Skip to main content

ORIGINAL RESEARCH article

Front. Mater., 22 December 2022
Sec. Mechanics of Materials
This article is part of the Research Topic Multi-field Coupling Mechanics of Smart Materials and Structures View all 6 articles

The morphologies of twins in tetragonal ferroelectrics

Ningbo He&#x;Ningbo He1Cuiping Li&#x;Cuiping Li1Chihou Lei
Chihou Lei2*Yunya Liu
Yunya Liu1*
  • 1Key Laboratory of Low Dimensional Materials and Application Technology of Ministry of Education, School of Materials Science and Engineering, Xiangtan University, Xiangtan, China
  • 2Department of Aerospace and Mechanical Engineering, Saint Louis University, Saint Louis, MO, United States

Twins, as a special structure, have been observed in a small number of experiments on ferroelectrics. A good understanding of the morphologies of twins is very important for twin engineering applications in ferroelectric materials. In this work, the morphologies of twins in tetragonal ferroelectrics are investigated using the compatibility analysis of the transformation strains and spontaneous polarization and the energetic analysis. The tetragonal BaTiO3 single crystal is chosen as an example for the material system. The results show that the lamellar twin structures with 111 as the twin plane have been identified by both compatibility analysis and energetic analysis, which is consistent with the experimental observations. In addition to 111 twin structures, 12¯1 and 21¯5 twin structures can also appear in tetragonal ferroelectrics. Furthermore, stable state uncharged twin boundaries and metastable charged twin boundaries are distinguished by energetic analysis, where the spontaneous polarization is compatible across the uncharged twin boundaries, while it is not compatible across the charged twin boundaries. The results predicted by the compatibility analysis and energetic analysis are consistent, thus providing a pathway for understanding the morphologies of twins.

1 Introduction

Due to their piezoelectric, dielectric, pyroelectric, ferroelectric, and electrocaloric properties, ferroelectric materials have shown great potential in sensors, actuators, capacitors, memories, solid-state refrigeration, and micro-electromechanical systems (Garcia and Bibes, 2012; Hwang et al., 2014; Martin and Rappe, 2016; Li et al., 2018a; Stadlober et al., 2019; Isaeva and Topolov, 2021; Manan et al., 2021; Shan et al., 2021; Asapu et al., 2022; Dan et al., 2022). The region with the same polarization is known as the ferroelectric domain, which corresponds to a ferroelectric variant. Different ferroelectric domains correspond to different ferroelectric variants with different transformation strains and spontaneous polarizations (Liu and Li, 2009). To form stable domain configurations, these variants must satisfy the compatibility conditions on transformation strains and spontaneous polarizations, leading to minimum energy. In tetragonal ferroelectrics, 90° and 180° domains have been observed (Le et al., 2013). In rhombohedral ferroelectrics, 71°, 109°, and 180° domains have been observed (Chen et al., 2007; Anthoniappen et al., 2017; Wan et al., 2021). In addition, the charged domain walls, across which the strain compatibility condition is satisfied, but the polarization compatibility condition is not satisfied, have been reported in the theoretical and experimental aspects (Liu et al., 2007; Liu and Li, 2009; Li et al., 2016). Therefore, the formation of domain structure is the basis for understanding the functional properties of ferroelectric materials. It is also the technological application of underpinned ferroelectric materials, making domain engineering being an effective way to enhance the performance of ferroelectrics (Shelke et al., 2011; Li et al., 2017; Geng et al., 2020; Qiu et al., 2020; Chen et al., 2022; Lei and Liu, 2022; Liu et al., 2022; Xu et al., 2022).

Domain engineering for ferroelectric materials has been extensively studied from both experimental and theoretical perspectives (Shelke et al., 2011; Li et al., 2017; Geng et al., 2020; Qiu et al., 2020; Chen et al., 2022; Lei and Liu, 2022; Liu et al., 2022; Xu et al., 2022). It is well known that twin structures are often observed in metals and alloys (Li et al., 2018b; Song et al., 2020; Li et al., 2023). Based on the formation of twins, twin engineering is widely used to enhance the mechanical properties of materials (Cheng et al., 2018; Li et al., 2022; Chen et al., 2023), thermal properties (Gao et al., 2021), electrical properties (Lei et al., 2018). Although domain engineering in ferroelectrics has received extensive attention, twins in ferroelectrics have rarely been reported. It is noticed that the 90° ferroelectric domain in some literature is called “twin”, but it is not a twin structure whose lattice should be symmetrical about the twin planes.

Twin structures have been experimentally observed in BaTiO3 powders (Eibl et al., 1987; Qin et al., 2010). For example, Qin et al. synthesized twinned BaTiO3 powders using BaCl2 and TiO2 at low temperatures, and their twin plane is 111 plane (Qin et al., 2010). By calcining a powder mixture of BaCO3 and TiO2 at 1100°C, Eibl et al. observed a 111 twin structures in the BaTiO3 ceramics (Eibl et al., 1987). Also, Cao et al. fabricated twinned structures in BaTiO3 films by a thermomechanical treatment technique, and BaTiO3 films with twin structures exhibit a smaller coercive field (Cao et al., 2015). This suggests that twin engineering may be used to enhance the properties of ferroelectrics. However, there are currently few reports on ferroelectric twinning, while the morphologies of ferroelectric twin, including its grain shape and orientation remain unclear. In addition to the 111 twin plane, it is still unclear whether ferroelectric materials can form other twin planes, while the revealing of morphologies and twin planes for ferroelectric twins is very important for utilizing twin engineering to enhance the performance of ferroelectrics. It is well known that the stable structures in ferroelectrics are energetically stable states, which should satisfy compatibility conditions on transformation strains and spontaneous polarization. Our previous work developed an energetic analysis method for ferroelectric domain patterns based on the equivalent inclusion method (Liu et al., 2007; Liu and Li, 2009). Thus, in this article, we analyze the morphologies of twins in tetragonal ferroelectrics by using compatibility analysis and energetic analysis, and the energy minimizing twin morphologies, including orientation and shape, are identified, as well as the determined twin planes.

