an increase in the number of conduction electrons, leading to a dramatic drop in resistivity. It is often attributed to electrical compensation of the integrated cation [181,182]. Few investigations have found as Table 2 that improving the piezoelectric characteristics of rare-earth dopants by adding other rare-earth elements is achievable.

    Table 2.  Rare earth elements as dopants engineering.
    Rare earth dopants d33 Ref.
    CeO2 132 [182]
    Yb2O3 390 [183]
    Nd2O3 250 [184]
    Y2O3 - [185]

     | Show Table
    DownLoad: CSV

    The applications for piezoceramics are extensive. Generators, sensors, actuators, and transducers employ them for a wide variety of tasks, including energy harvesting, strain sensing, active vibration reduction, ultrasonic sonar, distance metre, and so on. When it comes to piezoceramics, PZT is a popular choice. Lead-free alternatives are needed to PZT based piezoceramics because of its high toxicity. This transition from PZT to lead-free piezoceramics is being ensured by the legislation of several countries.

    KNN is a potential lead-free material for piezoelectric ceramics, but its poor characteristics and temperature dependence have slowed its development for nearly 50 years since its discovery in the 1950s [183]. Recent research has shown that phase boundary engineering (PBE) can improve KNN systems' electrical properties, making them more appropriate for industrial applications. To compensate for these weaknesses, lead-free piezoceramics like KNN can be doped to improve their characteristics. PBE improves material characteristics by changing phase boundaries. PBE enhances KNN's piezoelectric characteristics through synergistic contributions [184]. Phase engineering has been widely used to increase piezoelectricity in KNN-based ceramics by comparing the effects of standard additives on phase structure [185]. Previous section has discussed the results indicate that PBE has the potential to be a useful technique for enhancing the electrical properties of KNN systems, making them more applicable to industrial settings. Therefore, KNN-based piezoceramics can be improved to replace the role of lead-based ones. In addition to its other desirable properties, piezoceramics based on KNN have a high Curie temperature. However, KNN based piezoceramics have a number of limitations, including a high sintering temperature and a poor piezoelectric coefficient. Sintering aids such as ZnO, MnO2 and CuO have been studied. By this means, piezoceramic densification during sintering is enhanced. Reducing TO-T and raising TO-R to establish MPB is the key strategy for enhancing the piezoelectric coefficient for KNN based piezoceramics. At the same time, it is important to keep TC levels high. Other dopants are either TC decreasing or inactive. However, Li and Ag addition can be utilised to boost TC. The piezoelectric coefficient and Curie temperature of KNN-based piezoceramics have been reported to be rather high, for example d33 = 451 pC/N and TC = 258 ℃ [15], d33 = 440 pC/N and TC = 250 ℃ [16] and d33 = 308 pC/N and TC = 328 ℃ [25]. The piezoelectric ceramic substance KNN has recently attracted interest as a possible replacement for lead-based piezoceramics in a variety of electrical applications due to their high Curie temperature and piezoelectric coefficient. For instance, KNN has found widespread application as a dielectric material in various electronic applications. As a result of its high dielectric constant and low dielectric loss, it can be used in capacitors and other energy storage devices. By virtue of its dielectric properties, KNN can improve the functionality and efficiency of electrical equipment. However, KNN's appealing ferroelectric characteristics also find use in sensor and actuator applications [186]. Its piezoelectric properties enable it to convert mechanical energy into electrical signals and vice versa. This makes KNN appropriate for biomedical, automotive and industrial sensors, actuators and transducers. While the phase space and associated properties of KNN in ceramic form have been the subject of extensive research, a similarly robust body of literature for thin film KNN is still in its infancy. Micro-electro-mechanical systems (MEMS) have shown to be an excellent method for developing miniaturised systems and devices that combine mechanical and electrical functions. This innovation has prompted the development of KNN based on thin films. These thin films have high transparency and low absorption in the visible and near-infrared spectrums, in addition to exhibiting ferroelectric behaviour. Moreover, research has switched towards developing methods for producing high-quality films and, more recently, towards applying the material in micro- and nano-electronic devices and biomedical settings. KNN films with enhanced characteristics have been deposited and these have potential applications in micro- and nano electronic devices, as demonstrated by a number of studies that have revealed the film synthesis methodologies, highlighting key results and the challenges connected with synthesis approach [187,188]. As interest in developing lead-free piezoceramics grows, so too does the need for more study into KNN-based piezoceramics. One area where more work needs to be done is in the production of KNN-based piezoceramics. KNN based piezoceramics have been shown to have excellent piezoelectric capabilities and a high curie temperature; further investigation into their mechanical properties, electrical properties and mechanical and electrical fatigue behaviours is recommended. Improvements in these fields of study would aid in making informed choices about the range of possible KNN-based piezoceramics and their respective applications.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    Authors are grateful to the Ministry of Higher Education, Malaysia and Universiti Teknikal Malaysia Melaka (UTeM) under the Fundamental Research Grant Scheme (FRGS), grant No.: FRGS/1/2021/TK0/UTEM/01/2.

    The authors declare no conflict of interest.



    [1] Ringgaard E, Wurlitzer T, Wolny WW (2011) Properties of lead-free piezoceramics based on alkali niobates. Ferroelectrics 319: 97–107. http://doi.org/10.1080/00150190590965497 doi: 10.1080/00150190590965497
    [2] Cross E (2004) Lead-free at last. Nature 432: 24–25. https://doi.org/10.1038/nature03142 doi: 10.1038/nature03142
    [3] Jaffe B, Cook WR, Jaffe H (1971) Piezoelectric Ceramics, Non-Metallic Solids, London: Academic Press. https://doi.org/10.1016/0022-460X(72)90684-0
    [4] Kumar P, Pattanaik M, Sonia (2013) Synthesis and characterizations of KNN ferroelectric ceramics near 50/50 MPB. Ceram Int 39: 65–69. http://doi.org/10.1016/j.ceramint.2012.05.093 doi: 10.1016/j.ceramint.2012.05.093
    [5] Takenaka T, Nagata H (2005) Current status and prospects of lead-free piezoelectric ceramics. J Eur Ceram Soc 25: 2693–2700. http://doi.org/10.1016/j.jeurceramsoc.2005.03.125 doi: 10.1016/j.jeurceramsoc.2005.03.125
    [6] Shrout TR, Zhang SJ (2007) Lead-free piezoelectric ceramics: Alternatives for PZT? J Electroceramics 19: 111–124. http://doi.org/10.1007/s10832-007-9047-0 doi: 10.1007/s10832-007-9047-0
    [7] Rödel J, Jo W, Seifert KTP, et al. (2009) Perspective on the development of lead-free piezoceramics. J Am Ceram Soc 92: 1153–1177. http://doi.org/10.1111/j.1551-2916.2009.03061.x doi: 10.1111/j.1551-2916.2009.03061.x
    [8] Chen K, Zhang F, Li D, et al. (2016) Acceptor doping effects in (K0.5Na0.5)NbO3 lead-free piezoelectric ceramics. Ceram Int 42: 2899–2903. http://doi.org/10.1016/j.ceramint.2015.11.016 doi: 10.1016/j.ceramint.2015.11.016
    [9] Jo W, Dittmer R, Acosta M, et al. (2012) Giant electric-field-induced strains in lead-free ceramics for actuator applications—Status and perspective. J Electroceram 29: 71–93. http://doi.org/10.1007/s10832-012-9742-3 doi: 10.1007/s10832-012-9742-3
    [10] Saito Y, Takao H, Tani T, et al. (2004) Lead-free piezoceramics. Nature 432: 84–87. http://doi.org/10.1038/nature03028 doi: 10.1038/nature03028
    [11] Kim JS, Ahn CW, Lee SY, et al. (2011) Effects of LiNbO3 substitution on lead-free (K0.5Na0.5)NbO3 ceramics: Enhanced ferroelectric and electrical properties. Curr Appl Phys 11: S149–S153. http://doi.org/10.1016/j.cap.2011.03.049 doi: 10.1016/j.cap.2011.03.049
    [12] Liang W, Wu W, Xiao D, et al. (2011) Effect of the addition of CaZrO3 and LiNbO3 on the phase transitions and piezoelectric properties of K0.5Na0.5NbO3 lead-free ceramics. J Am Ceram Soc 94: 4317–4322. http://doi.org/10.1111/j.1551-2916.2011.04660.x doi: 10.1111/j.1551-2916.2011.04660.x
    [13] Wu J, Xiao D, Zhu J (2015) Potassium-sodium niobate lead-free piezoelectric materials: Past, present, and future of phase boundaries. Chem Rev 115: 2559–2595. http://doi.org/10.1021/cr5006809 doi: 10.1021/cr5006809
    [14] Akmal MMT, Warikh ARM, Azlan UA (2015) Physical and electrical properties enhancement of rare-earth doped-potassium sodium niobate (KNN): A review. Ceram-Silikatty 59: 158–163.
