Next Article in Journal
Effect of LDHs and Other Clays on Polymer Composite in Adsorptive Removal of Contaminants: A Review
Next Article in Special Issue
Spectroscopic, Raman, EMPA, Micro-XRF and Micro-XANES Analyses of Sulphur Concentration and Oxidation State of Natural Apatite Crystals
Previous Article in Journal
Synthesis of Enaminones-Based Benzo[d]imidazole Scaffold: Characterization and Molecular Insight Structure
Previous Article in Special Issue
Control of Thermophysical Properties of Langasite-Type La3Ta0.5Ga5.5O14 Crystals for Pressure Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molecular Orbital Calculation of Lead-Free Perovskite Compounds for Efficient Use of Alkaline and Alkaline Earth Metals

1
Department of Applied Physics, Graduate School of Science, Fukuoka University, Fukuoka 814-0180, Japan
2
Japan Atomic Energy Agency, Tsuruga 919-1279, Japan
*
Author to whom correspondence should be addressed.
Submission received: 28 August 2020 / Revised: 20 October 2020 / Accepted: 20 October 2020 / Published: 22 October 2020

Abstract

:
The effective ionic charges of lead-free perovskite dielectric complex compounds were investigated with molecular orbital calculation. The base model was a double perovskite cluster that consisted of octahedral oxygen cages with a transition metal ion of titanium, niobium, or zirconium located at each of their centers, and alkali and/or alkaline earth metal ions located at the body center, corners, edge centers, or face centers of the cluster. The results showed significant covalent bonds between the transition metals and the oxygens, and the alkali metals, especially sodium and oxygen. On the other hand, the alkaline earth metals have weak covalency. Calculation was also performed with the replacement of some of the oxygens with chlorine or fluorine; such replacement enhances the covalency of the transition metals. These trends provide good guidelines for the design properties of lead-free perovskite piezoelectrics based on ubiquitous sodium use.

1. Introduction

Lead (Pb)-based piezoelectrics have exclusively high electric polarization and piezoelectricity [1,2,3,4,5]. The typical substance is lead titanate, designated as PbTiO3 (PT) [6,7]. Its crystal structure is of a typical simple perovskite, which is denoted as ABO3 in Figure 1, where A, B, and O are ionic species; A and B each corresponds to cations of lead and titanium, and O is an anion of oxygen that forms an octahedral cage with the five other O ions. The chemical bonding of lead titanate was studied with first-principles calculation by R. E. Cohen [8]. The results indicated that the lead (Pb) ion is significantly covalent with the oxygen (O) and titanium (Ti) ions. This feature suggests that the covalency forms a molecular orbital (i.e., electronic channels) through which valence electrons are transferred back and forth between the cations and anions in response to external fields. The response may be soft or hard, depending on the significance of the covalency. Thus, covalency plays an important role in causing the piezoelectric strain, which is a reason PT and other Pb-based systems have excellent properties. Therefore, other ionic species that cause covalency at the site of Pb can be substitutes for harmless, lead-free piezoelectrics and their promising application. One of the authors, N. Takesue [9], confirmed with molecular orbital calculation that (Bi1/2Na1/2)TiO3 (BNT) [10,11] could be such a substitute. This substance is now widely recognized as a potential substitute and has been studied through many types of fabrication [11,12,13], long before N. Takesue’s confirmation [9]. In addition to BNT, many excellent Pb-free piezoelectrics have been discovered [14,15,16].
For piezoelectric perovskites, alkaline and alkaline-earth metals are traditionally regarded as representative ionic species occupying the site of Pb or A [17]. Although there are more modern systems such as BNT [10,11,12,13] and BNT-based solid solutions [18], we focused on traditional systems, namely, BiFeO3 [19] and GaFeO3 [20], since our work was on the efficient use of ubiquitous alkaline and alkaline-earth metals. Both species of traditional systems are ionic; their formal ionic charges are 1+ and 2+, respectively. However, this idea indicates only the valence charge of neutral atoms. A portion of the charge is transferred to neighboring atoms in substances consisting of more than one atom. Therefore, atoms—even those of alkaline and alkaline-earth metals—are not neutral in general. Rather, they are partially covalent and ionic, especially for metals with high-electronegativity atoms such as oxygen atoms as neighbors. This idea suggests that a table of effective ionic charges enables the control of the properties of lead-free piezoelectrics [8,21] and is considered compensative to tolerance factors [22,23], which is a well-established criterion for the stability of perovskite structures.
Among the representative species, alkaline metals (sodium [Na] and potassium [K]) and alkaline-earth metals (calcium [Ca] and barium [Ba]) are ubiquitous; in particular, sodium contained in marine resources is abundant. These are constituents of perovskite piezoelectrics such as NaNbO3 (Nb is niobium), KNbO3, CaTiO3, and BaTiO3, and they occupy the Pb site. We included BaZrO3 (zirconium [Zr] is chemically similar to Ti) in this category despite its cubic crystal structure (i.e., nonpolar) because it is frequently used as a solute component for this kind of piezoelectrics for their solid solutions.
The chemical bonding of BaTiO3 was also studied by R. E. Cohen [8]. The results showed that an effective charge of Ba was about 2+ (i.e., ionic) and a strong covalency of Ti with O, similar to the case of PT, was observed. However, the other elements including those in the alkaline and alkaline-earth columns of the periodic table, have not been considerably investigated. Na and Ca may be covalent because of their higher electronegativity than that of Ba. If their effective charges show an evident systematic variation along the two columns, then the piezoelectricity of these kinds of substances can be controlled through the synthesis of solid solutions and/or heterogeneous structuralization such as coherent interfaces; the way(s) may be realized by matching the one base piezoelectric with the other piezoelectric component(s). The successful case leads us to certain achievements, after which lead-free substance(s) with properties are as good as, or better than, those of Pb-based systems will become obtainable. This scheme motivated us to investigate the effective charges of perovskites with alkaline and/or alkaline-earth metals located at the Pb site including the charges of PT as the reference data through molecular orbital calculation. Then, we performed a calculation including the partial replacement of O with a halogen such as chlorine (Cl) and fluorine (F), or nitrogen (N) on the other side to investigate their electronegative effects on the effective charges of the cations.

