Defect Clustering in Rare-Earth-Doped BaTiO3 and SrTiO3 and Its

Oct 4, 2017 - All calculations were carried out using the general utility lattice program (GULP).(47). The binding energy (Ebind) between oppositely c...
0 downloads 9 Views 1MB Size
Subscriber access provided by TUFTS UNIV

Article 3

Defect Clustering in Rare-Earth-Doped BaTiO and SrTiO and Its Influence on Dopant Incorporation 3

Yohandys A. Zulueta, Teik-Cheng Lim, and James A. Dawson J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b08500 • Publication Date (Web): 04 Oct 2017 Downloaded from http://pubs.acs.org on October 5, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

Defect Clustering in Rare-Earth-Doped BaTiO3 and SrTiO3 and its Influence on Dopant Incorporation Y. A. Zulueta1,2, T.C. Lim3 and J. A. Dawson*4 1

Departamento de Física, Facultad de Ciencias Naturales, Universidad de Oriente, CP- 90500, Santiago de Cuba, Cuba

2

Department of Chemistry, KU Leuven, B-3001 Leuven, Belgium

3

School of Science and Technology, Singapore University of Social Sciences, Singapore

4

Department of Chemistry, University of Bath, Bath BA2 7AX, United Kingdom

1 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

*Corresponding author. E-mail: [email protected]

ABSTRACT Rare-earth (RE) doping of BaTiO3 and SrTiO3 has received tremendous research attention in recent decades. Although RE doping has been a particularly popular topic, little or no attention has been given to the contribution from different defect arrangements. In this study, we use a proven interatomic potential model to calculate the binding energies of different RE donor defect (Ba/Sr-site doping) configurations in BaTiO3 and SrTiO3. We consider two standard donor defect mechanisms, namely, charge compensation from cation vacancies. Consistently stronger negative binding energies are calculated for SrTiO3 compared to BaTiO3. We also show that SrTiO3 can accept a larger range of RE ion sizes at the divalent cation site than BaTiO3. This is highlighted by the simple trend of increasing binding energy with decreasing RE ion size found for the defect configurations with a Ti vacancy in SrTiO3. Conversely, defect clustering in BaTiO3 can cause significant strain on the lattice, resulting in positive binding energies. Our results show that one standard defect arrangement per incorporation mechanism cannot be applied to all dopants. We clearly show that the lowest energy defect configuration is dependent on both the dopant and host material. The consequences of these findings for the overall defect incorporation energetics in BaTiO3 and SrTiO3 are also presented.

2 Environment ACS Paragon Plus

Page 2 of 24

Page 3 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

1. Introduction The ABO3 perovskite structure endows BaTiO3 and SrTiO3 with a wide selection of useful electrical and structural properties. These powerful properties have promoted the use of BaTiO3 and SrTiO3 ceramics in a range of technological applications, including multilayer ceramic capacitors1,2, random access memories3,4, positive temperature coefficient of resistance thermistors5,6, thermoelectrics7,8 and microwave devices9,10. These two materials are generally considered as cornerstones of the electroceramics industry. The incorporation of dopant ions, in a wide range of concentrations, into the ABO3 perovskite structure is essential for refining the properties and applications of these materials. In particular, trivalent rare-earth (RE) ions have received prodigious attention because of their ability to tailor the electrical properties of BaTiO311–13 and the thermoelectric properties of SrTiO37,8,14. However, RE elements are expensive and potential methods to reduce their usage or use them more efficiently are therefore highly desired. There are a variety of RE incorporation mechanisms that can occur in these materials, depending on the RE3+ ion size and the A-/B-site ratio. For large RE ions, like La, doping is known to occur at the A-sites of the perovskite structure with charge compensation in oxidizing conditions coming from Sr or Ti vacancies for SrTiO3 and BaTiO3, respectively, or from reduction of Ti4+ to Ti3+ in reducing conditions. This has been confirmed both experimentally12,15–17 and computationally18–25. Mid-sized ions, such as Eu, Gd and Dy, whose ionic radii lie in-between the radii of the A-site and Ti ions, are reported to dope at the A-sites, Ti-sites or both sites