    [15] Zuo R, Rödel J, Chen R, et al. (2006) Sintering and electrical properties of lead-free Na0.5K0.5NbO3 piezoelectric ceramics. J Am Ceram Soc 89: 2010–2015. http://doi.org/10.1111/j.1551-2916.2006.00991.x doi: 10.1111/j.1551-2916.2006.00991.x
    [16] Zuo R, Lv D, Fu J, et al. (2009) Phase transition and electrical properties of lead free (Na0.5K0.5)NbO3-BiAlO3 ceramics. J Alloys Compd 476: 836–839. http://doi.org/10.1016/j.jallcom.2008.09.123 doi: 10.1016/j.jallcom.2008.09.123
    [17] Benabdallah F, Simon A, Khemakhem H, et al. (2011) Linking large piezoelectric coefficients to highly flexible polarization of lead free BaTiO3-CaTiO3-BaZrO3 ceramics. J Appl Phys 109: 124116. http://doi.org/1063/1.3599854 doi: 10.1063/1.3599854
    [18] Paul J, Nishimatsu T, Kawazoe Y, et al. (2007) Ferroelectric phase transitions in ultrathin films of BaTiO3. Phys Rev Lett 99: 077601. https://doi.org/10.1103/PhysRevLett.99.077601 doi: 10.1103/PhysRevLett.99.077601
    [19] Herber RP, Schneider GA, Wagner S, et al. (2007) Characterization of ferroelectric domains in morphotropic potassium sodium niobate with scanning probe microscopy. Appl Phys Lett 90: 2522905. http://doi.org/10.1063/1.2750395 doi: 10.1063/1.2750395
    [20] Wurfel P, Batra IP, Jacobs JT (1973) Polarization instability in thin ferroelectric films. Phys Rev Lett 30: 1218–1221. http://doi.org/10.1103/PhysRevLett.30.1218 doi: 10.1103/PhysRevLett.30.1218
    [21] Song HC, Kim HC, Kang CY, et al. (2009) Multilayer piezoelectric energy scavenger for large current generation. J Electroceramics 23: 301–304. https://doi.org/10.1007/s10832-008-9439-9 doi: 10.1007/s10832-008-9439-9
    [22] Abazari M, Akdoǧan EK, Safari A (2008) Effect of manganese doping on remnant polarization and leakage current in (K0.44, Na0.52, Li0.04) (Nb0.84, Ta0.10, Sb0.06) O3 epitaxial thin films on SrTiO3. Appl Phys Lett 92: 212903. http://doi.org/10.1063/1.2937000 doi: 10.1063/1.2937000
    [23] Yoo J, Lee K, Chung K, et al. (2006) Piezoelectric and dielectric properties of (LiNaK)(NbTaSb)O3 ceramics with variation in poling temperature. Jpn J Appl Phys 45: 7444–7448. https://dx.doi.org/10.1143/JJAP.45.7444 doi: 10.1143/JJAP.45.7444
    [24] Wang D, Cao M, Zhang S (2012) Investigation of ternary system PbHfO3-PbTiO3-Pb (Mg1/3 Nb2/3)O3 with morphotropic phase boundary compositions. J Am Ceram Soc 95: 3220–3228. https://doi.org/10.1111/j.1551-2916.2012.05300.x doi: 10.1111/j.1551-2916.2012.05300.x
    [25] Ahn CW, Jeong ED, Lee SY, et al. (2010) Enhanced ferroelectric properties of LiNbO3 substituted Na0.5K0.5NbO3 lead-free thin films grown by chemical solution deposition. Appl Phys Lett 93: 212905. http://doi.org/10.1063/1.3037214 doi: 10.1063/1.3037214
    [26] Hagh NM, Jadidian B, Ashbahian E, et al. (2008) Lead-free piezoelectric ceramic transducer in the donor-doped K1/2Na1/2NbO3 solid solution system. IEEE T Ultrason Ferr 55: 214–224. http://doi.org/10.1109/TUFFC.2008.630 doi: 10.1109/TUFFC.2008.630
    [27] Courths R, Steiner P, Höchst H, et al. (1980) Photoelectron-spectroscopy investigation and electronic properties of LiNbO3 crystal surfaces. Appl Phys 21: 345–352. https://doi.org/10.1007/BF00895926 doi: 10.1007/BF00895926
    [28] Lin D, Xiao D, Zhu J, et al. (2006) Piezoelectric and ferroelectric properties of lead-free [Bi1-y (Na1-xyLix)]0.5BayTiO3 ceramics. J Eur Ceram Soc 26: 3247–3251. https://doi.org/10.1016/j.jeurceramsoc.2005.09.038 doi: 10.1016/j.jeurceramsoc.2005.09.038
    [29] Reaney IM, Damjanovic D (1996) Crystal structure and domain-wall contributions to the piezoelectric properties of strontium bismuth titanate ceramics. J Appl Phys 80: 4223–4225. https://doi.org/10.1063/1.363301 doi: 10.1063/1.363301
    [30] Wu J, Tao H, Yuan Y, et al. (2015) Role of antimony in the phase structure and electrical properties of potassium-sodium niobate lead-free ceramics. RSC Adv 5: 14575–14583. https://doi.org/10.1039/C4RA14271C doi: 10.1039/C4RA14271C
    [31] Bernard J, Benčan A, Rojac T, et al. (2008) Low-temperature sintering of K0.5Na0.5NbO3 ceramics. J Am Ceram Soc 91: 2409–2411. https://doi.org/10.1111/j.1551-2916.2008.02447.x doi: 10.1111/j.1551-2916.2008.02447.x
    [32] Rubio-Marcos F, Romero JJ, Martín-Gonzalez MS, et al. (2010) Effect of stoichiometry and milling processes in the synthesis and the piezoelectric properties of modified KNN nanoparticles by solid state reaction. J Eur Ceram Soc 30: 2763–2771. https://doi.org/10.1016/j.jeurceramsoc.2010.05.027 doi: 10.1016/j.jeurceramsoc.2010.05.027
    [33] López-Juárez R, Novelo-Peralta O, González-García F, et al. (2011) Ferroelectric domain structure of lead-free potassium-sodium niobate ceramics. J Eur Ceram Soc 31: 1861–1864. https://doi.org/10.1016/j.jeurceramsoc.2011.02.031 doi: 10.1016/j.jeurceramsoc.2011.02.031
    [34] Shen ZY, Li JF, Wang K, et al. (2010) Electrical and mechanical properties of fine-grained Li/Ta-modified (Na, K)NbO3-based piezoceramics prepared by spark plasma sintering. J Am Ceram Soc 93: 1378–1383. https://doi.org/10.1111/j.1551-2916.2009.03542.x doi: 10.1111/j.1551-2916.2009.03542.x
    [35] Uršič H, Benčan A, Škarabot M, et al. (2010) Dielectric, ferroelectric, piezoelectric, and electrostrictive properties of K0.5Na0.5NbO3 single crystals. J Appl Phys 107: 033705. https://doi.org/10.1063/1.3291119 doi: 10.1063/1.3291119
    [36] Bobnar V, Bernard J, Kosec M. (2004) Relaxorlike dielectric properties and history-dependent effects in the lead-free K0.5Na0.5NbO3-SrTiO3 ceramic system. Appl Phys Lett 85: 994–996. http://doi.org/10.1063/1.1779947 doi: 10.1063/1.1779947
    [37] Wolny WW (2004) European approach to development of new environmentally sustainable electroceramics. Ceram Int 30: 1079–1083. https://doi.org/10.1016/j.ceramint.2003.12.025 doi: 10.1016/j.ceramint.2003.12.025
    [38] Uesu Y, Matsuda M, Yamada Y, et al. (2002) Symmetry of high-piezoelectric Pb-based complex perovskites at the morphotropic phase boundary: I. Neutron diffraction study on Pb(Zn1/3Nb2/3)O3-9%PbTiO3. J Phys Soc Jpn 71: 960–965. https://doi.org/10.1143/JPSJ.71.960 doi: 10.1143/JPSJ.71.960