2. Calculation

A discrete-variational (DV) Xα method of molecular orbital calculation [24,25] was used throughout this study by using the SCAT code for calculating the electronic states. First, the covalency and ionicity of alkaline metals lithium (Li), Na, and K (designated as MA) were investigated based on a fundamental compound cluster model (MA)2O with an antifluorite structure, as illustrated in Figure 2a. Previously reported lattice parameters [26] were used for these calculations. Second, the alkaline-earth metals magnesium (Mg), Ca, strontium (Sr), and Ba (designated as MAE) were investigated based on a cluster model (MAE)O with a sodium chloride structure, as illustrated in Figure 2b. The same lattice parameters [26] were used for these calculations.
In the third step, the perovskite was investigated based on double perovskite cluster models for PT, MANbO3, MAETiO3, BaZrO3, (Ba, Na)(Ti, Nb)O3, (Ba, Na)Ti(O, Cl)3, Na(Ti, Nb)(O, Cl)3, (Ba, Na)Ti(O, F)3, Na(Ti, Nb)(O, F)3, (Ba, Na)Nb(O, N)3, and Ba(Ti, Nb)(O, N)3, as illustrated in Figure 2c. The sites of ions in each model are indicated by colored spheres, and the correspondence of the colors to the ions is summarized in Table 1. Each model is a rectangular cluster and was given a nonpolar symmetry D2 (Schoenflies notation), common to all perovskite models employed; the reason is that the preliminary calculation with ferroelectric strains with a smaller symmetry than D2 indicated no essential effect on the bonds [9]. Furthermore, this simplified modeling reduced the time needed to complete the calculations while providing the essential chemical features. A typical value of the primitive cubic perovskite, 0.4 nm [7,17], was employed for the lattice parameters, so each of the three edge lengths of each model was given a value of 0.8 nm. The calculation on only PT was followed by a calculation based on its model, which incrementally provided an isotropic compression and expansion of 95%–105%; therefore, the above-mentioned effect of the electronic channels can be investigated [8,21].
All the above calculations were performed with the Madelung potential applied to each model, and Mulliken population analysis was conducted to determine the effective charges [24,25,27,28]. The lattice parameters used for modeling are listed in Table 2.

3. Results

The effective charges of the cations of each (MA)2O is listed in Table 3 in a downward sequence of the MA column in the periodic table. The charges are shown as a histogram in Figure 3a. Each effective charge is the average of the eight MA values in each model. The table and the figure indicate that the effective charges were noticeably below the formal charge of +1; in particular, the difference for sodium was significant. This evident charge transfer is considered to be caused by a large difference in electronegativity between neighboring MA and O; this transfer can be used to control the covalency and ionicity of the perovskite A site.
The effective charges of the cations of each (MAE)O are listed in Table 4 in the same manner as those of (MA)2O and are drawn as a histogram in Figure 3b. Two types of sodium chloride cluster models, namely, MAE- and O-body-centered, for each (MAE)O were employed. For the former model, the effective charge of MAE at the body center, bonded with the six nearest neighbor oxygens, was employed.
For the latter model, each of the eight MAE species at the corners was bonded with the three nearest neighbors, and each of the six at the face centers with the five; thus, the effective charges at these two non-body-centered locations were multiplied by 2 and 6/5, respectively, and the weighted values were averaged over each model.
The table and the figure indicate that the effective charges of MAE increased downward along the column toward the formal charge of +2. This finding differed from those of MA and was also related to the possible electronegativity of MAE and O. Therefore, the results of MAE may provide another aspect to control the chemical bonds of the perovskite A site.
The effective charges of the cations of the double perovskites of PT, MANbO3, MAETiO3, and BaZrO3 are listed in Table 5 and drawn as a histogram in Figure 4. The effective charge of each of the A-site ions Pb, MA, and MAE, each located at the body center, was employed; that of each of the B-site ions in the cages was averaged over each model. The table and the figure showed variations in the effective charges of MA and MAE for the perovskite models, similar to those along the columns of the periodic table for (MA)2O and (MAE)O, respectively. For the effective charges of the B site, Nb showed a value of almost +3, regardless of the value of MA in Table 5. Apart from PT, the values of Ti increased downward along the periodic column of MAE. This variation of Ti is similar to that of MAE and considered relevant to its electronegativity. The values of Zr of BaZrO3, which lies immediately below Ti along the periodic column, were higher than those of Ti. This variation was the same as that of MAE.
The effective charges of Pb of PT were approximately in the middle of the MA and MAE levels, and this Ti showed a smaller value than those of Ti and Zr in the other models. If the value of Pb (1.302) and Ti (2.369) were both magic numbers for the excellent performance of Pb-based substances, piezoelectricity may be designed or enhanced as a function of both the effective charges of the A and B sites averaged with respect to any arbitrarily modeled perovskite solid solution.
The effective charges of PT for each ratio of the isotropic deformation given for the model are listed in Table 6 and drawn as a histogram in Figure 5a. The effective charges were averaged as described above. The results show that the effective charges of Pb and Ti decreased with increasing expansion. Therefore, a compressed electronic state of the vacant 6s of Pb2+ and fully occupied 2p of O2− benefits from their ionic nature. The reason is that these oppositely charged ions, which are closely located, decrease the electrostatic energy [21]. In contrast, a covalent nature is favorable for the expanded state. The reason is that they reduce energy through their orbital hybridization of the low energy states [21]. The whole trend is illustrated in Figure 5b and is similar to the case of 3d0 of Ti4+ [21]. Therefore, the effective charges can be utilized as a measure of covalency, similar to the Born effective charge [29].
An example of averaging the effective charges is given below. The effective charges of our solid solution model (Ba, Na)(Ti, Nb)O3 are listed in Table 7 and drawn as a histogram in Figure 6. This case led us to assume two types of solid solution models, namely, Ba- and Na-body-centered. For the former model, the effective charge of Ba at the body center, bonded with the 12 nearest neighbor oxygens, was employed. For Na, each of the eight Na ions at the corners was bonded with the three nearest neighbors, and each of the six at the face centers with the eight. Thus, their charge transfers were multiplied by 4 and 12/8, respectively, to compute the effective charges at these two non-body-centered locations. These weighted transfers were each averaged over their equivalent locations of this model, and these averaged values were each subtracted from each formal charge to obtain each effective charge. Furthermore, the average effective charges of Ba and Na were further averaged in a manner we called the cluster chemical mean field. The further averaged value is also given in Table 7. In the same way, the effective charges of the latter model were obtained by reversing Ba and Na in the former model (reverse Na and Ba, respectively).
For the former and latter models, the effective charges of Ti were simply averaged over the four species, and the same was done for Nb. The averaged effective charges of Ti and Nb were further averaged, and this value is also given in Table 7.