3 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

simultaneously (self-compensation mechanism) in BaTiO3 and SrTiO3, depending on the thermodynamic conditions7,18–21,23,26–30. In BaTiO3, smaller RE ions, like Yb and Lu, dope exclusively at the Ti-site with oxygen vacancy charge compensation11,21,31–33. For SrTiO3, there are very few reports on Yb and Lu doping. One experimental study suggests that Sr-site incorporation is possible for Yb-doping34. Computationally, self-compensation is shown to be energetically preferred23,29,35. Similar to the mid-sized RE ions, it is highly likely that the smaller RE ions can dope at both the Sr- and Ti-sites, as well via selfcompensation, in SrTiO3. When calculating the energetics of dopant incorporation or defect formation for a material, it is essential to consider any potential cohesion or repulsion between the dopants and any resulting charge compensating defects. This is true regardless of whether the defect energetics are calculated by potential-based methods or density functional theory (DFT). Various computational studies have shown the importance of these binding interactions in a variety of perovskite materials.21–25,36–41 The magnitudes of the binding energies in BaTiO3 and SrTiO3 have been shown to be large and have a drastic effect on the overall lowest energy incorporation mechanism predicted21,23. This means that it is simply not enough to only consider isolated dopants in such calculations. However, this is often the approach used in the literature for most materials. Furthermore, when binding interactions are calculated, one low energy defect configuration is usually applied to all dopant ions, with little or no consideration of the binding between other potential configurations. This is unlikely to be an issue for simple defect pairs, such as the binding between one RE ion at an

4 Environment ACS Paragon Plus

Page 4 of 24

Page 5 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

A-site and one RE at a neighboring Ti-site (self-compensation), but for complex defect clusters, the binding energy will vary significantly for different configurations. In this study, we investigate the magnitude and variation of defect binding energies in numerous defect arrangements for two fundamental RE dopant incorporation mechanisms, namely, the donor-doping of RE ions into cubic BaTiO3 and SrTiO3 with charge compensation from either Ti or Ba/Sr vacancies:

• //// 2 RE2O3 + 4 BaBa / SrSr + TiTi → 4 REBa / Sr + VTi + 3BaO / SrO + BaTiO3 / SrTiO3

(1)

Scheme 1: Incorporation of RE3+ at Ba/Sr2+ with Ti4+ vacancy compensation.

• // RE2O3 + 3BaBa / SrSr → 2 REBa / Sr + VBa / Sr + 3BaO / SrO

Scheme 2: Incorporation of RE3+ at Ba/Sr2+ with Ba/Sr2+ vacancy compensation.

In scheme 1, four RE3+ ions dope at the A-sites, which creates a net charge of 4+ that is compensated for by one Ti vacancy. In scheme 2, two RE3+ ions dope at the A-sites and the total 2+ charge is compensated for by a single A-site vacancy. As discussed, schemes 1 and 2 are the dominant dopant incorporation mechanisms for large RE ions in BaTiO3 and SrTiO3, respectively. While we have only considered the two donor doping mechanisms in this study because of the wide range of complex defect clusters available for these mechanisms, the methods used and results found are certainly applicable to acceptor doping and self-compensation mechanisms.

2. Methodology The interatomic potential-based simulations performed in this work have been extensively used to calculate the defect properties and energetics of perovskite 5 Environment ACS Paragon Plus

(2)

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

materials and are presented in detail elsewhere42,43. In these calculations, the total energy of a system is made up of contributions from the long-range and short-range interactions with respect to the atomic positions in the lattice. The long-range forces are accounted for by Coulombic interactions and the short-range forces are represented by interatomic potentials that include electron repulsion and van der Waals interactions. In this work, we use the well-established potential model of Freeman et al.21 and Dawson et al.23, which have proven to be effective in modelling a wide variety of defect phenomena in perovskites14,29,30,44. The potential model was originally fitted for BaTiO321 using DFT calculations and was later extended to SrTiO3 and CaTiO323. A full description of the potential fitting procedure is available in Ref. 21. This model replicates both the experimental cell parameters and lattice energies of BaTiO3, SrTiO3 and their respective end-member oxides with precision. It is crucial that the perfect bulk structures are accurately reproduced before attempting to calculate the defect properties. A cutoff of 12 Å was used for all the interatomic potentials. The RE potentials were taken from Ref. 25. The ionic polarization in our simulations is accounted for by the shell model of Dick and Overhauser45. This model divides atoms into negatively charged massless shells and positively charged cores. The electrostatic interactions between the cores and shells are replaced by a harmonic term that is determined by the spring constant, k. The defect simulations were completed using the Mott-Littleton method46, whereby the crystal lattice is partitioned into two regions, an inner region and an outer region. Interactions are calculated explicitly in the inner region and are approximated using a dielectric continuum method in the outer region. All calculations were carried out using the general utility lattice program (GULP)47.