    [39] Zuo R, Xu Z, Li L (2008) Dielectric and piezoelectric properties of Fe2O3-doped (Na0.5K0.5)0.96Li0.04Nb0.86Ta0.1Sb0.04O3 lead-free ceramics. J Phys Chem Solids 69: 1728–1732. http://doi.org/10.1016/j.jpcs.2008.01.003 doi: 10.1016/j.jpcs.2008.01.003
    [40] Park SE, Shrout TR (1997) Ultrahigh strain and piezoelectric behavior in relaxor based ferroelectric single crystals. J Appl Phys 82: 1804–1811. http://doi.org/10.1063/1.365983 doi: 10.1063/1.365983
    [41] Damjanovic D (1998) Ferroelectric, dielectric and piezoelectric properties of ferroelectric thin films and ceramics. Rep Prog Phys 61: 1267–1324. http://doi.org/10.1088/0034-4885/61/9/002 doi: 10.1088/0034-4885/61/9/002
    [42] Kittel C (1946) Theory of the structure of ferromagnetic domains in films and small particles. Phys Rev 70: 965–971. http://doi.org/10.1103/PhysRev.70.965 doi: 10.1103/PhysRev.70.965
    [43] Egerton L, Dillon DM (1959) Piezoelectric and dielectric properties of ceramics in the system potassium-sodium niobate. J Am Ceram Soc 42: 438–442. http://doi.org/10.1111/j.1151-2916.1959.tb12971.x doi: 10.1111/j.1151-2916.1959.tb12971.x
    [44] Noheda B, Cox DE, Shirane G, et al. (1999) A monoclinic ferroelectric phase in the Pb(Zr1-xTix)O3 solid solution. Appl Phys Lett 74: 2059–2061. http://doi.org/10.1063/1.123756 doi: 10.1063/1.123756
    [45] Guo R, Cross LE, Park SE, et al. (2000) Origin of the high piezoelectric response in PbZr1-xTixO3. Phys Rev Lett 84: 5423–5426. http://doi.org/10.1103/PhysRevLett.84.5423 doi: 10.1103/PhysRevLett.84.5423
    [46] Kreisel J, Glazer AM, Jones G, et al. (2000) An X-ray diffraction and Raman spectroscopy investigation of A-site substituted perovskite compounds: The (Na1-xKx)0.5Bi0.5TiO3 (0 ≤ x ≤ 1) solid solution. J Phys Condens Matter 12: 3267–3280. http://doi.org/10.1088/0953-8984/12/14/305 doi: 10.1088/0953-8984/12/14/305
    [47] Singh AK, Pandey D (2003) Evidence for MB and MC phases in the morphotropic phase boundary region of (1-x)[Pb(Mg1/3Nb2/3)O3]-xPbTiO3: A rietveld study. Phys Rev B 67: 064102. http://doi.org/10.1103/PhysRevB.67.064102 doi: 10.1103/PhysRevB.67.064102
    [48] Wang P, Li Y, Lu Y (2011) Enhanced piezoelectric properties of (Ba0.85Ca0.15)(Ti0.9Zr0.1)O3 lead-free ceramics by optimizing calcination and sintering temperature. J Eur Ceram Soc 31: 2005–2012. http://doi.org/10.1016/j.jeurceramsoc.2011.04.023 doi: 10.1016/j.jeurceramsoc.2011.04.023
    [49] Bilc DI, Orlando R, Shaltaf R, et al. (2008) Hybrid exchange-correlation functional for accurate prediction of the electronic and structural properties of ferroelectric oxides. Phys Rev B 77: 165107. http://doi.org/10.1103/PhysRevB.77.165107 doi: 10.1103/PhysRevB.77.165107
    [50] Karaki T, Yan K, Miyamoto T, et al. (2007) Lead-free piezoelectric ceramics with large dielectric and piezoelectric constants manufactured from BaTiO3 nano-powder. Jpn J Appl Phys 46: L97. http://doi.org/10.1143/JJAP.46.L97 doi: 10.1143/JJAP.46.L97
    [51] Kakimoto KI, Akao K, Guo Y, et al. (2005) Raman scattering study of piezoelectric (N0.5K0.5)NbO3-LiNbO3 ceramics. Jpn J Appl Phys 44: 7064–7067. http://doi.org/10.1143/JJAP.44.7064 doi: 10.1143/JJAP.44.7064
    [52] Keeble DS, Benabdallah F, Thomas PA, et al. (2013) Revised structural phase diagram of (Ba0.7C0.3TiO3)(BaZr0.2Ti0.8O3). Appl Phys Lett 102: 092903. http://doi.org/10.1063/1.4793400 doi: 10.1063/1.4793400
    [53] Zhen Y, Li JF (2006) Normal sintering of (K, Na)NbO3-based ceramics: Influence of sintering temperature on densification, microstructure, and electrical properties. J Am Ceram Soc 89: 3669–3675. http://doi.org/10.1111/j.1551-2916.2006.01313.x doi: 10.1111/j.1551-2916.2006.01313.x
    [54] Buhret CF (1962) Some properties of bismuth perovskites. J Chem Phys 36: 798–803. http://doi.org/10.1063/1.1732613 doi: 10.1063/1.1732613
    [55] Schönau KA, Schmitt LA, Knapp M, et al. (2007) Nanodomain structure of Pb[Zr1-xTix]O3 at its morphotropic phase boundary: Investigations from local to average structure. Phys Rev B 75: 184117. http://doi.org/10.1103/PhysRevB.75.184117 doi: 10.1103/PhysRevB.75.184117
    [56] Wylie VEB, Damjanovic D, Klein N, et al. (2010) Structural complexity of (Na0.5Bi0.5)TiO3-BaTiO3 as revealed by Raman spectroscopy. Phys Rev B 82: 104112. http://doi.org/10.1103/PhysRevB.82.104112 doi: 10.1103/PhysRevB.82.104112
    [57] Woodward DI, Knudsen J, Reaney IM (2005) Review of crystal and domain structures in the PbZrxTi1-xO3 solid solution. Phys Rev B 72: 104110. http://doi.org/10.1103/PhysRevB.72.104110 doi: 10.1103/PhysRevB.72.104110
    [58] Ye ZG, Dong M (2000) Morphotropic domain structures and phase transitions in relaxor-based piezo-/ferroelectric (1-x)Pb(Mg1/3Nb2/3)O3-xPbTiO3 single crystals. J Appl Phys 87: 2312–2319. http://doi.org/10.1063/1.372180 doi: 10.1063/1.372180
    [59] Liu SF, Park SE, Shrout TR, et al. (1999) Electric field dependence of piezoelectric properties for rhombohedral 0.955Pb(Zn1/3Nb2/3)O3-0.045PbTiO3 single crystals. J Appl Phys 85: 2810–2814. http://doi.org/10.1063/1.369599 doi: 10.1063/1.369599
    [60] Anton EM, Jo W, Damjanovic D, et al. (2011) Determination of depolarization temperature of (Bi1/2Na1/2)TiO3-based lead-free piezoceramics. J Appl Phys 110: 094108. http://doi.org/10.1063/1.3660253 doi: 10.1063/1.3660253
    [61] Noblanc O, Gaucher P, Calvarin G (1996) Structural and dielectric studies of Pb(Mg1/3Nb2/3)O3-PbTiO3 ferroelectric solid solutions around the morphotropic boundary. J Appl Phys 79: 4291–4297. http://doi.org/10.1063/1.361865 doi: 10.1063/1.361865
    [62] Xue D, Zhou Y, Bao H, et al. (2011) Elastic, piezoelectric, and dielectric properties of Ba(Zr0.2Ti0.8)O3-50(Ba0.7Ca0.3)TiO3 Pb-free ceramic at the morphotropic phase boundary. J Appl Phys 109: 054110. http://doi.org/10.1063/1.3549173 doi: 10.1063/1.3549173
    [63] Wu J, Wang Y, Xiao D, et al. (2007) Effects of Ag content on the phase structure and piezoelectric properties of (K0.44-xNa0.52Li0.04 Agx)(Nb0.91Ta0.05Sb0.04)O3 lead-free ceramics. Appl Phys Lett 91: 132914. http://doi.org/10.1063/1.2793507 doi: 10.1063/1.2793507