4. Discussion and Guidelines for Designing Lead-Free Piezoelectrics

We checked the uniqueness of the above-mentioned magic numbers of PT and other Pb-based systems such as relaxor Pb(Mg1/3Nb2/3)O3, called PMN [30], which is a good component for high piezoelectric performance. Since there is no adequate symmetry for modeling the PMN cluster, we believed that the calculation of PMN would consume a long time. Thus, the calculation was not conducted; instead, the charges of Pb and Nb in Table 5 and that of Mg in Table 4 were employed to obtain the charges. The obtained values of A and B were approximately 1.3 and about 2.4, respectively, and these values for PMN were close to those for PT. Therefore, the magic numbers are likely true, although they must be verified through the fabrication of Pb-free substances, at least indirectly.
We also checked the relaxor PbZn1/3Nb2/3O3, called PZN [31], which is also a good component. The Zn contained in this substance did not belong to the related columns of the periodic table, so the effective charge was not calculated. However, the Zn of its oxide is expected to have a charge that is considerably less than the formal charge of +2 because Zn is a relatively highly electronegative metal. Let us assume that the effective charge of Zn in its oxide is +2, which is the maximum value. An estimated value of the effective charge of this B site, designated as Bavereged, given by averaging the formal charge of Zn with the charge of Nb of about +3 in Table 5, was about 2.7, which was evidently beyond the proposed magic number, but the inherent electronegativity of Zn causes Bavereged to be reduced in reality.
We also examined the well-known, excellent piezoelectric Pb(ZrxTi1−x)O3 (x~0.5) called PZT [32]. This Bavereged averaged with the charges of Ti and Zr in Table 5 was approximately 2.6, which was also considerably higher than the proposed magic number. Thus, the good property is considered purely as the morphotropic effect of PT with PbZrO3, which is the case with BT–BZ solid solutions [14,15,16].
A check of the uniqueness of the magic numbers did not reveal any morphotropic effect, piezoelectric coefficient, or electromechanical coupling coefficient, but it confirmed whether the effective charges of the modeled perovskites were close to the magic numbers. It also suggests the fabrication target systems that may have Pb–O-type covalency.
We set up two orthogonal axes so that they both lay on one plane. We assumed that one of the two axes indicated the averaged effective charges of either A or B ions for each model; the other axis, the other charge of B or A. In this way, both charges for each model give a coordinate in the plane. We marked, with a dot, each coordinate determined from the charges in Table 5, Table 6 and Table 7 including those of the BNT of the Na-centered model employed in a previous study [9]. The dots are given in Figure 7, which shows the following features.
  • The one circled dot is at the coordinate of the unstrained PT.
  • The three dots of MANbO3 are in the region where the effective charges of A and B are relatively low and high, respectively.
  • The three dots of MAETiO3 are in the region where the charges of A and B are high and low, respectively.
  • The two dots of (Ba, Na)(Ti, Nb)O3 are about in the middle of the regions of MANbO3 and MAETiO3.
  • The one dot of BaZrO3 is where the charges of A and B are both high.
  • The one dot of BNT is in the vicinity of PT.
Figure 7. Map of the values in Table 5, Table 6 and Table 7 including those of BNT (Bi1/2Na1/2TiO3) of the Na-centered model. LN, NN, KN are LiNbO3, NaNbO3, KNbO3. CT, ST, BT are CaTiO3, SrTiO3, and BaTiO3.
Figure 7. Map of the values in Table 5, Table 6 and Table 7 including those of BNT (Bi1/2Na1/2TiO3) of the Na-centered model. LN, NN, KN are LiNbO3, NaNbO3, KNbO3. CT, ST, BT are CaTiO3, SrTiO3, and BaTiO3.
Crystals 10 00956 g007
From the above features, the solid solutions of MANbO3 and MAETiO3 formed a narrow band from one region to the other, as shown in the figure. This band includes the locations of (Ba, Na)(Ti, Nb)O3, but not PT, so the lead-free property improvement through this synthesis of solid solutions is chemically limited. Moreover, BaZrO3 is not effective in the property improvement according to this idea. Nonetheless, this substance solely has a morphotropic effect with BT [14,15,16].