2 Theoretical frameworks

2.1 Tetragonal twin variants

In experiments, (111) twins have been observed in BaTiO3 powders and films at room temperature (Eibl et al., 1987; Qin et al., 2010; Cao et al., 2015; Cao et al., 2017). It is well known that BaTiO3 is tetragonal at room temperature. Guided by the experimental results (Eibl et al., 1987; Qin et al., 2010; Cao et al., 2015), the schematic of a tetragonal twin structure is plotted in Figure 1A, where the tetragonal unit cells TW1 and TW2 are symmetric about the twin plane. Each tetragonal unit cell has six possible polarization directions, corresponding to the six ferroelectric variants. Thus, there are a total of 12 ferroelectric variants in tetragonal twinned ferroelectrics. In their respective cubic crystallographic axes, such as TW1 in its local coordinate system x1TW1x2TW1x3TW1 and TW2 in its local coordinate system x1TW2x2TW2x3TW2, the transformation strain εi (i = 1,…,6) and the spontaneous polarization Pi of each variant are given as:

ε1=ε2=β000α000α,P1=P2=Pt00,ε3=ε4=α000β000α,P3=P4=0Pt0,ε5=ε6=α000α000β,P5=P6=00Pt,(1)

where α and β are lattice constants of the materials, and Pt is the module of the spontaneous polarization. It is noticed that the variants with opposite spontaneous polarizations have the same transformation strains. Considering that the tetragonal unit cells TW1 and TW2 have different local crystallographic axes, they have different transformation strains and spontaneous polarizations, which can be obtained by the tensor rotation of Eq. 1, which are denoted by εiSTW1,PiSTW1 and εjSTW2,PjSTW2, where the superscripts TW1 and TW2 indicate the symmetric unit cells about the twin boundary (TB), and the subscripts i and j (i, j = 1,…,6) label the ferroelectric variants on the sides of the TB.

FIGURE 1
www.frontiersin.org

FIGURE 1. (A) Schematic of the tetragonal ferroelectric twin structure, where twin boundary (TB) represents the twin plane, TW1 and TW2 respectively denote the symmetric unit cells about the twin boundary. x1TW1x2TW1x3TW1 and x1TW2x2TW2x3TW2 are the local coordinate systems along cubic crystallographic axes of TW1 and TW2; (B) the twin boundary formed by different variants, where n represents the normal to the twin plane, εiSTW1,PiSTW1 and εjSTW2,PjSTW2 are the transformation strains and spontaneous polarizations of the variants on either side of the twin boundary.

2.2 Compatibility condition

In ferroelectrics, the stable interface between various ferroelectric variants is a consequence of energy minimization, across which the transformation strains and spontaneous polarizations of the variants satisfy the compatibility conditions (Shu and Bhattacharya, 2001; Li and Liu, 2004; Liu and Li, 2009; O'Reilly et al., 2022). For twin structures, the variants quantified by εiSTW1,PiSTW1 and εjSTW2,PjSTW2 form a twin boundary, as shown in Figure 1B. Thus, the compatibility conditions on the transformation strains and the spontaneous polarizations across the twin boundary are given by:

PiSTW1PjSTW2n=0,(2)
εiSTW1εjSTW2=12an+na,(3)

where n represents the normal to the twin boundary, a represents the shear vector along the TB, and refers to the standard tensor product of two vectors. Thus, the compatibility conditions Eqs 2, 3 allow us to determine the variants εiSTW1,PiSTW1 and εjSTW2,PjSTW2 that form an energetically stable twin boundary plane.

2.3 Energetic of morphology

While the compatibility analysis is capable of identifying the orientation of twin structures, it cannot identify the grain shapes of the twins, as well as the charged interface, which is energetically metastable due to polarizations incompatibility (Liu et al., 2007; Liu and Li, 2009). These deficiencies can be overcome by energy analysis using the equivalent inclusion method, which has been successfully applied to analyze the morphologies of ferroelectric domain patterns (Liu et al., 2007) and the precipitates in thermoelectric compounds (Liu et al., 2014).

For a ferroelectric crystal possessing a transformation strain εs and spontaneous polarization Ps, the electromechanical behavior is governed by the following constitutive equations:

ε=FEσ+dtE+εs,(4)
D=dσ+kσE+Ps,(5)

where ε and σ are the strain and stress, respectively. D and E are the electric displacement and field, respectively. FE is the elastic compliance tensor under constant electric field, kσ is the dielectric constant under constant stress, d is the piezoelectric coefficient. The superscript t denotes the transpose of the tensor. By change of variables, the ferroelectric constitutive Eqs 4, 5 can be transformed into:

σ=CDεεshtDPs,(6)
E=hεεs+ζεDPs.(7)

The two types of constitutive Eqs. 47 can be rewritten as the following matrix forms:

Y=MX+Ys,(8)
X=LYYs,(9)

where

Y=εD,X=σE,Ys=εsPs,M=FEdtdkσ,L=M1=CDhthζε

with the superscript 1 denoting matrix inverse. Since the electromechanical moduli M and L are positive definite, it is convenient to analyze the energy of twin structures using the above constitutive Eqs. 47.

In order to analyze the stable morphologies of stable twin structures in tetragonal ferroelectrics, the equivalent inclusion method is adopted (Dunn and Taya, 1993; Liu et al., 2007; Liu and Li, 2009). The twin structure is treated as a matrix region D and an inhomogeneous inclusion Ω, where the matrix is composed of TW1 with electromechanical moduli denoted by LTW1 and transformation strain and spontaneous polarization given by YSTW1, while the inhomogeneous inclusion is composed of TW2 with LTW2 and YSTW2, as shown in Figure 2. Using the matrix TW1 as a reference, the effective transformation strain and spontaneous polarization of the inhomogeneous inclusion TW2 can be written as:

YS=YSTW2YSTW1.(10)

The electromechanical fields can be obtained by Eshleby’s classical solution (Liu et al., 2007; Liu and Li, 2009).

FIGURE 2
www.frontiersin.org

FIGURE 2. Schematic of the inhomogeneous inclusion, where TW1 and TW2 represent the unit cells in the twin structure.