    [64] Matsubara M, Yamaguchi T, Kikuta K, et al. (2004) Sinterability and piezoelectric properties of (K, Na)NbO3 ceramics with novel sintering aid. Jpn J Appl Phys 43: 7159–7163. http://doi.org/10.1143/JJAP.43.7159 doi: 10.1143/JJAP.43.7159
    [65] Kounga AB, Zhang ST, Jo W, et al. (2008) Morphotropic phase boundary in (1-x)Bi0.5Na0.5TiO3-xK0.5Na0.5NbO3 lead-free piezoceramics. Appl Phys Lett 92: 222902. http://doi.org/10.1063/1.2938064 doi: 10.1063/1.2938064
    [66] Wang K, Li JF (2007) Analysis of crystallographic evolution in (Na, K)NbO3-based lead-free piezoceramics by X-ray diffraction. Appl Phys Lett 91: 262902. http://doi.org/10.1063/1.2825280 doi: 10.1063/1.2825280
    [67] Akdoǧan EK, Kerman K, Abazari M, et al. (2008) Origin of high piezoelectric activity in ferroelectric (K0.44Na0.52Li0.04)-(Nb0.84Ta0.1Sb0.06)O3 ceramics. Appl Phys Lett 92: 112908. http://doi.org/10.1063/1.2897033 doi: 10.1063/1.2897033
    [68] Elkechai O, Manier M, Mercurio JP (1996) Na0.5Bi0.5TiO3-K0.5Bi0.5TiO3 (NBT-KBT) system: A structural and electrical study. Phys Stat Sol 157: 499–506. http://doi.org/10.1002/pssa.2211570234 doi: 10.1002/pssa.2211570234
    [69] Li F, Zhang S, Xu Z, et al. (2010) Composition and phase dependence of the intrinsic and extrinsic piezoelectric activity of domain engineered (1-x)Pb(Mg1/3Nb2/3)O3-xPbTiO3 crystals. J Appl Phys 108: 034106. http://doi.org/10.1063/1.3466978 doi: 10.1063/1.3466978
    [70] Zuo R, Fu J (2011) Rhombohedral-tetragonal phase coexistence and piezoelectric properties of (NaK)(NbSb)O3-LiTaO3-BaZrO3 lead-free ceramics. J Am Ceram Soc 94: 1467–1470. http://doi.org/10.1111/j.1551-2916.2010.04256.x doi: 10.1111/j.1551-2916.2010.04256.x
    [71] Fisher JG, Benčan A, Holc J, et al. (2007) Growth of potassium sodium niobate single crystals by solid state crystal growth. J Cryst Growth 303: 487–492. http://doi.org/10.1016/j.jcrysgro.2007.01.011 doi: 10.1016/j.jcrysgro.2007.01.011
    [72] Jo W, Seifert KTP, Anton E, et al. (2009) Perspective on the development of lead-free piezoceramics. J Am Ceram Soc 92: 1153–1177. http://doi.org/10.1111/j.1551-2916.2009.03061.x doi: 10.1111/j.1551-2916.2009.03061.x
    [73] Zhang S, Xia R, Shrout TR (2007) Lead-free piezoelectric ceramics vs. PZT? J Electroceram 19: 251–257. http://doi.org/10.1007/s10832-007-9056-z doi: 10.1007/s10832-007-9056-z
    [74] Zuo R, Fu J, Lv D, et al. (2010) Antimony tuned rhombohedral-orthorhombic phase transition and enhanced piezoelectric properties in sodium potassium niobate. J Am Ceram Soc 93: 2783–2787. http://doi.org/10.1111/j.1551-2916.2010.03804.x doi: 10.1111/j.1551-2916.2010.03804.x
    [75] Wu J, Xiao D, Wang Y, et al. (2008) CaTiO3-modified[(K0.5Na0.5)0.94Li0.06](Nb0.94Sb0.06)O3 lead-free piezoelectric ceramics with improved temperature stability. Scripta Mater 59: 750–752. http://doi.org/10.1016/j.scriptamat.2008.06.011 doi: 10.1016/j.scriptamat.2008.06.011
    [76] Wu J, Xiao D, Wang Y, et al. (2008) Improved temperature stability of CaTiO3-modified [(K0.5Na0.5)0.96Li0.04](Nb0.91Sb0.05Ta0.04)O3 lead-free piezoelectric ceramics. J Appl Phys 104: 024102. http://doi.org/10.1063/1.2956390 doi: 10.1063/1.2956390
    [77] Leontsev SO, Eitel RE (2010) Progress in engineering high strain lead-free piezoelectric ceramics. Sci Technol Adv Mat 11: 044302. http://doi.org/10.1088/1468-6996/11/4/044302 doi: 10.1088/1468-6996/11/4/044302
    [78] Kosec M, Malič B, Benčan A, et al. (2008) KNN-based piezoelectric ceramics, In: Safari A, Akdoğan EK, Piezoelectric and Acoustic Materials for Transducer Applications, Boston: Springer. https://doi.org/10.1007/978-0-387-76540-2_5
    [79] Liu W, Ren X (2009) Large piezoelectric effect in Pb-free ceramics. Phys Rev Lett 103: 257602. http://doi.org/10.1103/PhysRevLett.103.257602 doi: 10.1103/PhysRevLett.103.257602
    [80] Jo W, Schaab S, Sapper E, et al. (2011) On the phase identity and its thermal evolution of lead free (Bi1/2Na1/2)TiO3-6 mol BaTiO3. J Appl Phys 110: 074106. http://doi.org/10.1063/1.3645054 doi: 10.1063/1.3645054
    [81] Wang X, Wu J, Xiao D, et al. (2014) Giant piezoelectricity in potassium-sodium niobate lead-free ceramics. J Am Chem Soc 136: 2905–2910. http://doi.org/10.1021/ja500076h doi: 10.1021/ja500076h
    [82] Cheng X, Wu J, Lou X, et al. (2014) Achieving both giant d33 and high TC in potassium-sodium niobate ternary system. ACS Appl Mater Interfaces 6: 750–756. https://doi.org/10.1021/am404793e doi: 10.1021/am404793e
    [83] Cheng X, Wu J, Wang X, et al. (2013) Mediating the contradiction of d33 and TC in potassium-sodium niobate lead-free piezoceramics. ACS Appl Mater Interfaces 5: 10409–10417. http://doi.org/10.1021/am403448r doi: 10.1021/am403448r
    [84] Wang X, Wu J, Xiao D, et al. (2014) New potassium-sodium niobate ceramics with a giant d33. ACS Appl Mater Interfaces 6: 6177–6180. http://doi.org/10.1021/am500819v doi: 10.1021/am500819v
    [85] Cheng X, Wu J, Wang X, et al. (2014) New lead-free piezoelectric ceramics based on (K0.48Na0.52)(Nb0.95Ta0.05)O3-Bi0.5(Na0.7K0.2Li0.1)0.5ZrO3. Dalton Trans 43: 3434–3442. http://doi.org/10.1039/c3dt52603h doi: 10.1039/c3dt52603h
    [86] Liang W, Wang Z, Xiao D, et al. (2012) Effect of new phase boundary on the dielectric and piezoelectric properties of K0.5Na0.5NbO3-xBaZrO3-yBi0.5Na0.5TiO3 lead-free ceramics. Integr Ferroelectr 139: 63–74. http://doi.org/10.1080/10584587.2012.737221 doi: 10.1080/10584587.2012.737221
    [87] Zhang B, Wu J, Cheng X, et al. (2013) Lead-free piezoelectrics based on potassium-sodium niobate with giant d33. ACS Appl Mater Interfaces 5: 7718–7725. http://doi.org/10.1021/am402548x doi: 10.1021/am402548x
    [88] Liang W, Wu W, Xiao D, et al. (2011) Construction of new morphotropic phase boundary in 0.94(K0.42xNa0.6BaxNb1-xZrx)O3-0.06LiSbO3 lead-free piezoelectric ceramics. J Mater Sci 46: 6871–6876. http://doi.org/10.1007/s10853-011-5650-1 doi: 10.1007/s10853-011-5650-1
    [89] Munz D, Fett T (2000) Ceramics: Mechanical Properties, Failure Behaviour, Materials Selection, Heidelberg: Springer Berlin.