BNT is close to PT; BNT, which has been noteworthy for decades, is a successful case of solid solutions [10,11,12,13]. However, the bismuth (Bi) contained in BNT is volatile at high temperatures for sintering [33,34,35], which often deteriorates the quality during fabrication. Therefore, Bi may be an unwanted element from a fabrication point of view. Furthermore, the hardness of BNT is high compared with those of Pb-based systems. Therefore, an ultimate enhancement of the piezoelectricity may be naturally difficult, although BNT is a good piezoelectric, and solid solutions are promising because they can be fabricated at relatively low sintering temperatures [18], thereby suppressing vaporization.
Figure 7 shows that reducing the effective charge of the B site of (Ba, Na)(Ti, Nb)O3 improved its properties. The reduction can be done by replacing O with halogens such as F and Cl, which are more electronegative than O. The replacement is performed to enhance the covalency and reduce the charge of B, thus effectively mitigating the limitation of the property improvement. According to this idea, the replacement with N is the opposite case because N is less electronegative than O.
The results of the replacement are listed in Table 8 and drawn as a histogram in Figure 8. The four oxygens marked in Figure 2c were replaced with Cl, F, and N. Since their formal charges are different from that of O, the simple replacement causes the total charges, as the double perovskite unit cell, to be nonzero. Therefore, the zero charge was maintained by further replacing Ba with Na or the opposite, and Ti with Nb or the reverse, depending on each model given by replacing with Cl, F, and N. We marked at each coordinate with a dot as determined from the charges in Table 8. The table and figure indicate an expected reduction in the effective charges. The dots of the coordinates of the charges in the table are given in Figure 9. The dots of PT and (Ba, Na)(Ti, Nb)O3 in Figure 7 were also given to confirm any effect of the replacement. Figure 9 shows the two dots of (Ba, Na)Ti(O, Cl)3 and the two dots of (Ba, Na)Ti(O, F)3. Their locations were very close to PT, which means that the replacement with halogens is effective. Furthermore, Nb is not necessary in these substances; otherwise, the opposite effect is expected, since the charge of Nb is relatively high compared with that of Ti. The figure also shows the data of the other halogen-replaced models, namely, Na(Ti, Nb)(O, Cl)3, and Na(Ti, Nb)(O, F)3. The data indicate that the charges of A were excessively low because there was no Ba included, and B was also excessively high because Nb was included. Therefore, Ba is also effective for the design of piezoelectrics as well as halogens. Finally, as expected, the calculation based on the N-replaced models, namely, (Ba, Na)Nb(O, N)3 and Ba(Ti, Nb)(O, N)3, showed the reverse effect.

5. Conclusions

The effective ionic charges of perovskite dielectric complex compounds were investigated with molecular orbital calculation. Aside from PT, the results showed remarkable covalency for almost all cations and systematic variations in the effective charges of the A-site ions along their column in the periodic table. This variation corresponded to their electronegativity, and this was also the case with Ti and Zr, which belong to the same column. However, the charge of Pb for PT, which is an important component of good piezoelectrics, was a stand-alone value. Therefore, both charges of Pb and Ti may be magic numbers. In fact, the calculation results for the good piezoelectrics BNT and PMN were close to the proposed numbers. Hence, they may provide us with good guidelines for the design properties of lead-free piezoelectrics that will lead to excellent performance. Through the cluster chemical mean field, the calculation of (Ba, Na)(Ti, Nb)O3 indicates values close to those of PT, but this improvement was limited by the excessively high charge of B. Thus, the O in the model was replaced with the halogen Cl or F, which was more electronegative. The obtained charges were very close to that of PT, and Ba, Na, Ti, O, Cl, or F was necessary for the simplified synthesis of the perovskite substances. This idea does not explain the morphotropic effect, but may support it in terms of obtaining good Pb-free substances from a chemical point of view. This finding may also be very useful for the efficient use of ubiquitous alkaline metals and alkaline-earth metals, especially sodium.

Author Contributions

Conceptualization, N.T.; Methodology, N.T.; Validation, N.T. and J.-i.S.; Investigation, N.T. and J.-i.S.; Writing—original draft preparation, N.T.; Writing—review and editing, N.T. and J.-i.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the research team of science and technology of Fukuoka University.