6 Environment ACS Paragon Plus

Page 6 of 24

Page 7 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

The binding energy (Ebind) between oppositely charged defects is defined as the difference between the total energy of the isolated defects and the energy when the same defects are simulated together in a cluster: Ebind = E(X) + E(Y) – E(XY)

(3)

where a negative value implies a stable cluster and binding behavior.

We note that DFT could also be used to calculate the binding energies of defect clusters in these materials. However, given the significantly high number of calculations involved in this work, as a result of the large number of RE ions and defect configurations considered, the computational expense would be a limiting factor. Therefore, a computationally more efficient approach is necessary.

3. Results and discussion 3.1. Perovskite structure and defect configurations In the conventional ABX3 perovskite structure (see Figure 1(a)), the A and B sites are occupied by cations and the X site is an anion. In BaTiO3 and SrTiO3, the B-site Ti4+ ion is octahedrally coordinated to six O2- ions, while the A-site Ba2+ or Sr2+ ion is cubo-octahedrally coordinated to twelve O2- ions. The calculated lattice parameters for cubic BaTiO3 and SrTiO3 are 4.01 and 3.96 Å, respectively, in excellent agreement with X-ray diffraction values of 4.0048 and 3.91 Å49, respectively. We consider six defect configurations for scheme 1 and five for scheme 2, as illustrated in Figures 2 and 3, respectively. These configurations account for all possible defect arrangements of schemes 1 and 2 on the basis of a single perovskite unit cell. Only a single perovskite cell is considered since the interactions between oppositely charged neighboring defects will clearly be stronger than those between

7 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

defects with larger interatomic separations. This is in addition to the fact that the number of defect arrangements grows exponentially with increasing unit cells, thereby making the number of calculations impractical. A more detailed description of the six scheme 1 arrangements is given by Gröting et al.50, who used them to probe the chemical order and local structure in Na0.5Bi0.5TiO3. For scheme 2, the defects are always in a particular plane to produce layered structures. For configurations 1 and 2, the defects are ordered in the (001) plane. In configurations 3 and 4, there is defect ordering in the (110) plane, while in configuration 5, (111) defect ordering is present.

Fig. 1. (a) Cubic perovskite structure for BaTiO3 and SrTiO3. Blue spheres represent Ba/Sr ions, the TiO6 octahedra are given in grey and red spheres are oxygen ions.

8 Environment ACS Paragon Plus

Page 8 of 24

Page 9 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

1

2

3

aaaaaaaaa 4

5

6

Fig. 2. (a) Binding energy defect configurations considered for scheme 1. Purple spheres represent RE dopants and the Ti vacancy is indicated by the white sphere at the center of the unit cells.

9 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

1

2

4

Page 10 of 24

3

5

Fig. 3. Binding energy defect configurations considered for scheme 2. The Ba/Sr vacancy is indicated by the white sphere at the corner of the unit cells.

3.2. Binding energies

The calculated binding energies for scheme 1 in BaTiO3 and SrTiO3 are presented using heat maps in Figure 4. The first result to note is that the majority of binding energies are strongly negative. In the case of SrTiO3, all values are negative and in excess of -2 eV, thereby highlighting the significantly favorable defect interactions in these clusters. For BaTiO3, all of the calculated binding energies are weaker than the respective values for SrTiO3. This can be explained by simple Coulombic arguments, since the SrTiO3 lattice parameter of 3.96 Å is smaller than that for BaTiO3 (4.01 Å), resulting in a stronger electrostatic attraction between the positively charged RE ions

10 Environment ACS Paragon Plus

Page 11 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

and the negatively charged vacancy. Similar findings have been reported previously21,23, with even more substantial energies reported for CaTiO3. (a)

(b)

Fig. 4. Binding energies (eV) for each scheme 1 defect configuration in (a) BaTiO3 and SrTiO3.