    [90] Han C, Wu J, Pu C, et al. (2012) High piezoelectric coefficient of Pr2O3-doped Ba0.85Ca0.15Ti0.90Zr0.10O3 ceramics. Ceram Int 38: 6359–6363. http://doi.org/10.1016/j.ceramint.2012.05.008 doi: 10.1016/j.ceramint.2012.05.008
    [91] Kalyani AK, Senyshyn A, Ranjan R (2013) Polymorphic phase boundaries and enhanced piezoelectric response in extended composition range in the lead free ferroelectric BaTi1-xZrxO3. J Appl Phys 114: 014102. http://doi.org/10.1063/1.4812472 doi: 10.1063/1.4812472
    [92] Miclea C, Tanasoiu C, Miclea CF, et al. (2005) Effect of iron and nickel substitution on the piezoelectric properties of PZT type ceramics. J Eur Ceram Soc 25: 2397–2400. http://doi.org/10.1016/j.jeurceramsoc.2005.03.069 doi: 10.1016/j.jeurceramsoc.2005.03.069
    [93] Maurya D, Pramanick A, An K, et al. (2012) Enhanced piezoelectricity and nature of electric-field induced structural phase transformation in textured lead-free piezoelectric Na0.5Bi0.5TiO3-BaTiO3 ceramics. Appl Phys Lett 100: 172906. http://doi.org/10.1063/1.4709404 doi: 10.1063/1.4709404
    [94] Singh AK, Mishra SK, Ragini, et al. (2008) Origin of high piezoelectric response of Pb(ZrxTi1-x)O3 at the morphotropic phase boundary: Role of elastic instability. Appl Phys Lett 92: 022910. http://doi.org/10.1063/1.2836269 doi: 10.1063/1.2836269
    [95] Jo W, Daniels JE, Jones JL, et al. (2011) Evolving morphotropic phase boundary in lead-free (Bi1/2Na1/2)TiO3-BaTiO3 piezoceramics. J Appl Phys 109: 014110. http://doi.org/10.1063/1.3530737 doi: 10.1063/1.3530737
    [96] Zheng T, Wu J, Cheng X, et al. (2014) High strain in (K0.40Na0.60)(Nb0.955Sb0.045)O3-Bi0.50Na0.50ZrO3 lead-free ceramics with large piezoelectricity. J Mater Chem C 2: 8796–8803. http://doi.org/10.1039/c4tc01533a doi: 10.1039/c4tc01533a
    [97] Zhu X, Xu J, Meng Z (1997) Dielectric and piezoelectric properties of Pb(Ni1/3Nb2/3)O3-PbTiO3-PbZrO3 ceramics modified with bismuth and zinc substitutions. J Mater Sci 32: 4275–4282. http://doi.org/10.1023/A:1018655419424 doi: 10.1023/A:1018655419424
    [98] Gan BK, Yao K, He X (2007) Complex oxide ferroelectric ceramics Pb(Ni1/3Nb2/3)O3-Pb(Zn1/3Nb2/3)O3-Pb(Mg1/3Nb2/3)O3-PbZrO3-PbTiO3 with a low sintering temperature. J Am Ceram Soc 90: 1186–1192. http://doi.org/10.1111/j.1551-2916.2007.01617.x doi: 10.1111/j.1551-2916.2007.01617.x
    [99] Liu Y, Li Q, Qiao L, et al. (2022) Achieving giant piezoelectricity and high property uniformity simultaneously in a relaxor ferroelectric crystal through rare-earth element doping. Adv Sci 9: 2204631. http://doi.org/10.1002/advs.202204631 doi: 10.1002/advs.202204631
    [100] Tellier J, Malic B, Dkhil B, et al. (2009) Crystal structure and phase transitions of sodium potassium niobate perovskites. Solid State Sci 11: 320–324. https://doi.org/10.1016/j.solidstatesciences.2008.07.011 doi: 10.1016/j.solidstatesciences.2008.07.011
    [101] Baker DW, Thomas PA, Zhang N, et al. (2009) Structural study of KxNa1-xNbO3 (KNN) for compositions in the range x = 0.24–0.36. Acta Cryst B 65: 22–28. http://doi.org/10.1107/S0108768108037361 doi: 10.1107/S0108768108037361
    [102] Sharma JP, Kumar D, Sharma AK (2021) Structural and dielectric properties of pure potassium sodium niobate (KNN) lead free ceramics. Solid State Commun 334–335. http://doi.org/10.1016/j.ssc.2021.114345 doi: 10.1016/j.ssc.2021.114345
    [103] Zheng T, Wu J, Xiao D, et al. (2015) Potassium-sodium niobate lead-free ceramics: Modified strain as well as piezoelectricity. J Mater Chem A 3: 1868–1874. http://doi.org/10.1039/c4ta05423g doi: 10.1039/c4ta05423g
    [104] Lv X, Wu J, Xiao D, et al. (2018) Structural evolution of the R-T phase boundary in KNN-based ceramics. J Am Ceram Soc 101: 1191–1200. http://doi.org/10.1111/jace.15266 doi: 10.1111/jace.15266
    [105] Wang R, Bando H, Katsumata T, et al. (2009) Tuning the orthorhombic-rhombohedral phase transition temperature in sodium potassium niobate by incorporating barium zirconate. Phys Status Solidi-R 3: 142–144. http://doi.org/10.1002/pssr.200903090 doi: 10.1002/pssr.200903090
    [106] Li JF, Wang K, Zhu FY, et al. (2013) (K, Na)NbO3-based lead-free piezoceramics: Fundamental aspects, processing technologies, and remaining challenges. J Am Ceram Soc 96: 3677–3696. http://doi.org/10.1111/jace.12715 doi: 10.1111/jace.12715
    [107] Tennery VJ, Hang KW (1968) Thermal and X-ray diffraction studies of the NaNbO3 single bond sign KNbO3 system. J Appl Phys 39: 4749–4753. http://doi.org/10.1063/1.1655833 doi: 10.1063/1.1655833
    [108] Dai YJ, Zhang XW, Chen KP (2009) Morphotropic phase boundary and electrical properties of K1-xNaxNbO3 lead-free ceramics. Appl Phys Lett 94: 042905. http://doi.org/10.1063/1.3076105 doi: 10.1063/1.3076105
    [109] Karaki T, Katayama T, Yoshida K, et al. (2013) Morphotropic phase boundary slope of (K, Na, Li)NbO3-BaZrO3 binary system adjusted using third component (Bi, Na)TiO3 additive. Jpn J Appl Phys 52: 09KD11. http://doi.org/10.7567/jjap.52.09kd11 doi: 10.7567/jjap.52.09kd11
    [110] Chao X, Yang Z, Li Z, et al. (2012) Phase structures, electrical properties and temperature stability of (1-x)[(K0.458Na0.542)0.96Li0.04](Nb0.85Ta0.15)O3-xBiFeO3 ceramics. J Alloys Compd 518: 1–5. http://doi.org/10.1016/j.jallcom.2011.11.104 doi: 10.1016/j.jallcom.2011.11.104
    [111] Zhou JJ, Li JF, Cheng LQ, et al. (2012) Addition of small amounts of BiFeO3 to (Li, K, Na)(Nb, Ta)O3 lead-free ceramics: Influence on phase structure, microstructure and piezoelectric properties. J Eur Ceram Soc 32: 3575–3582. http://doi.org/10.1016/j.jeurceramsoc.2012.05.019 doi: 10.1016/j.jeurceramsoc.2012.05.019
    [112] Li X, Zhu J, Wang M, et al. (2010) BiScO3-modified (K0.475Na0.475Li0.05)(Nb0.95Sb0.05)O3 lead-free piezoelectric ceramics. J Alloys Compd 499: L1–L4. http://doi.org/10.1016/j.jallcom.2010.01.129 doi: 10.1016/j.jallcom.2010.01.129
    [113] Zhang C, Chen Z, Ji WJ, et al. (2011) Crystal structures and electrical properties of (1-x)K0.5Na0.5NbO3-xBi0.8La0.2FeO3 lead-free ceramics. J Alloys Compd 509: 2425–2429. http://doi.org/10.1016/j.jallcom.2010.11.037 doi: 10.1016/j.jallcom.2010.11.037
    [114] Wu W, Xiao D, Wu J, et al. (2011) Polymorphic phase transition-induced electrical behavior of BiCoO3-modified (K0.48Na0.52)NbO3 lead-free piezoelectric ceramics. J Alloys Compd 509: L284–L288. http://doi.org/10.1016/j.jallcom.2011.05.004 doi: 10.1016/j.jallcom.2011.05.004