Acknowledgments

N. T. thanks H. Zhang and S. Park for extending this work to the next fabrication stage. Figure 1 and Figure 2 in this text were partially drawn with graphic software VENUS developed by R. A. Dilanian and F. Izumi.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kuwata, J.; Uchino, K.; Nomura, S. Dielectric and piezoelectric properties of 0.91 Pb (Zn1/3Nb2/3) O3-0.09 PbTiO3 single crystals. Jpn. J. Appl. Phys. 1982, 21, 1298–1302. [Google Scholar] [CrossRef]
  2. Shrout, T.R.; Chang, Z.P.; Kim, N.; Markgraf, S. Dielectric behavior of single crystals near the (1−x) Pb(Mg1/3Nb2/3)O3-(x) PbTiO3 morphotropic phase boundary. Ferroelectr. Lett. 1990, 12, 63–69. [Google Scholar] [CrossRef]
  3. Yamashita, Y.; Harada, K. Crystal growth and electrical properties of lead scandium niobate-lead titanate binary single crystals. Jpn. J. Appl. Phys. 1997, 36, 6039–6042. [Google Scholar] [CrossRef]
  4. Yasuda, N.; Ohwa, H.; Kume, M.; Yamashita, Y. Piezoelectric properties of a high curie temperature Pb(In1/2Nb1/2)O3-PbTiO3 binary system single crystal near a morphotropic phase boundary. Jpn. J. Appl. Phys. 2000, 39, 66–68. [Google Scholar] [CrossRef]
  5. Sui, H.; Sun, H.; Liu, X.; Zhou, D.; Xu, R. Ferroelectric and dielectric behaviors of sol-gel derived perovskite PMN-PT/PZT heterostructures via compositional development: An interface-dependent study. J. Eur. Ceram. Soc. 2018, 38, 5382–5387. [Google Scholar] [CrossRef]
  6. Cole, S.S.; Espenschied, H. Lead titanate: Crystal structure, temperature of formation, and specific gravity data. J. Phys. Chem. 1937, 41, 445–451. [Google Scholar] [CrossRef]
  7. Megaw, H.D. Crystal structure of double oxides of the perovskite type. Proc. Phys. Soc. 1946, 58, 133–152. [Google Scholar] [CrossRef]
  8. Cohen, R.E. Origin of ferroelectricity in perovskite oxides. Nature 1992, 358, 136–138. [Google Scholar] [CrossRef]
  9. Takesue, N.; Ishibashi, K.; Asakura, K. Evaluation of covalency of ions in lead-free perovskite-type dielectric oxides. AIP Adv. 2017, 7, 105016. [Google Scholar] [CrossRef] [Green Version]
  10. Nagata, H.; Shinya, T.; Hiruma, Y.; Takenada, T.; Sakaguchi, I.; Haneda, H. Piezoelectric properties of bismuth sodium titanate ceramics. Ceram. Trans. 2005, 167, 213–221. [Google Scholar]
  11. Fan, P.; Zhang, Y.; Zhang, Q.; Xie, B.; Zhu, Y.; Mawat, M.A.; Ma, W.; Liu, K.; Xiao, J.; Zhang, H. Large strain with low hysteresis in Bi4Ti3O12 modified Bi1/2(Na0.82K0.18)1/2TiO3 lead-free piezoceramics. J. Eur. Ceram. Soc. 2018, 38, 4404–4413. [Google Scholar] [CrossRef]
  12. Kim, M.; Ito, R.; Kim, S.; Khanal, G.P.; Fujii, I.; Suzuki, T.S.; Uchikoshi, T.; Moriyoshi, C.; Kuroiwa, Y.; Wada, S. Fabrication of lead-free piezoelectric (Bi0.5Na0.5)TiO3-BaTiO3 ceramics using electrophoretic deposition. J. Mater. Sci. 2018, 53, 2396–2404. [Google Scholar] [CrossRef]
  13. Zhou, X.; Qi, H.; Yan, Z.; Xue, G.; Luo, H.; Zhang, D. Large energy density with excellent stability in fine-grained (Bi0.5Na0.5)TiO3-based lead-free ceramics. J. Eur. Ceram. Soc. 2019, 39, 4053–4059. [Google Scholar] [CrossRef]
  14. Saito, Y.; Takao, H.; Tani, T.; Nonoyama, T.; Takatori, K.; Homma, T.; Nagaya, T.; Nakamura, M. Lead-free piezoceramics. Nature 2004, 432, 84–87. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, W.; Ren, X. Large piezoelectric effect in Pb-free ceramics. Phys. Rev. Lett. 2009, 103, 257602. [Google Scholar] [CrossRef] [Green Version]
  16. Zhang, H.; Zhu, Y.; Fan, P.; Marwat, M.A.; Ma, W.; Liu, K.; Liu, H.; Xie, B.; Wang, K.; Koruza, J. Temperature-insensitive electric-field-induced strain and enhanced piezoelectric properties of <001> textured (K,Na)NbO3-based lead-free piezoceramics. Acta Mater. 2018, 156, 389–398. [Google Scholar] [CrossRef]
  17. Megaw, H.D. Ferroelectricity in Crystals; Methuen & Co. Ltd.: London, UK, 1957. [Google Scholar]
  18. Kujirai, T.; Takagi, Y.; Nagata, H.; Takenaka, T. Low sintering temperature and electrial properties on cu-doped (Bi1/2Na1/2)TiO3-based solid solution ceramics. In Proceedings of the IEEE International Symposium on Applications of Ferroelectrics (ISAF) 2019, Lausanne, Switzerland, 14–19 July 2019. [Google Scholar] [CrossRef]