Figure 4(b) shows a strong trend of increasing binding energy with decreasing RE ion size for all configurations in SrTiO3, as discussed previously23, with relatively little variation between the different defect configurations. However, this is not the case for BaTiO3. Although configurations 1 and 3 show a clear trend with ionic radii, the other configurations do not. For example, while the energies for La in the six configurations are similar, the variation for smaller RE ions is much greater, with many positive binding energies produced. This can perhaps be expected given that such mid- and small-sized RE ions cannot dope solely at the Ba-site in BaTiO3. These results suggest that SrTiO3 is capable of accommodating a greater range of RE ion sizes at the Sr site than BaTiO3 can for the Ba site. Our findings also indicate that the extent to which different defect configurations influence the binding energy, and

11 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

therefore the overall defect incorporation energetics, is dependent on both the host material and the dopant ion. Previous calculations have reported that configuration 1 is the lowest in energy for scheme 1 in both BaTiO321 and SrTiO323 as a result of the large dipole formed from the four RE ions being in the same plane. This is in agreement with our results for all dopants in BaTiO3 with the exception of La and Pr, where configurations 2 and 5 are the lowest in energy, respectively. Conversely, for SrTiO3, while configuration 1 is the most favorable for La doping, other configurations are preferred for smaller RE ions, particularly configuration 5. This highlights the importance of considering different binding configurations for different dopants as they can have a significant effect on the overall dopant formation energies. The binding energies of scheme 2 in BaTiO3 and SrTiO3 are displayed in Figure 5. Unlike scheme 1, every calculated binding energy is negative, suggesting that the interaction between two RE dopants and Ba or Sr vacancy is favorable regardless of the dopant, host material or configuration. The binding is SrTiO3 is again stronger than in BaTiO3 as a result of the reduced distance between the oppositely charged defects. As expected, the interactions between these defect clusters are generally weaker given the lower number of defects and reduced charges involved. Similar general trends have been reported previously21,23.

12 Environment ACS Paragon Plus

Page 12 of 24

Page 13 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

(a)

(b)

Fig. 5. Binding energies (eV) for each scheme 2 defect configuration in (a) BaTiO3 and (b) SrTiO3.

By comparing Figures 4 and 5, it is clear that there is no trend between RE ion size and binding energy for the scheme 2 defect clusters. This is particularly true for BaTiO3, where it can be seen from Figure 5(a) that the variation in energy with RE ion is small, with only one color dominating each defect configuration. The large separation between the A sites, which results in the lack of a strong network to link the sites is likely to be the reason for the absence of this trend21. Although there is little binding energy variation between different RE ions, there is significant variation between the different configurations, especially for smaller RE ions, such as Yb and Lu. For example, there is a difference of ~0.6 eV between the most (configuration 1) and least (configuration 2) favorable configurations in BaTiO3. A similar difference also exists between configurations 2 and 4 in SrTiO3. It is perhaps surprising that such variation can exist in a relatively simple defect cluster made up of only three defects ordered in five possible ways. These findings again illustrate the importance of testing different defect clusters.

13 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

3.3. Influence on dopant incorporation

Binding energies are often combined with defect solution or formation energies to produce a more accurate prediction of the total energy required for dopant incorporation. To illustrate the variation in binding energies reported here and how they influence the overall energetics of dopant incorporation, we combine our results with previously calculated solution energies for La-, Nd-, Eu-, Gd-, Ho-, Yb- and Ludoping in BaTiO321 and SrTiO323. The solution energy represents the energy required for a particular dopant to substitute into a host lattice by taking into account any charge compensating defects that may be required. More details are on how these solution energies are calculated are available in Refs. 21 and 23. The solution and total energies for schemes 1 and 2 in BaTiO3 and SrTiO3 are plotted as a function of RE ionic radius in Figure 6. The minimum and maximum total energies represent the solution energies combined with the strongest and weakest binding energies, respectively, obtained for each RE ion. The binding energy applied to the solution energy is divided by the number of dopants involved in the cluster, i.e., the binding energy per dopant.

14 Environment ACS Paragon Plus

Page 14 of 24

Page 15 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

(a)

(b)

Fig. 6. Solution and total energies plotted as a function of RE ionic radius for schemes 1 and 2 in (a) BaTiO3 and (b) SrTiO3. The solution energies for BaTiO3 and SrTiO3 are taken from Refs. 21 and 23, respectively. The minimum and maximum total energies represent the solution energies combined with the strongest and weakest binding energies, respectively, obtained for each RE ion.

15 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The most important finding from Figure 6 is that the lowest energy scheme can vary depending on the choice of binding energy/defect cluster used. It is well known that large RE ions, like La, dope at the Ba site of BaTiO3 with Ti vacancy charge compensation (scheme 1)12,21,51,52. This is confirmed by our calculations regardless of the defect complex considered (Figure 6(a)). However, for other RE ions, the maximum scheme 1 and minimum scheme 2 energies are very similar, meaning that is not now clear what scheme will dominate. This problem is even clearer for SrTiO3 (Figure 6(b)), where it is known that donor doping of large RE ions occurs at the Sr sites with Sr vacancy charge compensation (scheme 2)16,17,23,53. Our calculations also predict this, but only when the strongest binding energies are added to the solution energy. When this is not the case, La is predicted to incorporate via scheme 1. Essentially, our results show that not determining the lowest energy defect cluster for each dopant ion can lead to incorrect predictions of its incorporation scheme. It is important to bear in mind that we have only considered two donor doping schemes in this work. In reality, the prediction of a dopant incorporation scheme in these materials is more complex as it also involves acceptor doping and self-compensation mechanisms. This means that the potential for incorrectly predicting the doping site as well as the compensating defects becomes higher. Many works have shown the importance of binding energies in the calculation of defect formation energies. In this study, we have shown the importance of calculating the lowest possible binding energies from numerous defect cluster configurations. These findings have been presented using interatomic potential-based calculations of perovskites, but they could potentially apply to any defect cluster in any material, regardless of the calculation methods.

16 Environment ACS Paragon Plus

Page 16 of 24

Page 17 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

4. Conclusions Using interatomic potential-based defect calculations, we have explored defect binding in RE-donor-doped BaTiO3 and SrTiO3, and analyzed its overall impact on the energetics of dopant incorporation in these materials. The key findings of this study are as follows: -

Binding between the RE dopants and their charge compensating defects is shown to be generally favorable, particularly for SrTiO3.

-

Our results suggest that SrTiO3 can accept a larger range of RE ion sizes at its A site than BaTiO3.

-

Defect clustering in BaTiO3 can cause significant strain on the lattice, resulting in positive binding energies.

-

One standard defect arrangement per incorporation mechanism cannot be applied to all dopants as the lowest energy defect configuration is dependent on both the dopant and host material.

-

Not considering defect clustering or considering the wrong defect cluster can result in incorrect predictions of the dominant dopant incorporation scheme.

The atomic-scale insights presented in this study have fully illustrated the importance of defect clustering in oxide perovskite materials and how the binding behavior of these clusters can vary dramatically. In addition, the findings presented here are certainly not limited to only these two materials, or indeed only potential-based calculations. We hope that this work will encourage more in-depth studies of the interactions between clustered defects in a wide variety of materials.

ACKNOWLEDGEMENTS

17 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

Page 18 of 24

The authors thank the Japan Society for the Promotion of Science (JSPS) for funding through a JSPS fellowship (Grant No. 2503370). JAD is grateful to Dr. Pedro “Pete” Canepa for insightful discussions regarding data presentation.

REFERENCES (1)

Kishi, H.; Mizuno, Y.; Chazono, H. Base-Metal Electrode-Multilayer Ceramic Capacitors: Past, Present and Future Perspectives. Jpn. J. Appl. Phys. 2003, 42, 1–15.

(2)

Pan, M. J.; Randall, C. A. A Brief Introduction to Ceramic Capacitors. IEEE Electrical Insulation Magazine. 2010, pp. 44–50.

(3)

Scott, J. F.; de Araujo, C. A. Ferroelectric Memories. Science 1989, 246, 1400– 1405.

(4)

Sirenko, A. A.; Bernhard, C.; Golnik, A.; Clark, A. M.; Hao, J.; Si, W.; Xi, X. X. Soft-Mode Hardening in SrTiO3 Thin Films. Nature 2000, 404, 373–376.

(5)

Sinclair, D. C.; West, A. R. Effect of Atmosphere on the PTCR Properties of BaTiO3 Ceramics. J. Mater. Sci. 1994, 29, 6061–6068.

(6)

Huybrechts, B.; Ishizaki, K.; Takata, M. The Positive Temperature Coefficient of Resistivity in Barium Titanate. J. Mater. Sci. 1995, 30, 2463–2474.

(7)

Muta, H.; Kurosaki, K.; Yamanaka, S. Thermoelectric Properties of Rare Earth Doped SrTiO3. J. Alloys Compd. 2003, 350, 292–295.

(8)

Ohta, S.; Nomura, T.; Ohta, H.; Koumoto, K. High-Temperature Carrier Transport and Thermoelectric Properties of Heavily La- or Nb-Doped SrTiO3 Single Crystals. J. Appl. Phys. 2005, 97, 34106.

(9)

Tagantsev, A. K.; Sherman, V. O.; Astafiev, K. F.; Venkatesh, J.; Setter, N. Ferroelectric

Materials

for

Microwave

Tunable

18 Environment ACS Paragon Plus

Applications.

J.

Page 19 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

Electroceramics 2003, 11, 5–66. (10)

Haeni, J. H.; Irvin, P.; Chang, W.; Uecker, R.; Reiche, P.; Li, Y. L.; Choudhury, S.; Tian, W.; Hawley, M. E.; Craigo, B.; et al. Room-Temperature Ferroelectricity in Strained SrTiO3. Nature 2004, 430, 758–761.

(11)

Jung, Y.-S.; Na, E.-S.; Paik, U.; Lee, J.; Kim, J. A Study on the Phase Transition and Characteristics of Rare Earth Elements Doped BaTiO3. Mater. Res. Bull. 2002, 37, 1633–1640.

(12)

Morrison, F. D.; Sinclair, D. C.; West, A. R. Electrical and Structural Characteristics of Lanthanum-Doped Barium Titanate Ceramics. J. Appl. Phys. 1999, 86, 6355–6366.

(13)

Ismail, F. A.; Osman, R. A. M.; Idris, M. S. Review on Dielectric Properties of Rare Earth Doped Barium Titanate. AIP Conf. Proc. 2016, 1756, 90005.

(14)

Dawson, J. A.; Tanaka, I. Local Structure and Energetics of Pr-and La-Doped SrTiO3 Grain Boundaries and the Influence on Core–Shell Structure Formation. J. Phys. Chem. C 2014, 118, 25765–25778.

(15)

Morrison, F. D.; Sinclair, D. C.; West, A. R. An Alternative Explanation for the Origin of the Resistivity Anomaly in La-Doped BaTiO3. J. Am. Ceram. Soc. 2001, 84, 474–476.

(16)

Yu, Z.; Ang, C. Dielectric and Conduction Behavior of La-Doped SrTiO3 with Suppressed Quantum-Paraelectric Background. Appl. Phys. Lett. 2002, 80, 643–645.

(17)

Savaniu, C.-D.; Irvine, J. T. S. Reduction Studies and Evaluation of Surface Modified A-Site Deficient La-Doped SrTiO3 as Anode Material for IT-SOFCs. J. Mater. Chem. 2009, 19, 8119–8128.

(18)

Lewis, G. V.; Catlow, C. R. A.; Casselton, R. E. W. PTCR Effect in BaTiO3. J.

19 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

Am. Ceram. Soc. 1985, 68, 555–558. (19)

Buscaglia, M. T.; Buscaglia, V.; Viviani, M.; Nanni, P. Atomistic Simulation of Dopant Incorporation in Barium Titanate. J. Am. Ceram. Soc. 2001, 84, 376–384.

(20)

Moriwake, H.; Fisher, C. A. J.; Kuwabara, A. First-Principles Calculations of Rare-Earth Dopants in BaTiO3. Jpn. J. Appl. Phys. 2009, 48, 09KC03.

(21)

Freeman, C. L.; Dawson, J. A.; Chen, H.-R.; Harding, J. H.; Ben, L.-B.; Sinclair, D. C. A New Potential Model for Barium Titanate and its Implications for Rare-Earth Doping. J. Mater. Chem. 2011, 21, 4861–4868.

(22)

Dawson, J. A.; Harding, J. H.; Chen, H.; Sinclair, D. C. First-Principles Study of Intrinsic Point Defects in Hexagonal Barium Titanate. J. Appl. Phys. 2012, 111, 94108.

(23)

Dawson, J. A.; Li, X.; Freeman, C. L.; Harding, J. H.; Sinclair, D. C. The Application of a New Potential Model to the Rare-Earth Doping of SrTiO3 and CaTiO3. J. Mater. Chem. C 2013, 1, 1574–1582.

(24)

Freeman, C. L.; Dawson, J. A.; Chen, H.; Ben, L.; Harding, J. H.; Morrison, F. D.; Sinclair, D. C.; West, A. R. Energetics of Donor‐Doping, Metal Vacancies, and Oxygen‐Loss in A‐Site Rare‐Earth‐Doped BaTiO3. Adv. Funct. Mater. 2013, 23, 3925–3928.

(25)

Lewis, G. V; Catlow, C. R. A. Defect Studies of Doped and Undoped Barium Titanate Using Computer Simulation Techniques. J. Phys. Chem. Solids 1986, 47, 89–97.

(26)

Makovec, D.; Samardžija, Z.; Drofenik, M. Solid Solubility of Holmium, Yttrium, and Dysprosium in BaTiO3. J. Am. Ceram. Soc. 2004, 87, 1324–1329.

(27)

Mizuno, Y.; Kishi, H.; Ohnuma, K.; Ishikawa, T.; Ohsato, H. Effect of Site

20 Environment ACS Paragon Plus

Page 20 of 24

Page 21 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

Occupancies of Rare Earth Ions on Electrical Properties in Ni-MLCC Based on BaTiO3. J. Eur. Ceram. Soc. 2007, 27, 4017–4020. (28)

Jiang, C.; Fang, L.; Shen, M.; Zheng, F.; Wu, X. Effects of Eu Substituting Positions and Concentrations on Luminescent, Dielectric, and Magnetic Properties of SrTiO3 Ceramics. Appl. Phys. Lett. 2009, 94, 71110.

(29)

Zulueta, Y. A.; Dawson, J. A.; Mune, P. D.; Froeyen, M.; Nguyen, M. T. Oxygen Vacancy Generation in Rare‐Earth‐Doped SrTiO3. Phys. Status Solidi 2016, 253, 2197–2203.

(30)

Freeman, C. L.; Dawson, J. A.; Harding, J. H.; Ben, L.; Sinclair, D. C. The Influence of A‐Site Rare Earth Ion Size in Controlling the Curie Temperature of Ba1−xRExTi1−x/4O3. Adv. Funct. Mater. 2013, 23, 491–495.

(31)

Dunbar, T. D.; Warren, W. L.; Tuttle, B. A.; Randall, C. A.; Tsur, Y. Electron Paramagnetic

Resonance

Investigations

of

Lanthanide-Doped

Barium

Titanate:  Dopant Site Occupancy. J. Phys. Chem. B 2004, 108, 908–917. (32)

Dawson, J. A.; Freeman, C. L.; Ben, L.-B.; Harding, J. H.; Sinclair, D. C. An Atomistic Study into the Defect Chemistry of Hexagonal Barium Titanate. J. Appl. Phys. 2011, 109, 84102.

(33)

Zulueta, Y. A.; Dawson, J. A.; Leyet, Y.; Guerrero, F.; Anglada‐Rivera, J.; Nguyen, M. T. The Potential Existence of Mixed Defect Incorporation Modes for Rare‐Earth Doped Cubic BaTiO3. Phys. Status Solidi 2016, 253, 733–737.

(34)

Muralidharan, M.; Anbarasu, V.; Elaya Perumal, A.; Sivakumar, K. Carrier Induced Ferromagnetism in Yb Doped SrTiO3 Perovskite System. J. Mater. Sci. Mater. Electron. 2014, 25, 4078–4087.

(35)

Akhtar, M. J.; Akhtar, Z.-U.-N.; Jackson, R. A.; Catlow, C. R. A. Computer Simulation Studies of Strontium Titanate. J. Am. Ceram. Soc. 1995, 78, 421–

21 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

428. (36)

Erhart, P.; Albe, K. Thermodynamics of Mono- and Di-Vacancies in Barium Titanate. J. Appl. Phys. 2007, 102, 84111.

(37)

Klyukin, K.; Alexandrov, V. Effect of Intrinsic Point Defects on Ferroelectric Polarization Behavior of SrTiO3. Phys. Rev. B 2017, 95, 35301.

(38)

Stokes, S. J.; Islam, M. S. Defect Chemistry and Proton-Dopant Association in BaZrO3 and BaPrO3. J. Mater. Chem. 2010, 20, 6258–6264.

(39)

Buin, A.; Comin, R.; Xu, J.; Ip, A. H.; Sargent, E. H. Halide-Dependent Electronic Structure of Organolead Perovskite Materials. Chem. Mater. 2015, 27, 4405–4412.

(40)

Dawson, J. A.; Tanaka, I. Proton Trapping in Y and Sn Co-Doped BaZrO3. J. Mater. Chem. A 2015, 3, 10045–10051.

(41)

Dawson, J. A.; Chen, H.; Tanaka, I. Combined Ab Initio and Interatomic Potentials Based Assessment of the Defect Structure of Mn-Doped SrTiO3. J. Phys. Chem. C 2014, 118, 14485–14494.

(42)

Harding, J. H. Computer Simulation of Defects in Ionic Solids. Reports Prog. Phys. 1990, 53, 1403.

(43)

Catlow, C. R. A. Computer Modelling in Inorganic Chemistry; Academic Press: San Diego, 1997.

(44)

Dawson, J. A.; Sinclair, D. C.; Harding, J. H.; Freeman, C. L. A-Site Strain and Displacement in Ba1–xCaxTiO3 and Ba1–xSrxTiO3 and the Consequences for the Curie Temperature. Chem. Mater. 2014, 26, 6104–6112.

(45)

Dick, B. G.; Overhauser, A. W. Theory of the Dielectric Constants of Alkali Halide Crystals. Phys. Rev. 1958, 112, 90–103.

(46)

Mott, N. F.; Littleton, M. J. Conduction in Polar Crystals. I. Electrolytic

22 Environment ACS Paragon Plus

Page 22 of 24

Page 23 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

The Journal of Physical Chemistry

Conduction in Solid Salts. Trans. Faraday Soc. 1938, 34, 485–499. (47)

Gale, J. D.; Rohl, A. L. The General Utility Lattice Program (GULP). Mol. Simul. 2003, 29, 291–341.

(48)

Wyckoff, R. W. G. Crystal Structures. John Wiley, New York, 1963.

(49)

Lytle,

F.

W.

X‐Ray

Diffractometry

of

Low‐Temperature

Phase

Transformations in Strontium Titanate. J. Appl. Phys. 1964, 35, 2212–2215. (50)

Gröting, M.; Hayn, S.; Albe, K. Chemical Order and Local Structure of the Lead-Free Relaxor Ferroelectric Na1/2Bi1/2TiO3. J. Solid State Chem. 2011, 184, 2041–2046.

(51)

Morrison, F. D.; Sinclair, D. C.; West, A. R. Doping Mechanisms and Electrical Properties of La-Doped BaTiO3 Ceramics. Int. J. Inorg. Mater. 2001, 3, 1205–1210.

(52)

Morrison, F. D.; Sinclair, D. C.; West, A. R. Characterization of LanthanumDoped Barium Titanate Ceramics Using Impedance Spectroscopy. J. Am. Ceram. Soc. 2001, 84, 531–538.

(53)

Gomann, K.; Borchardt, G.; Schulz, M.; Gomann, A.; Maus-Friedrichs, W.; Lesage, B.; Kaitasov, O.; Hoffmann-Eifert, S.; Schneller, T. Sr Diffusion in Undoped and La-Doped SrTiO3 Single Crystals under Oxidizing Conditions. Phys. Chem. Chem. Phys. 2005, 7, 2053–2060.

23 Environment ACS Paragon Plus

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 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

TABLE OF CONTENTS GRAPHIC

24 Environment ACS Paragon Plus

Page 24 of 24