    [115] Yang H, Zhou C, Zhou Q, et al. (2013) Lead-free (Li, Na, K)(Nb, Sb)O3 piezoelectric ceramics: Effect of Bi(Ni0.5Ti0.5)O3 modification and sintering temperature on microstructure and electrical properties. J Mater Sci 48: 2997–3002. http://doi.org/10.1007/s10853-012-7078-7 doi: 10.1007/s10853-012-7078-7
    [116] Liu Y, Chu R, Xu Z, et al. (2011) Effects of BiAlO3 on structure and electrical properties of K0.5Na0.5NbO3-LiSbO3 lead-free piezoceramics. Mater Sci Eng B 176: 1463–1466. http://doi.org/10.1016/j.mseb.2011.09.002 doi: 10.1016/j.mseb.2011.09.002
    [117] Zuo R, Fang X, Ye C (2007) Phase structures and electrical properties of new lead-free (Na0.5K0.5)NbO3-(Bi0.5Na0.5)TiO3 ceramics. Appl Phys Lett 90: 092904. http://doi.org/10.1063/1.2710768 doi: 10.1063/1.2710768
    [118] Du H, Zhou W, Luo F, et al. (2008) High Tm lead-free relaxor ferroelectrics with broad temperature usage range: 0.04BiScO3-0.96(K0.5Na0.5)NbO3. J Appl Phys 104: 044104. http://doi.org/10.1063/1.2969773 doi: 10.1063/1.2969773
    [119] Du H, Zhou W, Luo F, et al. (2008) Polymorphic phase transition dependence of piezoelectric properties in (K0.5Na0.5)NbO3-(Bi0.5K0.5)TiO3 lead-free ceramics. J Phys D Appl Phys 41: 115413. http://doi.org/10.1088/0022-3727/41/11/115413 doi: 10.1088/0022-3727/41/11/115413
    [120] Jiang XP, Yang Q, Yu ZD, et al. (2010) Microstructure and electrical properties of Li0.5Bi0.5TiO3-modified (Na0.5K0.5)NbO3 lead-free piezoelectric ceramics. J Alloys Compd 493: 276–280. http://doi.org/10.1016/j.jallcom.2009.12.079 doi: 10.1016/j.jallcom.2009.12.079
    [121] Wang R, Xie RJ, Hanada K, et al. (2008) Enhanced piezoelectricity around the tetragonal/orthorhombic morphotropic phase boundary in (Na, K)NbO3-ATiO3 solid solutions. J Electroceramics 21: 263–266. http://doi.org/10.1007/s10832-007-9136-0 doi: 10.1007/s10832-007-9136-0
    [122] Park HY, Cho KH, Paik DS, et al. (2007) Microstructure and piezoelectric properties of lead-free (1-x)(Na0.5K0.5)NbO3-xCaTiO3 ceramics. J Appl Phys 102: 124101. http://doi.org/10.1063/1.2822334 doi: 10.1063/1.2822334
    [123] Kim MR, Song HC, Choi JW, et al. (2009) Synthesis and piezoelectric properties of (1-x)(Na0.5K0.5)NbO3-x(Ba0.95Sr0.05)TiO3 ceramics. J Electroceramics 23: 502–505. http://doi.org/10.1007/s10832-008-9519-x doi: 10.1007/s10832-008-9519-x
    [124] Chen Z, He X, Yu Y, et al. (2009) Piezoelectric and dielectric properties of (Na0.5K0.5)NbO3-(Bi0.5Na0.5)0.94Ba0.06TiO3 lead-free piezoelectric ceramics. Jpn J Appl Phys 48: 030204. http://doi.org/10.1143/JJAP.48.030204 doi: 10.1143/JJAP.48.030204
    [125] Zuo R, Lv D, Fu J, et al. (2009) Phase transition and electrical properties of lead free (Na0.5K0.5)NbO3-BiAlO3 ceramics. J Alloys Compd 476: 836–839. http://doi.org/10.1016/j.jallcom.2008.09.123 doi: 10.1016/j.jallcom.2008.09.123
    [126] Du H, Zhou W, Luo F, et al. (2008) Design and electrical properties' investigation of (K0.5Na0.5) NbO3-BiMeO3 lead-free piezoelectric ceramics. J Appl Phys 104: 034104. https://doi.org/10.1063/1.2964100 doi: 10.1063/1.2964100
    [127] Tao H, Wu J, Zheng T, et al. (2015) New (1-x)K0.45Na0.55Nb0.96Sb0.04O3-xBi0.5Na0.5HfO3 lead-free ceramics: Phase boundary and their electrical properties. J Appl Phys 118: 044102. http://doi.org/10.1063/1.4927281 doi: 10.1063/1.4927281
    [128] Gao Y, Zhang J, Qing Y, et al. (2011) Remarkably strong piezoelectricity of lead-free (K0.45Na0.55)0.98Li0.02(Nb0.77Ta0.18Sb0.05)O3 ceramic. J Am Ceram Soc 94: 2968–2973. http://doi.org/10.1111/j.1551-2916.2011.04468.x doi: 10.1111/j.1551-2916.2011.04468.x
    [129] Xu P, Jiang M, Liu X (2011) Effects of low concentration BiFeO3 additions on microstructure and piezoelectric properties of (K0.5Na0.5)NbO3 ceramics. Adv Mat Res 335–336: 968–975. http://doi.org/10.4028/www.scientific.net/AMR.335-336.968 doi: 10.4028/www.scientific.net/AMR.335-336.968
    [130] Deng Y, Wang J, Zhang C, et al. (2020) Structural and electric properties of MnO2-doped KNN-LT lead-free piezoelectric ceramics. Crystals 10: 1–8. http://doi.org/10.3390/cryst10080705 doi: 10.3390/cryst10080705
    [131] Zhang Y, Zhai J, Xue S (2020) Effect of three step sintering on piezoelectric properties of KNN-based lead-free ceramics. Chem Phys Lett 758: 137906. http://doi.org/10.1016/j.cplett.2020.137906 doi: 10.1016/j.cplett.2020.137906
    [132] Jiang J, Li H, Zhao C, et al. (2022) Broad-temperature-span and improved piezoelectric/dielectric properties in potassium sodium niobate-based ceramics through diffusion phase transition. J Alloys Compd 925: 166708. http://doi.org/10.1016/j.jallcom.2022.166708 doi: 10.1016/j.jallcom.2022.166708
    [133] Liu Y, Pan Y, Bai X, et al. (2021) Structure evolution and piezoelectric properties of K0.40Na0.60Nb0.95Sb0.05O3-Bi0.5K0.5HfO3-SrZrO3 ternary lead-free ceramics with R-O-T phase boundary. J Mater Sci Mater Electron 32: 9032–9043. http://doi.org/10.1007/s10854-021-05573-7 doi: 10.1007/s10854-021-05573-7
    [134] Jiang M, Liu X, Chen G (2009) Phase structures and electrical properties of new lead-free Na0.5K0.5NbO3-LiSbO3-BiFeO3 ceramics. Scripta Mater 60: 909–912. http://doi.org/10.1016/j.scriptamat.2009.02.017 doi: 10.1016/j.scriptamat.2009.02.017
    [135] Minhong J, Manjiao D, Huaxin L, et al. (2011) Piezoelectric and dielectric properties of K0.5Na0.5NbO3-LiSbO3-BiScO3 lead-free piezoceramics. Mater Sci Eng B 176: 167–170. http://doi.org/10.1016/j.mseb.2010.10.007 doi: 10.1016/j.mseb.2010.10.007
    [136] Li F, Tan Z, Xing J, et al. (2017) Investigation of new lead free (1−x)KNNS-xBKZH piezo-ceramics with R-O-T phase boundary. J Mater Sci Mater Electron 28: 8803–8809. http://doi.org/10.1007/s10854-017-6607-1 doi: 10.1007/s10854-017-6607-1
    [137] Li F, Gou Q, Xing J, et al. (2017) The piezoelectric and dielectric properties of sodium-potassium niobate ceramics with new multiphase boundary. J Mater Sci Mater Electron 28: 18090–18098. http://doi.org/10.1007/s10854-017-7753-1 doi: 10.1007/s10854-017-7753-1
    [138] Qiao L, Li G, Tao H, et al. (2020) Full characterization for material constants of a promising KNN-based lead-free piezoelectric ceramic. Ceram Int 46: 5641–5644. http://doi.org/10.1016/j.ceramint.2019.11.009 doi: 10.1016/j.ceramint.2019.11.009
    [139] Shi C, Ma J, Wu J, et al. (2020) Coexistence of excellent piezoelectric performance and high curie temperature in KNN-based lead-free piezoelectric ceramics. J Alloys Compd 846: 156245. http://doi.org/10.1016/j.jallcom.2020.156245 doi: 10.1016/j.jallcom.2020.156245