  19. Tuluk, A.Y.; Mahon, T.R.; Van Der Zwaag, S.; Groen, P. BiFeO3 synthesis by conventional solid-state reaction. In Proceedings of the IEEE International Symposium on Applications of Ferroelectrics (ISAF) 2019, Lausanne, Switzerland, 14–19 July 2019. [Google Scholar] [CrossRef]
  20. Tachiyama, K.; Yasui, S.; Rao, B.N.; Dazai, T.; Usami, T.; Taniyama, T.; Katayama, T.; Hamasaki, Y.; Yu, J.; He, H.; et al. Magnetic properties of single crystal GaFeO3. MRS Adv. 2019, 4, 61–66. [Google Scholar] [CrossRef]
  21. Hill, N.A. Why are there so few magnetic ferroelectrics? J. Phys. Chem. 2000, 104, 6694–6709. [Google Scholar] [CrossRef]
  22. Goldschmidt, V.M. Die gesetze der krystallochemie. Naturwissenschaften 1926, 21, 477–485. [Google Scholar] [CrossRef]
  23. Bartel, C.J.; Sutton, C.; Goldsmith, B.R.; Ouyang, R.; Musgrave, C.B.; Ghiringhelli, L.M.; Scheffler, M. New tolerance factor to predict the stability of perovskite oxides and halides. Sci. Adv. 2019, 5, 1–9. [Google Scholar] [CrossRef] [Green Version]
  24. Slater, J.C. Quantum, Theory of Molecules and Solids; McGraw Hill: New York, NY, USA, 1974; Volume 4. [Google Scholar]
  25. Adachi, H.; Tsukada, M.; Satoko, C. Discrete variational Xα cluster calculations. I. Application to metal clusters. J. Phys. Soc. Jpn. 1978, 45, 875–883. [Google Scholar] [CrossRef]
  26. Editorial Commitees of Encyclopedic Dictionary of Chemistry. Encyclopedic Dictionary of Chemistry Maruzen eBook Library; Kyoritsu Shuppan Co., Ltd.: Tokyo, Japan, 1963. [Google Scholar]
  27. Mulliken, R.S. Electronic population analysis on LCAO-MO molecular wave functions. I. J. Chem. Phys. 1955, 23, 1833–1840. [Google Scholar] [CrossRef] [Green Version]
  28. Csizmadia, I.G. Theory and Practice of MO Calculations on Organic Molecules; Elsevier: Amsterdam, The Netherlands, 1976. [Google Scholar]
  29. Rabe, K.; Ahn, C.H.; Tricsone, J.M. Physics of Ferroelectrics-Modern Perspective; Springer: Berlin/Heidelberg, Germany, 2007. [Google Scholar]
  30. Cowley, R.A.; Gvasaliya, S.N.; Lushnikov, S.; Roessli, B.; Rotaru, G.-M. Relaxing with relaxors: A review of relaxor ferroelectrics. Adv. Phys. 2011, 60, 229–327. [Google Scholar] [CrossRef]
  31. Kuwata, J.; Uchino, K.; Nomura, S. Phase transitions in the Pb (Zn1/3Nb2/3)O3-PbTiO3 system. Ferroelectrics 1981, 37, 579–582. [Google Scholar] [CrossRef]
  32. Rouquette, J.; Haines, J.; Bornand, V.; Pintard, M.; Papet, P.; Bousquet, C.; Konczewicz, L.; Gorelli, F.A.; Hull, S. Pressure tuning of the morphotropic phase boundary in piezoelectric lead zirconate titanate. Phys. Rev. 2004, 70, 014108. [Google Scholar] [CrossRef]
  33. König, J.; Spreitzer, M.; Jancar, B.; Suvorov, D.; Samardzija, Z.; Popović, A. The thermal decomposition of K0.5Bi0.5TiO3 ceramics. J. Eur. Ceram. Soc. 2009, 29, 1695–1701. [Google Scholar] [CrossRef]
  34. Kainz, T.; Naderer, M.; Schutz, D.; Fruhwirth, O.; Mautner, F.; Reichmann, K. Solid state synthesis and sintering of solid solutions of BNT–xBKT. J. Eur. Ceram. Soc. 2014, 34, 3685–3697. [Google Scholar] [CrossRef]
  35. Taghaddos, E.; Charalambous, H.; Tsakalakos, T.; Safari, A. Electromechanical properties of flash sintered BNT-based piezoelectric ceramic. J. Eur. Ceram. Soc. 2019, 39, 2882–2888. [Google Scholar] [CrossRef]
Figure 1. Simple perovskite structure expressed as ABO3.
Figure 1. Simple perovskite structure expressed as ABO3.
Crystals 10 00956 g001
Figure 2. Cluster models for the calculation. (a) is an antifluorite structure (MA)2O with oxygens colored red. (b) is a sodium chloride structure (MAE)O with MAE body-centered; MAE and O are each colored green and blue, and exchanging the colors provides the oxygen body-centered structure. (c) is a double perovskite structure. The green spheres and the blue spheres are MA, MAE, Pb, Na, or Ba. The gold spheres and purple spheres are Ti, Nb, or Zr. The black spheres are O, Cl, F, or N. The red spheres are oxygens. Chemical species are located so that symmetry D2 is maintained. The colors of ions in each model are summarized in Table 1.
Figure 2. Cluster models for the calculation. (a) is an antifluorite structure (MA)2O with oxygens colored red. (b) is a sodium chloride structure (MAE)O with MAE body-centered; MAE and O are each colored green and blue, and exchanging the colors provides the oxygen body-centered structure. (c) is a double perovskite structure. The green spheres and the blue spheres are MA, MAE, Pb, Na, or Ba. The gold spheres and purple spheres are Ti, Nb, or Zr. The black spheres are O, Cl, F, or N. The red spheres are oxygens. Chemical species are located so that symmetry D2 is maintained. The colors of ions in each model are summarized in Table 1.
Crystals 10 00956 g002
Figure 3. Histograms of the effective charges of the cations of (a) (MA)2O and (b) (MAE)O. The values in Table 3 and Table 4 were used.
Figure 3. Histograms of the effective charges of the cations of (a) (MA)2O and (b) (MAE)O. The values in Table 3 and Table 4 were used.
Crystals 10 00956 g003
Figure 4. Histograms of the effective charges of the cations of double simple perovskites ABO3. The values in Table 5 were used.
Figure 4. Histograms of the effective charges of the cations of double simple perovskites ABO3. The values in Table 5 were used.
Crystals 10 00956 g004
Figure 5. Case similar to 3d0 of Ti4+: (a) histogram of the effective charges of the cations of strained and unstrained PT, (b) physical and chemical image of energy gain and loss with respect to strain. (a) was drawn by using the values in Table 6. r, rc, re in (b) are the distance between the two ions in general meaning, the distance in the compressed state, that in the expanded state, respectively.
Figure 5. Case similar to 3d0 of Ti4+: (a) histogram of the effective charges of the cations of strained and unstrained PT, (b) physical and chemical image of energy gain and loss with respect to strain. (a) was drawn by using the values in Table 6. r, rc, re in (b) are the distance between the two ions in general meaning, the distance in the compressed state, that in the expanded state, respectively.
Crystals 10 00956 g005
Figure 6. Histograms of the effective charges of the cations of the solid solution model (Ba, Na)(Ti, Nb)O3. A corresponds to (Ba, Na), and B (Ti, Nb). The values in Table 7 were used.
Figure 6. Histograms of the effective charges of the cations of the solid solution model (Ba, Na)(Ti, Nb)O3. A corresponds to (Ba, Na), and B (Ti, Nb). The values in Table 7 were used.
Crystals 10 00956 g006
Figure 8. Histogram of the effective charges of A and B of cluster model (Ba, Na)(Ti, Nb)O3 replaced with Cl, F, N, respectively. The values in Table 8 were used. The values of PT in Table 5 and (Ba, Na)(Ti, Nb)O3 in Table 7 were also used for comparison.
Figure 8. Histogram of the effective charges of A and B of cluster model (Ba, Na)(Ti, Nb)O3 replaced with Cl, F, N, respectively. The values in Table 8 were used. The values of PT in Table 5 and (Ba, Na)(Ti, Nb)O3 in Table 7 were also used for comparison.
Crystals 10 00956 g008
Figure 9. Map of the values in Table 8. The dots of PT and (Ba, Na)(Ti, Nb)O3 in Figure 7 are given for comparison.
Figure 9. Map of the values in Table 8. The dots of PT and (Ba, Na)(Ti, Nb)O3 in Figure 7 are given for comparison.
Crystals 10 00956 g009
Table 1. Colors of the ions in each perovskite model.
Table 1. Colors of the ions in each perovskite model.
ModelIonColor
PTPbgreen, blue
Tigold, purple
Ored, black
MANbO3MAgreen, blue
Nbgold, purple
Ored, black
MAETiO3MAEgreen, blue
Tigold, purple
Ored, black
BaZrO3Bagreen, blue
Zrgold, purple
Ored, black
(Ba, Na)(Ti, Nb)O3
Ba-centered
Bagreen
Nablue
Tigold
Nbpurple
Ored, black
(Ba, Na)(Ti, Nb)O3
Na-centered
Nagreen
Bablue
Tigold
Nbpurple
Ored, black
(Ba, Na)Ti(O, Cl)3
Ba-centered
Bagreen
Nablue
Tigold, purple
Ored
Clblack
(Ba, Na)Ti(O, Cl)3
Na-centered
Nagreen
Bablue
Tigold, purple
Ored
Clblack
Na(Ti, Nb)(O, Cl)3Nagreen, blue
Tigold
Nbpurple
Ored
Clblack
(Ba, Na)Ti(O, F)3
Ba-centered
Bagreen
Nablue
Tigold, purple
Ored
Fblack
(Ba, Na)Ti(O, F)3
Na-centered
Nagreen
Bablue
Tigold, purple
Ored
Fblack
Na(Ti, Nb)(O, F)3Nagreen, blue
Tigold
Nbpurple
Ored
Fblack
(Ba, Na)Nb(O, N)3
Ba-centered
Bagreen
Nablue
Nbgold, purple
Ored
Nblack
(Ba, Na)Nb(O, N)3
Na-centered
Nagreen
Bablue
Nbgold, purple
Ored
Nblack
Ba(Ti, Nb)(O, N)3Bagreen, blue
Tigold
Nbpurple
Ored
Nblack
Table 2. Lattice parameters used for modeling.
Table 2. Lattice parameters used for modeling.