Consider a uniform external field X0 is applied to the twin structure in Figure 2. If an inhomogeneous inclusion does not exist, a uniform Y0 will be induced in matrix. But due to the emergence of inhomogeneous inclusion TW2, the additional disturbance fields X and Y are induced. Thus, the electromechanical field for the inhomogeneous inclusion TW2 can be expressed by (Liu and Li, 2009):

X0+X=LTW2Y0+YYS,(11)

which can be computed by Eshelby’s equivalent inclusion method (Liu et al., 2007; Liu and Li, 2009), such as:

LTW2Y0+YYS=LTW1Y0+YYSY**,(12)

where the inhomogeneous inclusion TW2 is replaced by an equivalent inclusion, which has the same electromechanical moduli LTW1 of matrix and additional transformation strain and spontaneous polarization Y**, which ensures the equivalency of electromechanical fields between TW2 and the equivalent inclusion. For ellipsoidal inclusion, it is well known that the disturbance electromechanical field in the inclusion can be calculated by (Wang, 1992; Dunn and Taya, 1993; Liu et al., 2007; Liu and Li, 2009):

Y=TYS+Y**),(13)

where

T=N2SN3N

with

N2=I0etkε,N3=00etkε,N=Idkσ10kσ1.

In the above expressions, I is the unit four-rank tensor,

e=FE1d,kε=kσdtFE1d.

S is the piezoelectric Eshelby’s tensors, which is a function of the electromechanical modulus LTW1 of the matrix TW1, as well as a function of the shape and orientation of the inclusion TW2 (Dunn and Taya, 1993). Combining Eqs 12, 13, the fictitious transformation strain and spontaneous polarization Y** can be expressed as

Y**=LTW1LTW2TLTW11LTW2LTW1Y0LTW1LTW2TLTW11LTW2YSYS,(14)

from which the electromechanical field in the twin structure can be determined.

The potential energy of the twin structure system can then be expressed by:

W=D12X0+X)Y0+YYS)X0Y0+Y)dx,(15)

where the first term represents the elastic and electric energies, and the second term represents the work done by external loading. In the absence of the inhomogeneous inclusion TW2, the effective transformation strain and spontaneous polarization YS and the disturbance fields vanish, leading to the potential energy expressed by:

W0=12DX0Y0dx.(16)

Thus, due to the existence of the inhomogeneous inclusion TW2, which infers the formation of a twin structure, the resulting energy variation is

W=WW0=12ΩX0Y**+XYS+2X0YSdx,(17)

where the Hill’s condition is adopted for simplification (Dunn and Taya, 1993; Li and Dunn, 1999). In this work, we consider the scenario of no external field, and the energy variation is further reduced to:

W=12ΩXYSdx.(18)

Eq. 18 allows us to analyze the energy variation due to the emergence of twin structures, including its orientation and shape of the inhomogeneous inclusion TW2. The equilibrium morphologies of the twins can be identified by minimizing the energy variation with respect to the orientation and shape of the inhomogeneous inclusion TW2. For the inhomogeneous inclusion TW2, its shape can be assumed to be ellipsoidal, and the orientation can be described using Euler angles in the global coordinate system.

3 Results and discussions

In order to determine the morphologies of twins in tetragonal ferroelectrics, we employ the above theory to implement the compatibility analysis and energetic analysis of twin structures. We choose BaTiO3 as the material system, whose electromechanical moduli associated with the single domain with the polarization along 001 in its cubic crystallographic axes is M. The transformation strain and spontaneous polarization are listed in Table 1. Notice that the physical quantities of other variants in different coordinate systems can be obtained by tensor rotations.

TABLE 1
www.frontiersin.org

TABLE 1. The electromechanical moduli M of single-domain with polarization along [001] in its cubic crystallographic axes (FE: 10–12 m2/N; d: 10–12 C/N; κσ/κ0; where κ0 = 8.85 × 10–12 C2/Nm2 is the permittivity of free space), transformation strain ε; and spontaneous polarization P(C/m2) of tetragonal BaTiO3 single crystal (Zgonik et al., 1994; Liu and Li, 2009).

3.1 Compatibility analysis

Guided by the experimental results (Eibl et al., 1987; Qin et al., 2010; Cao et al., 2015), the tetragonal twin structure is schematically illustrated in Figure 1A. There are six variants with polarizations along the cubic crystallographic axes of TW1 unit cell. More explicitly, the polarizations are :P1STW1100TW1, P2STW11¯00TW1, P3STW1010TW1, P4STW101¯0TW1, P5STW1001TW1, P6STW1001¯TW1 in the coordinate system x1TW1x2TW1x3TW1. Also, there are six variants with polarizations along the cubic crystallographic axes of TW2 unit cells, which are denoted by P1STW212¯2¯TW1, P2STW21¯22TW1, P3STW22¯12¯TW1, P4STW221¯2TW1, P5STW2221¯TW1, P6STW22¯2¯1TW1 in the coordinate system x1TW1x2TW1x3TW1. Note that the orientational correspondences between TW1 unit cells and TW2 unit cells are: 100TW2132323TW1, 010TW2231323TW1, and 001TW2232313TW1. The transformation matrix that rotates from the x1TW2x2TW2x3TW2 coordinate to the x1TW1x2TW1x3TW1 coordinate is then given by:

RTW2TW1=13122212221.(19)

In the x1TW1x2TW1x3TW1 coordinate, the transformation strain εiSTW1 and spontaneous polarization PiSTW1 of each variant for TW1 have been given in Eq. 1, while the transformation strain εiSTW2 and spontaneous polarization PiSTW2 of TW2 can be obtained by the tensor rotation of Eq. 1 as:

PiSTW2=RTW2TW1Pi,(20)
εiSTW2=RTW2TW1εiRTW2TW11,(21)

resulting in:

ε1STW2=ε2STW2=8α+β92α2β92α2β92α2β95α+4β94α+4β92α2β94α+4β95α+4β9,P1STW2=P2STW2=Pt32Pt32Pt3,ε3STW2=ε4STW2=5α+4β92α2β94α+4β92α2β98α+β92α2β94α+4β92α2β95α+4β9,P3STW2=P4STW2=2Pt3Pt32Pt3,ε5STW2=ε6STW2=5α+4β94α+4β92α2β94α+4β95α+4β92α2β92α2β92α2β98α+β9,P5STW2=P6STW2=2Pt32Pt3Pt3.(22)

For the twin structure composed of TW1 εiSTW1,PiSTW1 and TW2 εjSTW2,PjSTW2, the normal to the twin plane can be computed by solving Eqs. 2 and 3. Based on this compatibility analysis of the transformation strains and spontaneous polarizations, the twin planes between different tetragonal variants from TW1 and TW2 that may form twin structures are summarized in Table 2. It is found that the following pairs of variants can form (111)-twin in tetragonal ferroelectric: variants ε1STW1,P1STW1 with ε2STW2,P2STW2, variants ε2STW1,P2STW1 with ε1STW2,P1STW2, variants ε3STW1,P3STW1 with ε4STW2,P4STW2, variants ε4STW1,P4STW1 with ε3STW2,P3STW2, variants ε5STW1,P5STW1 with ε5STW2,P5STW2, and variants ε6STW1,P6STW1 with ε6STW2,P6STW2, which have also been observed in experiments (Eibl et al., 1987; Cheng et al., 2006; Wu et al., 2006; Qin et al., 2010; Cao et al., 2015). Three representatives of these (111)-twins are schematically shown in Figure 3. Additionally, Table 2 shows that 12¯1 and 21¯5 twinned structures may also exist in tetragonal ferroelectrics, which have not been observed in currently reported experiments. Some predicted structures of 12¯1 and 21¯5 twins are schematically shown in Figure 4.

TABLE 2
www.frontiersin.org

TABLE 2. The possible twin planes formed between various tetragonal variants identified by compatibility analysis. Note that all the polarization directions and twin plane normal are indicated in the local coordinate system x1TW1x2TW1x3TW1.

FIGURE 3
www.frontiersin.org

FIGURE 3. Schematics of the possible (111)-twins in tetragonal ferroelectrics; (A) Tetragonal ferroelectric variants ε5STW1,P5STW1 with ε5STW2,P5STW2; (B) Tetragonal ferroelectric variants ε1STW1,P1STW1 with ε2STW2,P2STW2; (C) Tetragonal ferroelectric variants ε3STW1,P3STW1 with ε4STW2,P4STW2. Here x1TW1x2TW1x3TW1 and x1TW2x2TW2x3TW2 indicate the local coordinate system along cubic crystallographic axes of TW1 and TW2, respectively.

FIGURE 4
www.frontiersin.org

FIGURE 4. Schematics of the possible 12¯1-twin and 21¯5-twin in tetragonal ferroelectrics. (A) Tetragonal ferroelectric variants ε5STW1,P5STW1 with ε1STW2,P1STW2; (B) Tetragonal ferroelectric variants ε5STW1,P5STW1 with ε4STW2,P4STW2. Here x1TW1x2TW1x3TW1 and x1TW2x2TW2x3TW2 indicate the local coordinate system along cubic crystallographic axes of TW1 and TW2, respectively.

3.2 Energetic analysis

Via the compatibility analysis, it is found that twin structures can be formed between different ferroelectric variants, and the twin planes are mainly 111,12¯1,21¯5. But the compatibility analysis cannot determine the shapes of the twin grains. To this end, we apply Eq. 18 to perform an energetic analysis of tetragonal ferroelectric twins. Guided by the results of compatibility analysis, three types of twinned structures with variants ε5STW1,P5STW1 with ε5STW2,P5STW2, variants ε5STW1,P5STW1 with ε1STW2,P1STW2, variants ε5STW1,P5STW1 with ε4STW2,P4STW2 are considered in energetic analysis, which may form 111,12¯1,21¯5 twinned structures respectively.

3.2.1 Twinned structure with variants ε5STW1,P5STW1 and ε5STW2,P5STW2

To be specific, the inhomogeneous inclusion TW2 is assumed to be spheroidal. For the twinned structure with variants ε5STW1,P5STW1 and ε5STW2,P5STW2, for the convenience of calculation, we choose the global coordinate system x1x2x3 as shown in Figure 5A, where x1100G12120TW1, x2010G12120TW1 and x3001G001TW1, which makes the polarization vectors of the two variants lie on the same plane denoted by HJOQ, as shown in Figure 5B. The orientation of inhomogeneous inclusion TW2 can then be described by only one Euler angle θ, which is the angle between the x3 axis and the rotational axis X3 of the spheroid, as shown in Figure 5B. Notice that x2X2. The transformation matrix that rotates from local x1TW1x2TW1x3TW1 and x1TW2x2TW2x3TW2 coordinates to the global x1x2x3 coordinate are respectively:

RTW1G=1212012120001,RTW2G=13213243212120232313.(23)

FIGURE 5
www.frontiersin.org

FIGURE 5. Energetic analysis of twinned structure with variants ε5STW1,P5STW1 and ε5STW2,P5STW2: (A) Schematic of the relationship between the global coordinate system x1x2x3 and local coordinate system x1TW1x2TW1x3TW1; (B) Schematic of variant ε5STW2,P5STW2 as an inhomogeneous inclusion within the variant ε5STW1,P5STW1 as the matrix; (C) the potential energy variation as a function of the shape aspect ratio a3/a1 and orientation angle θ of the inhomogeneous inclusion TW2; and (D) the corresponding potential energy variation at a3/a1 = 0.0001; The schematics of the uncharged (E) and charged (F) twin boundaries corresponding to the two wells in the energy landscape.

The shape aspect ratio of the inhomogeneous inclusion TW2 can be defined by a3/a1=a3/a2, where a1,a2,a3 are the dimensions along the principal axes of the spheroidal. Based on Eq. 18, we calculate the potential energy variation as a function of the shape aspect ratio a3/a1 and the orientation angle θ of inhomogeneous inclusion TW2, as shown in Figure 5C. It is found that, when the aspect ratio a3/a1 approaches zero, two energy wells exist in the energy landscape, indicating that the stable twin structure is lamellar. Indeed, lamellae twins have been observed in BaTiO3 ceramics (Oppolzer and Schmelz, 1983). Furthermore, one of the energy wells occurs at θ=125.3°, while another occurs at θ=36.6°. The well at θ=125.3° refers to the uncharged twin boundary with 111 as the twin plane, across which the transformation strains and spontaneous polarizations are compatible, as shown in Figure 5E, which is consistent with those predicted by the compatibility analysis [as listed in Table 2]. On the other hand, the well at θ=36.6° refers to the charged twin boundary with 112¯ as the twin plane, across which only the transformation strain is compatible, but the spontaneous polarization is not compatible, as shown in Figure 5F. Figure 5D reveals that the energy well at θ=125.3° is lower than that at θ=36.6°, suggesting that the uncharged twin boundary is in a stable state, while the charged twin boundary is in a metastable state. A steep energy barrier exists between the two energy wells, making it hard to convert from the charged twin boundary to the uncharged twin boundary. The charged domain walls have been observed in many experiments (Sluka et al., 2012; Li et al., 2016; Wu et al., 2021), which are similar to the charged twin boundaries. Both of them are in metastable states, across which the spontaneous polarizations are also not compatible.

3.2.2 Twinned structure with variants ε5STW1,P5STW1 and ε1STW2,P1STW2

For the twinned structure with ε5STW1,P5STW1 and ε1STW2,P1STW2, for the convenience of calculation, we choose the global coordinate system x1x2x3 as shown in Figure 6A, where x1100G15250TW1, x2010G25150TW1, and x3001G001TW1, such that polarization vectors of the two variants lie on the same plane HJOQ, as shown in Figure 6B. For this case, the potential energy variation due to the emergence of the inhomogeneous inclusion TW2 as a function of its shape aspect ratio a3/a1 and the orientation angle θ is plotted in Figure 6C. It is observed that one energy well occurs at θ=114.1° in the energy landscape with a3/a1 approaching zero, which corresponds to the uncharged twin boundary with 12¯1 as the twin plane, across which the transformation strains and spontaneous polarizations are compatible, as shown in Figure 6E, consistent with the prediction by the compatibility analysis [as listed in Table 2].

FIGURE 6
www.frontiersin.org

FIGURE 6. Energetic analysis of twinned structure with variants ε5STW1,P5STW1 and ε1STW2,P1STW2: (A) Schematic of the relationship between the global coordinate system x1x2x3 and local coordinate system x1TW1x2TW1x3TW1; (B) Schematic of variant ε1STW2,P1STW2 as an inhomogeneous inclusion within the variant ε5STW1,P5STW1 as the matrix; (C) the potential energy variation as a function of the shape aspect ratio a3/a1 and orientation angle θ of the inhomogeneous inclusion TW2; and (D) the corresponding potential energy variation at a3/a1 = 0.0001; The schematic of the uncharged (E) twin boundary corresponding to the well in the energy landscape.

3.2.3 Twinned structure with variants ε5STW1,P5STW1 and ε4STW2,P4STW2

For the twinned structure with ε5STW1,P5STW1 and ε4STW2,P4STW2, we choose the global coordinate x1x2x3, as shown in Figure 7A, where x1100G25150TW1, x2010G15250TW1, and x3001G001TW1. In the global coordinate x1x2x3, the polarization vectors of the two variants lie on the same plane HJOQ, as shown in Figure 7B. We calculate the potential energy variation due to the emergence of the inhomogeneous inclusion TW2 as a function of its shape aspect ratio a3/a1 and orientation angle θ in Figure 7C. It can be seen that, when the aspect ratio a3/a1 approaches zero, there exists two energy wells, at θ=155.9° and θ=69.3° respectively. The energy well at θ=155.9° corresponds to the uncharged twin boundary with 21¯5 as the twin plane. The transformation strains and spontaneous polarizations across this twin plane are compatible, as shown in Figure 7E, which is consistent with the prediction by the compatibility analysis [as listed in Table 2]. The other energy well located at θ=69.3° is metastable, resulting in the charged twin boundary with 2¯11 as the twin plane, across which only the transformation strains are compatible, while the spontaneous polarizations are not compatible, as shown in Figure 7F.

FIGURE 7
www.frontiersin.org

FIGURE 7. Energetic analysis of twinned structure with variants ε5STW1,P5STW1 and ε4STW2,P4STW2: (A) Schematic of the relationship between the global coordinate system x1x2x3 and local coordinate system x1TW1x2TW1x3TW1; (B) Schematic of variant ε4STW2,P4STW2 as an inhomogeneous within the variant ε5STW1,P5STW1 as the matrix; (C) the potential energy variation as a function of the shape aspect ratio a3/a1 and orientation angle θ of the inhomogeneous inclusion TW2; and (D) the corresponding potential energy variation at a3/a1 = 0.0001; The schematics of the uncharged (E) and charged (F) twin boundaries corresponding to the two wells in the energy landscape.

3.3 Comparison between compatibility analysis and energetic analysis

For different combinations of variants, including the pairs ε5STW1,P5STW1 with ε5STW2,P5STW2, ε5STW1,P5STW1 with ε1STW2,P1STW2, ε5STW1,P5STW1 with ε4STW2,P4STW2, the predicted results of the compatibility analysis and energetic analysis are summarized in Table 3. It can be found that the compatibility analysis on transformation strains can identify twin planes, but it cannot distinguish whether the twin plane is uncharged or charged. The energetic analysis not only can identify the twin plane, but also can further determine whether the twin plane is uncharged in a steady state or charged in a metastable state. The compatibility analysis on both transformation strains and spontaneous polarizations can only determine uncharged twin planes, as charged twin planes do not satisfy the compatibility in spontaneous polarizations. Overall, the results from compatibility analysis and energetic analysis are consistent. It has been demonstrated that twin engineering can reduce the coercive field of ferroelectric materials (Cao et al., 2015) due to the increased ferroelectric variants, suggesting that twin engineering may be used to enhance the other properties of ferroelectrics.

TABLE 3
www.frontiersin.org

TABLE 3. Comparison of prediction results between compatibility analysis and energetic analysis. Note that the twin planes are specified in the x1TW1x2TW1x3TW1 coordinate.

4 Conclusion

We have investigated the morphologies of twins in tetragonal BaTiO3 single crystals by using the compatibility analysis of the transformation strains and spontaneous polarizations and the energetic analysis. It is found that the lamellar twin structures with 111 as the twin plane have been identified by both compatibility analysis and energetic analysis, which agrees well with experimental observations. The results also show that the 12¯1 and 21¯5 twin structures can exist in tetragonal ferroelectrics. In addition, the stable state uncharged twin boundaries and metastable charged twin boundaries are distinguished by energetic analysis, since the spontaneous polarizations are compatible across the uncharged twin boundaries, while the spontaneous polarizations are not compatible across the charged twin boundaries. The results predicted by the compatibility analysis and energetic analysis agree well with each other, offering a pathway for understanding the morphologies of twins in ferroelectrics.

Data availability statement

The raw data supporting the conclusion of this article will be made available by the authors, upon reasonable request.

Author contributions

NH and CL have contributed equally to this work. All authors contributed to conceptualization, theoretical calculations, analysis and writing the original draft. CL and YL revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially supported by the National Natural Science Foundation of China (11572276 and 12172318), the Hunan Provincial Natural Science Foundation of China (2021JJ10006), and the science and technology innovation Program of Hunan Province (2022RC3069). He also acknowledges support from Hunan Provincial Innovation Foundation for Postgraduate (CX20200628). Lei acknowledges the startup fund supported jointly by the Dean of Parks College and the Provost of Saint Louis University.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Anthoniappen, J., Chang, W. S., Soh, A. K., Tu, C. S., Vashan, P., and Lim, F. S. (2017). Electric field induced nanoscale polarization switching and piezoresponse in Sm and Mn co-doped BiFeO3 multiferroic ceramics by using piezoresponse force microscopy. Acta Mater. 132, 174–181. doi:10.1016/j.actamat.2017.04.034

CrossRef Full Text | Google Scholar

Asapu, S., Pagaduan, J. N., Zhuo, Y., Moon, T., Midya, R., Gao, D., et al. (2022). Large remnant polarization and great reliability characteristics in W/HZO/W ferroelectric capacitors. Front. Mater. 9, 969188. doi:10.3389/fmats.2022.969188

CrossRef Full Text | Google Scholar

Cao, S. G., Li, Y. S., Wu, H. H., Wang, J., Huang, B. L., and Zhang, T. Y. (2017). Stress-induced cubic-to-hexagonal phase transformation in perovskite nanothin films. Nano Lett. 17 (8), 5148–5155. doi:10.1021/acs.nanolett.7b02570

PubMed Abstract | CrossRef Full Text | Google Scholar

Cao, S. G., Wu, H. H., Ren, H., Chen, L. Q., Wang, J., Li, J. Y., et al. (2015). A novel mechanism to reduce coercive field of ferroelectric materials via {111} twin engineering. Acta Mater. 97, 404–412. doi:10.1016/j.actamat.2015.07.009

CrossRef Full Text | Google Scholar

Chen, C., Liu, H., Lai, Q., Mao, X., Fu, J., Fu, Z., et al. (2022). Large-scale domain engineering in two-dimensional ferroelectric CuInP2S6 via giant flexoelectric effect. Nano Lett. 22 (8), 3275–3282. doi:10.1021/acs.nanolett.2c00130

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, L., Li, F., Gao, B. T., Zhou, C., Wu, J., Deng, S. Q., et al. (2023). Excellent energy storage and mechanical performance in hetero-structure BaTiO3-based relaxors. Chem. Eng. J. 452, 139222. doi:10.1016/j.cej.2022.139222

CrossRef Full Text | Google Scholar

Chen, Y. B., Katz, M. B., Pan, X. Q., Das, R. R., Kim, D. M., Baek, S. H., et al. (2007). Ferroelectric domain structures of epitaxial (001) BiFeO3 thin films. Appl. Phys. Lett. 90 (7), 072907. doi:10.1063/1.2472092

CrossRef Full Text | Google Scholar

Cheng, S.-Y., Ho, N.-J., and Lu, H.-Y. (2006). Crystallographic relationships of the {111} growth twins in tetragonal barium titanate determined by electron-backscatter diffraction. J. Am. Ceram. Soc. 89 (11), 3470–3474. doi:10.1111/j.1551-2916.2006.01231.x

CrossRef Full Text | Google Scholar

Cheng, Z., Zhou, H. F., Lu, Q. H., Gao, H. J., and Lu, L. (2018). Extra strengthening and work hardening in gradient nanotwinned metals. Science 362 (6414), eaau1925. doi:10.1126/science.aau1925

PubMed Abstract | CrossRef Full Text | Google Scholar

Dan, H. Y., Li, H. Y., and Yang, Y. (2022). Flexible ferroelectric materials-based triboelectric nanogenerators for mechanical energy harvesting. Front. Mater. 9, 939173. doi:10.3389/fmats.2022.939173

CrossRef Full Text | Google Scholar

Dunn, M. L., and Taya, M. (1993). An analysis of piezoelectric composite materials containing ellipsoidal inhomogeneities. Proc. R. Soc. Lond. A 443 (1918), 265–287. doi:10.1098/rspa.1993.0145

CrossRef Full Text | Google Scholar

Eibl, O., Pongratz, P., Skalicky, P., and Schmelz, H. (1987). Formation of (111) twins in BaTiO3 ceramics. J. Am. Ceram. Soc. 70 (8), C-195–C-197. doi:10.1111/j.1151-2916.1987.tb05724.x

CrossRef Full Text | Google Scholar

Gao, Y. F., Ning, W. B., Zhang, X. L., Liu, Y. Z., Zhou, Y. G., and Tang, D. W. (2021). The effective regulation of nanotwinning on the multichannel thermal transport in hybrid organic–inorganic halide perovskite. Nano Energy 82, 105747. doi:10.1016/j.nanoen.2021.105747

CrossRef Full Text | Google Scholar

Garcia, V., and Bibes, M. (2012). Inside story of ferroelectric memories. Nature 483 (7389), 279–280. doi:10.1038/483279a

PubMed Abstract | CrossRef Full Text | Google Scholar

Geng, W. R., Guo, X. W., Zhu, Y. L., Wang, Y. J., Tang, Y. L., Han, M. J., et al. (2020). Oxygen octahedral coupling mediated ferroelectric-antiferroelectric phase transition based on domain wall engineering. Acta Mater. 198, 145–152. doi:10.1016/j.actamat.2020.08.007

CrossRef Full Text | Google Scholar

Hwang, G. T., Park, H., Lee, J. H., Oh, S., Park, K. I., Byun, M., et al. (2014). Self-powered cardiac pacemaker enabled by flexible single crystalline PMN-PT piezoelectric energy harvester. Adv. Mater. 26 (28), 4880–4887. doi:10.1002/adma.201400562

PubMed Abstract | CrossRef Full Text | Google Scholar

Isaeva, A. N., and Topolov, V. Y. (2021). Lead-free 0–3-type composites: From piezoelectric sensitivity to modified figures of merit. J. Adv. Dielect. 11 (02), 2150010. doi:10.1142/S2010135X21500107

CrossRef Full Text | Google Scholar

Le, D.-T., Kwon, S.-J., Yeom, N.-R., Lee, Y.-J., Jeong, Y.-H., Chun, M.-P., et al. (2013). Effects of the domain size on local d33 in tetragonal (Na0.53K0.45Li0.02)(Nb0.8Ta0.2)O3 ceramics. J. Am. Ceram. Soc. 96 (1), 174–178. doi:10.1111/j.1551-2916.2012.05445.x

CrossRef Full Text | Google Scholar

Lei, C. H., and Liu, Y. Y. (2022). Correlations between local electrocaloric effect and domains in ferroelectric crystals. Appl. Phys. Lett. 121 (10), 102902. doi:10.1063/5.0094473

CrossRef Full Text | Google Scholar

Lei, S. H., Fan, H. Q., Fang, J. W., Ren, X. H., Ma, L. T., and Tian, H. L. (2018). Unusual devisable high-performance perovskite materials obtained by engineering in twins, domains, and antiphase boundaries. Inorg. Chem. Front. 5 (3), 568–576. doi:10.1039/C7QI00711F

CrossRef Full Text | Google Scholar

Li, F., Zhang, S. J., Xu, Z., and Chen, L.-Q. (2017). The contributions of polar nanoregions to the dielectric and piezoelectric responses in domain-engineered relaxor-PbTiO3 crystals. Adv. Funct. Mater. 27 (18), 1700310. doi:10.1002/adfm.201700310

CrossRef Full Text | Google Scholar

Li, J. L., Li, F., Xu, Z., and Zhang, S. J. (2018a). Multilayer lead-free ceramic capacitors with ultrahigh energy density and efficiency. Adv. Mater. 30 (32), 1802155. doi:10.1002/adma.201802155

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, J. Y., and Dunn, M. L. (1999). Analysis of microstructural fields in heterogeneous piezoelectric solids. Int. J. Eng. Sci. 37 (6), 665–685. doi:10.1016/S0020-7225(98)00091-3

CrossRef Full Text | Google Scholar

Li, J. Y., and Liu, D. (2004). On ferroelectric crystals with engineered domain configurations. J. Mech. Phys. Solids 52 (8), 1719–1742. doi:10.1016/j.jmps.2004.02.011

CrossRef Full Text | Google Scholar

Li, L. H., Liu, W. H., Qi, F. G., Wu, D., and Zhang, Z. Q. (2022). Effects of deformation twins on microstructure evolution, mechanical properties and corrosion behaviors in magnesium alloys - a review. J. Magnes. Alloy. 10 (9), 2334–2353. doi:10.1016/j.jma.2022.09.003

CrossRef Full Text | Google Scholar

Li, L. L., Zhang, Z. J., Zhang, P., and Zhang, Z. F. (2023). A review on the fatigue cracking of twin boundaries: Crystallographic orientation and stacking fault energy. Prog. Mater. Sci. 131, 101011. doi:10.1016/j.pmatsci.2022.101011

CrossRef Full Text | Google Scholar

Li, L. Z., Britson, J., Jokisaari, J. R., Zhang, Y., Adamo, C., Melville, A., et al. (2016). Giant resistive switching via control of ferroelectric charged domain walls. Adv. Mater. 28 (31), 6574–6580. doi:10.1002/adma.201600160

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, Q., Xue, S. C., Wang, J., Shao, S., Kwong, A. H., Giwa, A., et al. (2018b). High-strength nanotwinned al alloys with 9R phase. Adv. Mat. 30 (11), 1704629. doi:10.1002/adma.201704629

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, Y., Yang, B., Lan, S., Pan, H., Nan, C.-W., and Lin, Y.-H. (2022). Perspectives on domain engineering for dielectric energy storage thin films. Appl. Phys. Lett. 120 (15), 150501. doi:10.1063/5.0090739

CrossRef Full Text | Google Scholar

Liu, Y. Y., Chen, L. Q., and Li, J. Y. (2014). Precipitate morphologies of pseudobinary Sb2Te3–PbTe thermoelectric compounds. Acta Mater. 65, 308–315. doi:10.1016/j.actamat.2013.10.072

CrossRef Full Text | Google Scholar

Liu, Y. Y., and Li, J. Y. (2009). Energetic analysis of ferroelectric domain patterns by equivalent inclusion method. J. Mater. Sci. 44 (19), 5214–5224. doi:10.1007/s10853-009-3536-2

CrossRef Full Text | Google Scholar

Liu, Y. Y., Liu, J. J., Xie, S. H., and Li, J. Y. (2007). Energetics of charged domain walls in ferroelectric crystals. Appl. Phys. Lett. 91 (17), 172910. doi:10.1063/1.2803765

CrossRef Full Text | Google Scholar

Manan, A., Rehman, M. U., Ullah, A., Ahmad, A. S., Iqbal, Y., Qazi, I., et al. (2021). High energy storage density with ultra-high efficiency and fast charging–discharging capability of sodium bismuth niobate lead-free ceramics. J. Adv. Dielect. 11 (04), 2150018. doi:10.1142/S2010135X21500181

CrossRef Full Text | Google Scholar

Martin, L. W., and Rappe, A. M. (2016). Thin-film ferroelectric materials and their applications. Nat. Rev. Mater. 2 (2), 16087. doi:10.1038/natrevmats.2016.87

CrossRef Full Text | Google Scholar

O'Reilly, T., Holsgrove, K., Gholinia, A., Woodruff, D., Bell, A., Huber, J., et al. (2022). Exploring domain continuity across BaTiO3 grain boundaries: Theory meets experiment. Acta Mater. 235, 118096. doi:10.1016/j.actamat.2022.118096

CrossRef Full Text | Google Scholar

Oppolzer, H., and Schmelz, H. (1983). Investigation of twin lamellae in BaTiO3 ceramics. J. Am. Ceram. Soc. 66 (6), 444–446. doi:10.1111/j.1151-2916.1983.tb10078.x

CrossRef Full Text | Google Scholar

Qin, S. B., Liu, D., Zheng, F. F., Zuo, Z. Y., Liu, H., and Xu, X. G. (2010). (111) Twinned BaTiO3 microcrystallites. CrystEngComm 12 (10), 3003–3007. doi:10.1039/B925863A

CrossRef Full Text | Google Scholar

Qiu, C., Wang, B., Zhang, N., Zhang, S., Liu, J., Walker, D., et al. (2020). Transparent ferroelectric crystals with ultrahigh piezoelectricity. Nature 577 (7790), 350–354. doi:10.1038/s41586-019-1891-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Shan, D. L., Lei, C. H., Cai, Y. C., Pan, K., and Liu, Y. Y. (2021). Mechanical control of electrocaloric response in epitaxial ferroelectric thin films. Int. J. Solids Struct. 216, 59–67. doi:10.1016/j.ijsolstr.2021.01.020

CrossRef Full Text | Google Scholar

Shelke, V., Mazumdar, D., Srinivasan, G., Kumar, A., Jesse, S., Kalinin, S., et al. (2011). Reduced coercive field in BiFeO3 thin films through domain engineering. Adv. Mater. 23 (5), 669–672. doi:10.1002/adma.201000807

PubMed Abstract | CrossRef Full Text | Google Scholar

Shu, Y. C., and Bhattacharya, K. (2001). Domain patterns and macroscopic behaviour of ferroelectric materials. Philos. Mag. B 81 (12), 2021–2054. doi:10.1080/13642810108208556

CrossRef Full Text | Google Scholar

Sluka, T., Tagantsev, A. K., Damjanovic, D., Gureev, M., and Setter, N. (2012). Enhanced electromechanical response of ferroelectrics due to charged domain walls. Nat. Commun. 3 (1), 748. doi:10.1038/ncomms1751

PubMed Abstract | CrossRef Full Text | Google Scholar

Song, M., Zhou, G., Lu, N., Lee, J., Nakouzi, E., Wang, H., et al. (2020). Oriented attachment induces fivefold twins by forming and decomposing high-energy grain boundaries. Science 367 (6473), 40–45. doi:10.1126/science.aax6511

PubMed Abstract | CrossRef Full Text | Google Scholar

Stadlober, B., Zirkl, M., and Irimia-Vladu, M. (2019). Route towards sustainable smart sensors: Ferroelectric polyvinylidene fluoride-based materials and their integration in flexible electronics. Chem. Soc. Rev. 48 (6), 1787–1825. doi:10.1039/C8CS00928G

PubMed Abstract | CrossRef Full Text | Google Scholar

Wan, H., Luo, C. T., Liu, C., Chang, W. Y., Yamashita, Y., and Jiang, X. N. (2021). Alternating current poling on sliver-mode rhombohedral Pb(Mg1/3Nb2/3)O3-PbTiO3 single crystals. Acta Mater. 208, 116759. doi:10.1016/j.actamat.2021.116759

CrossRef Full Text | Google Scholar

Wang, B. (1992). Three-dimensional analysis of an ellipsoidal inclusion in a piezoelectric material. Int. J. Solids Struct. 29 (3), 293–308. doi:10.1016/0020-7683(92)90201-4

CrossRef Full Text | Google Scholar

Wu, H. J., Ning, S. C., Waqar, M., Liu, H. J., Zhang, Y., Wu, H.-H., et al. (2021). Alkali-deficiency driven charged out-of-phase boundaries for giant electromechanical response. Nat. Commun. 12 (1), 2841. doi:10.1038/s41467-021-23107-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, Y.-C., Lee, C.-C., Lu, H.-Y., McCauley, D. E., and Chu, M. S. H. (2006). The {111} growth twins in tetragonal barium titanate. J. Am. Ceram. Soc. 89 (5), 1679–1686. doi:10.1111/j.1551-2916.2006.00948.x

CrossRef Full Text | Google Scholar

Xu, X. Y., Wang, T. X., Chen, P. C., Zhou, C., Ma, J. N., Wei, D. Z., et al. (2022). Femtosecond laser writing of lithium niobate ferroelectric nanodomains. Nature 609 (7927), 496–501. doi:10.1038/s41586-022-05042-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Zgonik, M., Bernasconi, P., Duelli, M., Schlesser, R., Günter, P., Garrett, M. H., et al. (1994). Dielectric, elastic, piezoelectric, electro-optic, and elasto-optic tensors of BaTiO3 crystals. Phys. Rev. B 50 (9), 5941–5949. doi:10.1103/PhysRevB.50.5941

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: ferroelectric, twin, compatibility analysis, energetic analysis, ferroelectric variant

Citation: He N, Li C, Lei C and Liu Y (2022) The morphologies of twins in tetragonal ferroelectrics. Front. Mater. 9:1100964. doi: 10.3389/fmats.2022.1100964

Received: 17 November 2022; Accepted: 25 November 2022;
Published: 22 December 2022.

Edited by:

Qingzhen Yang, Xi’an Jiaotong University, China

Reviewed by:

Hong-Hui Wu, University of Science and Technology Beijing, China
Jiaming Zhu, Shandong University, China

Copyright © 2022 He, Li, Lei and Liu. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Chihou Lei, chihou.lei@slu.edu; Yunya Liu, yyliu@xtu.edu.cn

These authors have contributed equally to this work

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.