    [140] Pan D, Guo Y, Zhang K, et al. (2017) Phase structure, microstructure, and piezoelectric properties of potassium-sodium niobate-based lead-free ceramics modified by Ca. J Alloys Compd 693: 950–954. http://doi.org/10.1016/j.jallcom.2016.09.277 doi: 10.1016/j.jallcom.2016.09.277
    [141] Wang C, Fang B, Qu Y, et al. (2020) Preparation of KNN based lead-free piezoelectric ceramics via composition designing and two-step sintering. J Alloys Compd 832: 153043. http://doi.org/10.1016/j.jallcom.2019.153043 doi: 10.1016/j.jallcom.2019.153043
    [142] Politova ED, Kaleva GM, Mosunov AV, et al. (2020) Dielectric and local piezoelectric properties of lead-free KNN-based perovskite ceramics. Ferroelectrics 569: 201–208. http://doi.org/10.1080/00150193.2020.1822677 doi: 10.1080/00150193.2020.1822677
    [143] Brenda CJ, Armando RM, Villafuerte CME, et al. (2018) Piezoelectric, dielectric and ferroelectric properties of (1-x)(K0.48Na0.52)0.95Li0.05Nb0.95Sb0.05O3-xBa0.5(Bi0.5Na0.5)0.5ZrO3 lead-free solid solution. J Electron Mater 47: 6053–6058. http://doi.org/10.1007/s11664-018-6488-y doi: 10.1007/s11664-018-6488-y
    [144] Zhang K, Guo Y, Pan D, et al. (2016) Phase transition and piezoelectric properties of dense (K0.48Na0.52)0.95Li0.05SbxNb(1-x)O3-0.03Ca0.5(Bi0.5Na0.5)0.5ZrO3 lead free ceramics. J Alloys Compd 664: 503–509. http://doi.org/10.1016/j.jallcom.2015.12.256 doi: 10.1016/j.jallcom.2015.12.256
    [145] Wendari TP, Arief S, Mufti N, et al. (2020) Ratio effect of salt fluxes on structure, dielectric and magnetic properties of La, Mn-doped PbBi2Nb2O9 aurivillius phase. Ceram Int 46: 14822–14827. http://doi.org/10.1016/j.ceramint.2020.03.007 doi: 10.1016/j.ceramint.2020.03.007
    [146] Difeo M, Rubio-Marcos F, Gibbs F, et al. (2023) Effect of antimony content on transition behavior and electric properties of (K0.44Na0.52Li0.04)(Nb0.9-xTa0.10Sbx)O3 ceramics. Appl Sci 13: 992. https://doi.org/10.3390/app13020992 doi: 10.3390/app13020992
    [147] Wu J (2018) Advances in Lead-Free Piezoelectric Materials, Singapore: Springer Singapore. https://doi.org/10.1007/978-981-10-8998-5
    [148] Nandini RN, Krishna M, Suresh AV, et al. (2018) Effect of MWCNTs on piezoelectric and ferroelectric properties of KNN composites. Mater Sci Eng B 231: 40–56. http://doi.org/10.1016/j.mseb.2018.09.001 doi: 10.1016/j.mseb.2018.09.001
    [149] Shirane G, Newnham R, Pepinsky R (1954) Dielectric properties and phase transitions of NaNbO3 and (Na, K)NbO3. Phys Rev 96: 581–588. http://doi.org/10.1103/PhysRev.96.581 doi: 10.1103/PhysRev.96.581
    [150] Jaeger Re, Egerton L (1962) Hot pressing of potassium-sodium niobates. J Am Ceram Soc 45: 209–213. http://doi.org/10.1111/j.1151-2916.1962.tb11127.x doi: 10.1111/j.1151-2916.1962.tb11127.x
    [151] Zhang B, Wu J, Wang X, et al. (2013) Rhombohedral-orthorhombic phase coexistence and electrical properties of Ta and BaZrO3 co-modified (K, Na)NbO3 lead-free ceramics. Curr Appl Phys 13: 1647–1650. http://doi.org/10.1016/j.cap.2013.06.010 doi: 10.1016/j.cap.2013.06.010
    [152] Lv YG, Wang CL, Zhang JL, et al. (2009) Tantalum influence on physical properties of (K0.5Na0.5)(Nb1-xTax)O3 ceramics. Mater Res Bull 44: 284–287. http://doi.org/10.1016/j.materresbull.2008.06.019 doi: 10.1016/j.materresbull.2008.06.019
    [153] Zuo R, Ye C, Fang X (2007) Dielectric and piezoelectric properties of lead free Na0.5K0.5NbO3-BiScO3 ceramics. Jpn J Appl Phys 46: 6733–6736. http://doi.org/10.1143/jjap.46.6733 doi: 10.1143/JJAP.46.6733
    [154] Zhou C, Zhang J, Yao W, et al. (2018) Piezoelectric performance, phase transitions and domain structure of 0.96(K0.48Na0.52)(Nb0.96Sb0.04)O-0.04(Bi0.50Na0.50)ZrO3 ceramics. J Appl Phys 124: 164101. http://doi.org/10.1063/1.5048345 doi: 10.1063/1.5048345
    [155] Zhang Y, Li P, Shen B, et al. (2018) Effect of shifting orthorhombic-tetragonal phase transition on structure and properties of K0.5Na0.5NbO3-based lead-free ceramics. J Alloys Compd 735: 1328–1330. http://doi.org/10.1016/j.jallcom.2017.11.281 doi: 10.1016/j.jallcom.2017.11.281
    [156] Guo Y, Kakimoto KI, Ohsato H (2004) Phase transitional behavior and piezoelectric properties of (Na0.5K0.5)NbO3-LiNbO3 ceramics. Appl Phys Lett 85: 4121–4123. http://doi.org/10.1063/1.1813636 doi: 10.1063/1.1813636
    [157] Guo Y, Kakimoto K, Ohsato H (2005) (Na0.5K0.5)NbO3-LiTaO3 lead-free piezoelectric ceramics. Mater Lett 59: 241–244. https://doi.org/10.1016/j.matlet.2004.07.057 doi: 10.1016/j.matlet.2004.07.057
    [158] Wang K, Li JF, Liu N (2008) Piezoelectric properties of low-temperature sintered Li-modified (Na, K)NbO3 lead-free ceramics. Appl Phys Lett 93: 092904. http://doi.org/10.1063/1.2977551 doi: 10.1063/1.2977551
    [159] Wang R, Bando H, Itoh M (2009) Universality in phase diagram of (K, Na)NbO3-MTiO3 solid solutions. Appl Phys Lett 95: 092905. http://doi.org/10.1063/1.3224196 doi: 10.1063/1.3224196
    [160] Wang R, Bando H, Kidate M, et al. (2011) Effects of A-site ions on the phase transition temperatures and dielectric properties of (1–x)(Na0.5K0.5)NbO3-xAZrO3 solid solutions. Jpn J Appl Phys 50: 09ND10. http://doi.org/10.1143/JJAP.50.09ND10 doi: 10.1143/JJAP.50.09ND10
    [161] Zheng T, Wu J, Cheng X, et al. (2014) Wide phase boundary zone, piezoelectric properties, and stability in 0.97(K0.4Na0.6)(Nb1-xSbx)O3-0.03Bi0.5Li0.5ZrO3 lead-free ceramics. Dalton Trans 43: 9419–9426. http://doi.org/10.1039/c4dt00768a doi: 10.1039/c4dt00768a
    [162] Fong DD, Kolpak AM, Eastman JA, et al. (2006) Stabilization of monodomain polarization in ultrathin PbTiO3 films. Phys Rev Lett 96: 127601. http://doi.org/10.1103/PhysRevLett.96.127601 doi: 10.1103/PhysRevLett.96.127601
    [163] Wang X, Wu J, Xiao D, et al. (2014) Giant piezoelectricity in potassium-sodium niobate lead-free ceramics. J Am Chem Soc 136: 2905–2910. http://doi.org/10.1021/ja500076h doi: 10.1021/ja500076h
    [164] Wang X, Wu J, Cheng X, et al. (2013) Compositional dependence of phase structure and electrical properties in (K0.50Na0.50)0.97Bi0.01(Nb1-xZrx)O3 lead-free ceramics. Ceram Int 39: 8021–8024. http://doi.org/10.1016/j.ceramint.2013.03.071 doi: 10.1016/j.ceramint.2013.03.071
    [165] Zheng T, Wu J, Cheng X, et al. (2014) New potassium-sodium niobate material system: A giant-d33 and high-TC lead-free piezoelectric. Dalton Trans 43: 11759–11766. http://doi.org/10.1039/c4dt01293c doi: 10.1039/c4dt01293c