ModelCrystal SystemLattice Parameter (nm)
Li2Ocubica = 0.461
Na2Ocubica = 0.555
K2Ocubica = 0.644
MgOcubica = 0.421
CaOcubica = 0.480
SrOcubica = 0.5144
BaOcubica = 0.55391
Double perovskitesa, b, c orthogonal, D2a = b = c = 0.4 × 2 = 0.8
Table 3. Effective charges of MA of reverse fluoride structures.
Table 3. Effective charges of MA of reverse fluoride structures.
ModelEffective Charge of MA
Li2O0.601
Na2O0.409
K2O0.561
Table 4. Effective charge of MAE of the sodium chloride structures.
Table 4. Effective charge of MAE of the sodium chloride structures.
ModelEffective Charge of MAE
MgO Mg-centered1.325
MgO O-centered1.295
CaO Ca-centered1.597
CaO O-centered1.584
SrO Sr-centered1.791
SrO O-centered1.762
BaO Ba-centered1.824
BaO O-centered1.800
Table 5. Effective charges of the cations of the ABO3 double perovskites.
Table 5. Effective charges of the cations of the ABO3 double perovskites.
ModelMA/MAE/PbEffective Charge of ATi/Zr/NbEffective Charge of B
LiNbO3MA0.703Nb3.020
NaNbO3MA0.608Nb3.007
KNbO3MA0.791Nb3.032
CaTiO3MAE1.581Ti2.451
SrTiO3MAE1.838Ti2.480
BaTiO3MAE1.900Ti2.507
BaZrO3MAE1.910Zr2.823
PbTiO3Pb1.302Ti2.369
Table 6. Effective charges of PT (PbTiO3) for each ratio of the isotropic deformation.
Table 6. Effective charges of PT (PbTiO3) for each ratio of the isotropic deformation.
ModelPbTi
PbTiO3 95%1.3812.584
PbTiO3 97.5%1.3372.471
PbTiO3 100%1.3022.369
PbTiO3 102.5%1.2752.274
PbTiO3 105%1.2542.188
Table 7. Effective charges of the solid solution models (Ba, Na)(Ti, Nb)O3. Each of the detailed chemical notations of the models is given in parenthesis in each cell on the left.
Table 7. Effective charges of the solid solution models (Ba, Na)(Ti, Nb)O3. Each of the detailed chemical notations of the models is given in parenthesis in each cell on the left.
ModelBa/NaA AveragedTi/NbB Averaged
(Ba, Na)(Ti, Nb)O3
((Ba1/2Na1/2)(Ti1/2Nb1/2)O3)
Ba-centered
Ba 1.900
Na 0.659
1.279Ti 2.592
Nb 2.832
2.712
(Ba, Na)(Ti, Nb)O3
((Ba1/2Na1/2)(Ti1/2Nb1/2)O3)
Na-centered
Ba 1.914
Na 0.607
1.261Ti 2.583
Nb 2.841
2.712
Table 8. Effective charges of solid solution models (Ba, Na)(Ti, Nb)O3; O ions were partially replaced with halogen Cl or F, or not halogen N. Each of the detailed chemical notations of the models is given in parenthesis in each cell on the left.
Table 8. Effective charges of solid solution models (Ba, Na)(Ti, Nb)O3; O ions were partially replaced with halogen Cl or F, or not halogen N. Each of the detailed chemical notations of the models is given in parenthesis in each cell on the left.
ModelAA AveragedBB Averaged
(Ba, Na)Ti(O, Cl)3
((Ba4Na4)Ti8(O20Cl4))
Ba-centered
Ba 1.920
Na 0.500
1.210Ti 2.3442.344
(Ba, Na)Ti(O, Cl)3
((Ba4Na4)Ti8O20Cl4)
Na-centered
Ba 1.873
Na 0.459
1.166Ti 2.3552.355
Na(Ti, Nb)(O, Cl)3
(Na8(Ti4Nb4)(O20Cl4))
Na 0.4450.445Ti 2.434
Nb 2.611
2.523
(Ba, Na)Ti(O, F)3
((Ba4Na4)Ti8(O20F4))
Ba-centered
Ba 1.865
Na 0.523
1.194Ti 2.4222.422
(Ba, Na)Ti(O, F)3
((Ba4Na4)Ti8(O20F4))
Na-centered
Ba 1.856
Na 0.528
1.192Ti 2.4232.423
Na(Ti, Nb)(O, F)3
(Na8(Ti4Nb4)(O20F4))
Na 0.5450.545Ti 2.542
Nb 2.761
2.651
(Ba, Na)Nb(O, N)3
((Ba4Na4)Nb8(O20N4))
Ba-centered
Ba 1.912
Na 0.724
1.318Nb 2.9832.983
(Ba, Na)Nb(O, N)3
((Ba4Na4)Nb8(O20N4))
Na-centered
Ba 1.943
Na 0.609
1.276Nb 2.9832.983
Ba(Ti, Nb)(O, N)3
(Ba8(Ti4Nb4)(O20N4))
Ba 1.8991.899Ti 2.589
Nb 2.817
2.703
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Takesue, N.; Saito, J.-i. Molecular Orbital Calculation of Lead-Free Perovskite Compounds for Efficient Use of Alkaline and Alkaline Earth Metals. Crystals 2020, 10, 956. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst10110956

AMA Style

Takesue N, Saito J-i. Molecular Orbital Calculation of Lead-Free Perovskite Compounds for Efficient Use of Alkaline and Alkaline Earth Metals. Crystals. 2020; 10(11):956. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst10110956

Chicago/Turabian Style

Takesue, Naohisa, and Jun-ichi Saito. 2020. "Molecular Orbital Calculation of Lead-Free Perovskite Compounds for Efficient Use of Alkaline and Alkaline Earth Metals" Crystals 10, no. 11: 956. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst10110956

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

Article Metrics

Back to TopTop