    [166] Yang W, Li P, Wu S, et al. (2020) Coexistence of excellent piezoelectric performance and thermal stability in KNN-based lead-free piezoelectric ceramics. Ceram Int 46: 1390–1395. http://doi.org/10.1016/j.ceramint.2019.09.102 doi: 10.1016/j.ceramint.2019.09.102
    [167] Wong JYY, Zhang N, Ye ZG (2016) High-temperature solution growth and vapour transport equilibration of (1−x)K1−yNayNbO3-xLiNbO3 lead-free piezo-/ferroelectric single crystals. J Cryst Growth 452: 125–130. http://doi.org/10.1016/j.jcrysgro.2016.01.022 doi: 10.1016/j.jcrysgro.2016.01.022
    [168] Priya S, Nahm S (2013) Lead-Free Piezoelectrics, New York: Springer New York. https://doi.org/10.1007/978-1-4419-9598-8
    [169] Condurache OA, Radan K, Prah U, et al. (2019) Heterogeneity challenges in multiple-element-modified lead-free piezoelectric ceramics. Materials 12: 4049. https://doi.org/10.3390/ma12244049 doi: 10.3390/ma12244049
    [170] Chen R, Jiang L, Zhang T, et al. (2019) Eco-friendly highly sensitive transducers based on a new KNN-NTK-FM lead-free piezoelectric ceramic for high-frequency biomedical ultrasonic imaging applications. IEEE Trans Biomed Eng 66: 1580–1587. http://doi.org/10.1109/TBME.2018.2876063 doi: 10.1109/TBME.2018.2876063
    [171] Zhao Z, Lv Y, Dai Y, et al. (2020) Ultrahigh electro-strain in acceptor-doped KNN lead-free piezoelectric ceramics via defect engineering. Acta Mater 200: 35–41. http://doi.org/10.1016/j.actamat.2020.08.073 doi: 10.1016/j.actamat.2020.08.073
    [172] Zhang Y, Li JF (2019) Review of chemical modification on potassium sodium niobate lead-free piezoelectrics. J Mater Chem C 7: 4284–4303. http://doi.org/10.1039/c9tc00476a doi: 10.1039/c9tc00476a
    [173] Schütz D, Deluca M, Krauss W, et al. (2012) Lone-pair-induced covalency as the cause of temperature and field-induced instabilities in bismuth sodium titanate. Adv Funct Mater 22: 2285–2294. http://doi.org/10.1002/adfm.201102758 doi: 10.1002/adfm.201102758
    [174] Yin J, Zhao C, Zhang Y, et al. (2018) Ultrahigh strain in site engineering-independent Bi0.5Na0.5TiO3-based relaxor-ferroelectrics. Acta Mater 147: 70–77. http://doi.org/10.1016/j.actamat.2018.01.054 doi: 10.1016/j.actamat.2018.01.054
    [175] Akmal MHM, Warikh ARM, Azlan UAA, et al. (2016) Structural evolution and dopant occupancy preference of yttrium-doped potassium sodium niobate thin films. J Electroceramics 37: 50–57. http://doi.org/10.1007/s10832-016-0039-9 doi: 10.1007/s10832-016-0039-9
    [176] Akmal MHM, Warikh ARM (2021) Electrical behaviour of yttrium-doped potassium sodium niobate thin film for piezoelectric energy harvester applications. J Aust Ceram Soc 57: 589–596. http://doi.org/10.1007/s41779-021-00569-2 doi: 10.1007/s41779-021-00569-2
    [177] Knauth P (2006) Ionic and electronic conduction in nanostructured solids: Concepts and concerns, consensus and controversies. Solid State Ionics 177: 2495–2502. http://doi.org/10.1016/j.ssi.2006.02.039 doi: 10.1016/j.ssi.2006.02.039
    [178] Maziati A, Umar A, Warikh M, et al. (2015) Enhanced structural and electrical properties of lead-free Y-doped (K, Na)NbO3 thin films. Jurnal Teknologi 77: 67–71. https://doi.org/10.11113/jt.v77.6609 doi: 10.11113/jt.v77.6609
    [179] Li Y, Chen W, Xu Q, et al. (2007) Piezoelectric and dielectric properties of CeO2-doped Bi0.5Na0.44K0.06TiO3 lead-free ceramics. Ceram Int 33: 95–99. http://doi.org/10.1016/j.ceramint.2005.08.001 doi: 10.1016/j.ceramint.2005.08.001
    [180] Deng A, Wu J (2020) Effects of rare-earth dopants on phase structure and electrical properties of lead-free bismuth sodium titanate-based ceramics. J Materiomics 6: 286–292. http://doi.org/10.1016/j.jmat.2020.03.005 doi: 10.1016/j.jmat.2020.03.005
    [181] Bathelt R, Soller T, Benkert K, et al. (2012) Neodymium doping of KNNLT. J Eur Ceram Soc 32: 3767–3772. http://doi.org/10.1016/j.jeurceramsoc.2012.05.025 doi: 10.1016/j.jeurceramsoc.2012.05.025
    [182] Akmal M, Hatta M, Warikh ARM, et al. (2016) Influence of yttrium dopant on the structure and electrical conductivity of potassium sodium niobate thin films. Mat Res 19: 1417–1422. https://doi.org/10.1590/1980-5373-MR-2016-0076 doi: 10.1590/1980-5373-mr-2016-0076
    [183] Lv X, Zhu J, Xiao D, et al. (2020) Emerging new phase boundary in potassium sodium-niobate based ceramics. Chem Soc Rev 49: 671–707. http://doi.org/10.1039/c9cs00432g doi: 10.1039/c9cs00432g
    [184] Lv X, Zhang J, Liu Y, et al. (2020) Synergetic contributions in phase boundary engineering to the piezoelectricity of potassium sodium niobate lead-free piezoceramics. ACS Appl Mater Interfaces 12: 39455–39461. http://doi.org/10.1021/acsami.0c12424 doi: 10.1021/acsami.0c12424
    [185] Ma ZY, Zheng HJ, Zhao L, et al. (2023) Comparison of impact from typical additives for phase structure in (K, Na)NbO3-based ceramics. Ceram Int 49: 18629–18637. http://doi.org/10.1016/j.ceramint.2023.02.239 doi: 10.1016/j.ceramint.2023.02.239
    [186] Dahiya A, Thakur OP, Juneja JK (2013) Sensing and actuating applications of potassium sodium niobate: Use of potassium sodium niobate in sensor and actuator. 2013 Seventh International Conference on Sensing Technology (ICST), New Zealand, 383–386.
    [187] Maziati Akmal MH, Warikh ARM, Azlan UAA, et al. (2017) The effects of different annealing temperatures and number of deposition layers on the crystallographic properties sodium niobate (KNN) thin films synthesized by sol-gel spin coating technique. J Adv Manuf Technol 11: 91–102. Available from: https://jamt.utem.edu.my/jamt/article/view/1232.
    [188] Akmal MHM, Warikh ARM, Azlan UAA, et al. (2018) Optimizing the processing conditions of sodium potassium niobate thin films prepared by sol-gel spin coating technique. Ceram Int 44: 317–325. https://doi.org/10.1016/j.ceramint.2017.09.175 doi: 10.1016/j.ceramint.2017.09.175
  • This article has been cited by:

    1. Di You, Shaofeng Zhang, Jichao Liu, Wei Shi, Qian Xiao, Yong Tang, Hua Tan, Feng Chen, Sixian Ye, Tran Nguyen Minh An, Junfeng Wang, Junya Wang, Ting Zhang, Haibo Zhang, Enhanced Strain Properties and Temperature Stability in Textured Potassium Niobate Sodium-Based Multilayer Piezoelectric Actuators, 2024, 2637-6113, 10.1021/acsaelm.4c01684
    2. Rahul Bhatnagar, Varsha Yadav, Upendra Kumar, Marcos Flores Carrasco, Piezoelectric Energy Harvesting: From Fundamentals to Advanced Applications, 2024, 2194-4288, 10.1002/ente.202401455
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2657) PDF downloads(260) Cited by(2)

Figures and Tables

Figures(2)  /  Tables(2)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog