Bi nonstoichiometry and composition engineering in (1 − x)Bi1+yFeO3+3y/2xBaTiO3 ceramics

Ting Zheng, Yi Ding and Jiagang Wu*
Department of Materials Science, Sichuan University, 610064, P. R. China. E-mail: wujiagang0208@163.com; msewujg@scu.edu.cn

Received 20th April 2016 , Accepted 15th September 2016

First published on 15th September 2016


Abstract

Bi is easily evaporated during the preparation of BiFeO3, and moreover its phase structure and electrical properties can also be modified by doping BaTiO3. In this work, the effects of Bi nonstoichiometry as well as BaTiO3 on the phase purity, microstructure, and electrical properties of (1 − x)Bi1+yFeO3+3y/2xBaTiO3 (0.25 ≤ x ≤ 0.35 and −0.1 ≤ y ≤ 0.25) ceramics were studied. A rhombohedral–pseudocubic (R–C) phase boundary was formed in the ceramics with a wide composition range of x = 0.30 and −0.05 ≤ y ≤ 0.15, leading to an enhanced piezoelectricity (d33 ∼ 180 pC/N). In addition, one can find that dense microstructure as well as the improvement of Curie temperature (TC ∼ 506 °C) and thermal stability of piezoelectricity (Ta = 400 °C) can be achieved in the ceramics with excess Bi (x = 0.30, y = 0.05) with respect to the ones with Bi deficiency. We expect that the investigation on Bi nonstoichiometry and the addition of BaTiO3 will point out an effective method to fabricate high-performance BFO-based ceramics.


1. Introduction

Lead-free BiFeO3–BaTiO3 (BFO–BTO) solid solutions have become one of the most promising candidates for high-temperature piezoelectric applications due to their environment-friendly characteristics and relatively high piezoelectricity.1–7 The key properties of BFO–BTO ceramics, such as piezoelectricity and ferroelectricity, have been extensively investigated in the past several years.2–6 Particularly, the electrical properties of BFO–BTO ceramics can be improved by the construction of phase boundaries.3,7 Among these studies, the main methods to improve their electrical properties can be shown here: (1) improvement of preparation process.8–14 It was previously reported that the modified conventional solid-state method with a quenched process is effective for the enhancement of electrical properties due to the freezing of disordered defect states.8,9,14 (2) Site engineering.15–21 Ion substitutions for Bi (e.g., La, Nd, Eu) or Fe (e.g., Al, Sc, Co, Cr) in BFO–BTO ceramics were carried out, and improved electrical properties were also achieved.15–21 (3) Construction of ternary systems.22–26 A series of third members [e.g., Bi0.5Na0.5TiO3, Bi0.5K0.5TiO3, BiM0.5Ti0.5O3 (M: Zn, Ni, Mg)] were utilized to modify BFO–BTO ceramics with the aim of improving their resistivity and electrical properties.22–26 Recently, a large piezoelectricity (d33 = 402 pC/N) together with a high Curie temperature (TC = 454 °C) have been reported in BFO–BTO–BiGaO3 ceramics, indicating that the BFO–BTO material systems may become the most promising candidate.27 However, little attention was paid to the phase boundary and electrical properties of pure BFO–BTO ceramics with Bi nonstoichiometry.

For BFO-based ceramics, one of the tough issues is low electrical resistivity, thus leading to the difficulty in poling process. Previously, it was reported that the Bi evaporation during high-temperature process can induce the formation of some defects (i.e., free electrons or charged vacancies), thus resulting in a high conductivity.28 It was found that the Bi content strongly affected electrical properties of Bi-based piezoceramics. For example, the difficulty in the poling process for (Bi0.5+xNa)TiO3 ceramics may be ascribed to the pinning of domain walls because the Bi evaporation can bring about oxygen vacancies.29 An enhancement of piezoelectricity and ferroelectricity can also be achieved in Bi2O3-doped (K0.5Na0.5)NbO3 ceramics.30 Therefore, it is essential to investigate the effects of Bi nonstoichiometry on phase structure and electrical properties of BFO–BTO ceramics. Although the effects of Bi excess on the structure and piezoelectricity of BFO–BTO ceramics were carried out, there are few reports on the effects of Bi deficiency on their electrical properties, and the corresponding ferroelectricty as well as thermal stability were also few mentioned.31 Therefore, it is necessary to conduct the systematic researches about the effects of Bi nonstoichiometry as well as BaTiO3 on the phase structure and electrical properties of BFO ceramics.

In this study, the effects of Bi nonstoichiometry and BaTiO3 on the phase structure, microstructure, and electrical properties of (1 − x)Bi1+yFeO3+3y/2xBaTiO3 (x = 0.25–0.35 and y = −0.1–0.25) ceramics were studied systematically. Rhombohedral–pseudocubic (R–C) phase boundary was formed in the ceramics through the addition of BaTiO3. A large piezoelectric coefficient (d33 ∼ 180 pC/N) can be achieved in a wide composition range of x = 0.30 and −0.05 ≤ y ≤ 0.15. In addition, appropriate excess Bi can promote the ferroelectricity, the Curie temperature, and the thermal stability of BFO–BTO ceramics. The related physical mechanisms were also illuminated.

2. Experimental procedure

Lead-free (1 − x)Bi1+yFeO3+3y/2xBaTiO3 (0.25 ≤ x ≤ 0.35 and −0.1 ≤ y ≤ 0.25) ceramics were synthesized by the conventional solid-state method with a quenching method. Here, we need to point out that the y range of −0.1 to 0.25 is the nominal value, and then the actual y composition range was identified to be −0.02–0.22 according to the EDS data. In addition, the error of atom percent of each sample is not more than 10%, indicating that chemical composition is relatively stable in all samples. Therefore, in order to simplify the reader's understanding, we will still discuss related phenomenon using the nominal composition range. And we will also point out the corresponding actual Bi content in the manuscript. Raw materials including Bi2O3 (99%), Fe2O3 (99%), BaCO3 (99%), and TiO2 (98%) were weighted and ball milled for 24 h with alcohol. Those mixing slurries were dried and calcined at 700 °C for 2 h. After that, the pellets with 10 mm diameter and 0.6 mm thickness were pressed under a pressure of 10 MPa using 6–8 wt% polyvinyl alcohol (PVA) as a binder. After burning off PVA at 500 °C for 3 h, the pellets were sintered at 960–1000 °C for 3 h in air and then quenched in water, and the heating and cooling rates are 10 °C min−1. For electrical measurement, both sides of the sintered samples were pasted on silver slurry and then fired at 600 °C for 10 min. The samples were poled at 120 °C in a silicone oil bath under a dc field of 5 kV mm−1.

X-ray diffraction (XRD) (Bruker D8 Advanced XRD, Bruker AXS Inc., Madison, WI, CuKα) was used to measure the phase structure. The temperature dependence of capacitance and dielectric loss were tested using an LCR analyzer in the temperature range of 20–550 °C. The field emission-scanning electron microscopy (FE-SEM) (JSM-7500, Japan) was used to characterize the surface microstructure of the sintered samples. Their polarization–electric (PE) hysteresis loops were measured by Radiant Precision Workstation (USA) at f = 10 Hz and room temperature. The piezoelectric constant (d33) was measured by piezo-d33 meter (ZJ-3 A, China) and approximately 24 h after poling.

3. Results and discussion

Due to the close relationships between phase structure and piezoelectricity, it is essential to identify the phase structure of BFO–BTO ceramics. Fig. 1(a) and (b) show the room-temperature XRD patterns of the ceramics as a function of x (y = 0.05) and y (x = 0.30), measured in the 2θ range of 20–70°. The addition of BaTiO3 into BiFeO3 matrix can form a stable solid solution, resulting in a pure perovskite structure. The phase transition of the ceramics is confirmed by the expanded XRD patterns at 2θ = 31–32° and 38–40°. It is clearly to find that there is a composition dependence of their phase structure. For example, the ceramics can endure the phase transition from rhombohedral to pseudocubic phases with the increase of x, leading to the formation of R–C phase boundary for 0.30 ≤ x ≤ 0.33. In addition, a pseudocubic phase can be detected in the ceramics with y = −0.1 (actual y content: y = −0.02) [Fig. 1(b)], which can also be identified by its εrT curve in Fig. 5(c). With the further increase of y contents, the multiphase coexistence containing rhombohedral and pseudocubic phases was formed. Therefore, composition design has induced the formation of a wide R–C phase boundary in the ceramics. In addition, it can be observed from Fig. 1(b) that the characteristic peak shifted to low angle with the increase of y contents (0 ≤ y ≤ 0.15), and then shifted to high angle when y = 0.25 (actual y content: y = 0.22).
image file: c6ra10264f-f1.tif
Fig. 1 XRD patterns of (1 − x)Bi1+yFeO3+3y/2xBaTiO3 ceramics with different (a) x (y = 0.05) and (b) y (x = 0.30) contents.

As we all know, the ion occupying can result in the formation of point defects, which has a close relationship with microstructure and electrical properties of perovskite piezoceramics.32 Here, we systematically explored the doping mechanisms of the ceramics with different y contents. Obviously, the severe loss of Bi2O3 induced the formation of Bi vacancies (VBi3+). The volatilization of Bi2O3 also caused the formation of both Bi (VBi3+) and O vacancies (VO2+). However, the excess Bi contents compensated the Bi volatilization and resulted in different ion occupying mechanisms. The ionic radii of Bi3+, Ba2+, and Ti4+ are 0.103 nm, 0.135 nm, and 0.0605 nm respectively. Therefore, excess Bi3+ occupied the Ba or Ti site. Combined with the characteristic peaks shift, the variations of crystal cell volume and the ionic radius, we can deduce that slight excess of Bi3+ (0.05 ≤ y ≤ 0.15) can substitute Ti site, leading to the expansion of crystal cell volume as well as the peak shift to low angle. However, the excessive Bi contents of y = 0.25 (actual y content: y = 0.22) can substitute Ba site, resulting in the shrinkage of crystal cell volume and the peak shift to high angle. The defect reactions were listed below:

 
2Bi3+ + 3O2− → Bi2O3↑ + 2VBi3− + 3VO2+ (1)
 
Bi2O3 → 2BiTi + 3OO + VO2+ (2)
 
Bi2O3 → 2BiBa+ + 3OO + VBa2− (3)

As a result, the point defects contain a large number of Bi vacancies (VBi3−) and some O vacancies (VO2+) for the ceramics with Bi deficiency due to both the severe loss of Bi2O3 and the volatilization of Bi2O3. The eqn (1) exhibits the Bi volatilization as well as the formation of both Bi VBi3− and VO2+ for the ceramics with y = 0. Eqn (2) illustrates that the Bi occupied Ti site, leading to the formation of a small number of VO2+ in the ceramics with 0.05 ≤ y ≤ 0.15. Finally, the formation of VBa2− for the ceramics with y = 0.25 (actual y content: y = 0.22) was due to the occupying of Bi for Ba site.

In order to analyze the valence state and oxygen vacancies of BFO–BTO ceramics with Bi nonstoichiometry, we carried out the element composition by XPS spectra, as shown in Fig. 2(a)–(c). In addition, the fitting results were included in Table 1. Fig. 2(a) shows the fitting results of Bi 4f for the ceramics with y = −0.1 (actual y content: y = −0.02), 0.05, and 0.25 (actual y content: y = 0.22). The main peaks corresponding to Bi 4f5/2 and Bi 4f7/2 were observed, and the separation between two peaks is about 5.3 eV, confirming the existence of Bi3+.33–35 One can also see a small amount of atomic or metallic Bi exists located at ∼159 eV, indicating the existence of Bi+ (ref. 34) In addition, the percentage of Bi+ decreased firstly and then increased with the increase of Bi contents, reaching the lowest value of 5% for y = 0.05 [Table 1]. Fig. 2(b) shows the asymmetrical O 1s photoelectron peaks divided into three sub-peaks. The lowest peaks in binding energy located at ∼529 eV correspond to the cation–oxygen bonds, while two further higher energy peaks located at ∼530 eV and at ∼532 eV were assigned to adsorbed surface H2O as well as the presence of oxygen vacancies.35–37 The percentage of oxygen vacancies of the ceramics with the increase of y content was 64.0%, 63.1%, and 59.6%, respectively. This result indicated that oxygen vacancies decreased gradually as the Bi contents increased. The experimental results just coincide with the previous theoretical analysis. At last, the high resolution XPS scans of Fe 2p region was exhibited in Fig. 2(c), and the energy peak corresponding to Fe 2p3/2 can be used to calculate the ratio of Fe3+ and Fe2+. Fe3+ locates at ∼711 eV, while Fe2+ locates at a lower energy peak of 709 eV. According to the fitting results, the concentration ratios of Fe3+ to Fe2+ were respectively 50.5[thin space (1/6-em)]:[thin space (1/6-em)]49.5, 53.7[thin space (1/6-em)]:[thin space (1/6-em)]46.3, and 59.6[thin space (1/6-em)]:[thin space (1/6-em)]40.4 with increasing y contents (see Table 1), demonstrating that the concentrations of Fe2+ decreased gradually with the increase of Bi. Generally speaking, Fe2+ and oxygen vacancies can emerge simultaneously in BFO for charge compensation, and thus less Fe2+ means less oxygen vacancies.35,38 In this work, the concentration of Fe2+ decreased gradually with the increase of Bi contents, which is just coincident with the results of the concentration variation of oxygen vacancies. This result indicates that excess Bi contents are indeed beneficial to the decrease of Fe2+ and oxygen vacancies.


image file: c6ra10264f-f2.tif
Fig. 2 Valence states of the elements in (1 − x)Bi1+yFeO3+3y/2xBaTiO3 ceramics with y = −0.1, 0.05, and 0.25. (a) Bi 4f, (b) O 1s, and (c) Fe 2p.
Table 1 Fitted peaks of Bi 4f, O 1s, and Fe 2p
Bi 4f Peaks Atomic% O 1s Peaks Atomic% Fe 2p3/2 Peaks Atomic%
y = −0.1 163.8 42.3 y = −0.1 531.6 64.0 y = −0.1 711.1 50.5
159.3 6.1 531.0 3.1    
158.5 51.6 529.1 32.9 709.6 49.5
y = 0.05 164.1 43.6 y = 0.05 531.9 63.1 y = 0.05 711.1 53.7
159.5 5.0 530.4 10.5    
158.8 51.4 529.3 26.4 709.5 46.3
y = 0.25 163.9 43.5 y = 0.25 531.7 59.6 y = 0.25 711.0 59.6
159.3 5.9 530.4 12.4    
158.5 50.5 529.1 28.0 709.5 40.4


Fig. 3 displays the surface microstructure of the ceramics with different x (y = 0.05) and y (x = 0.30). As shown in Fig. 3(a)–(d), dense microstructure, conspicuous grain boundaries as well as similar grain sizes can be found in those ceramics with different x, demonstrating that the microstructure cannot be greatly affected by the variations of BT content. Nevertheless, there is an obvious change in microstructure for the ceramics with different y, as shown in Fig. 3(e)–(h). For example, the seriously refined grain sizes can be formed in the ceramics with y = −0.1 (actual y content: y = −0.02), a much denser microstructure can be obtained when appropriate excess of Bi2O3 were introduced, while too many Bi2O3 results in a loose microstructure of the ceramics with y = 0.25 (actual y content: y = 0.22). Thus a severe deficiency of Bi2O3 prohibited the grain growths and caused the refinement of grain sizes due to the formation of a large number of VBi3− and VO2+ [Fig. 3(e)]. However, too excessive Bi2O3 led to a loose microstructure [Fig. 3(h)]. As a result, appropriate compensation of Bi2O3 can contribute to the densification of the microstructure in BFO–BTO ceramics due to the easy volatilization of Bi during their high-temperature sintering. In order to clearly exhibit the variations of grain sizes, their grain size distribution can be calculated and presented in Fig. 4. One can see that the average grain sizes increased firstly and then decreased as both x and y increased, reaching a maximum value of about 7.07 μm for x = 0.30 and y = 0.05. The reduced grain sizes for the ceramics with y = −0.1(actual y content: y = −0.02) and y = 0.25 (actual y content: y = 0.22) may be due to the prohibited grain growth caused by massive point defects, such as VBa2−, VBi3− and VO2+.


image file: c6ra10264f-f3.tif
Fig. 3 The surface microstructure of (1 − x)Bi1+yFeO3+3y/2xBaTiO3 ceramics with different x (y = 0.05) and y (x = 0.30) contents.

image file: c6ra10264f-f4.tif
Fig. 4 Grain size distributions of the (1 − x)Bi1+yFeO3+3y/2xBaTiO3 ceramics with different x (y = 0.05) and y (x = 0.30) content.

As discussed above, Bi nonstoichiometry seriously affected the microstructure of the ceramics. Therefore, we measured the EDS data of different areas of the ceramics and then calculated the average values of the actual atom percent, as shown in Table 2. First of all, it can be seen that the measured Bi atom percent increased with the increase of y contents. The average atom percent of Bi element is 13.76%, 14.02%, and 16.62% for the ceramics with y = −0.1, 0, and 0.25, respectively. In addition, the error of atom percent of each sample is not more than 10%, indicating that chemical compositions are relatively stable in all samples. At last, according the EDS data, the actual composition range of y can be considered to be −0.02 to 0.22. Therefore, we have given out the nominal and actual composition range of y in this manuscript at the same time for the easier understanding.

Table 2 EDS data of Bi element in the ceramics with y = −0.1, 0, and 0.25
Composition Area 1 Area 2 Average value Nominal value Error (%)
y = −0.1 13.67 13.86 13.76 12.78 7
y = 0 14.06 13.99 14.02 14 0.1
y = 0.25 16.92 16.32 16.62 16.91 2


Fig. 5(a) and (c) exhibit the temperature-dependent dielectric constant (εrT) of the ceramics as a function of x (y = 0.05) and y (x = 0.30), measured at room temperature∼550 °C and at f = 100 kHz. First, it is obvious to find that the εrT curve of the ceramics with y = −0.1 (actual content: y = −0.02) was seriously suppressed. The severe deficiency of Bi caused the formation of massive VBi3−, then leading to the increase of disorder degree in perovskite structure and the serious suppression of εrT curves. Besides, the composition dependence of Curie temperature (TC) of the ceramics was plotted in the inset of Fig. 5(a) and (c). We can know that their TC declined rapidly with increasing x contents. However, their TC values increased firstly and then decreased slightly with the increase of y, reaching a maximum value of ∼506 °C for x = 0.30 and y = 0.05. Consequently, appropriate compensation for Bi2O3 can greatly promote the TC values of BFO–BTO ceramics. Fig. 5(b) and (d) exhibit the temperature dependence of tan[thin space (1/6-em)]δ of the ceramics with x (y = 0.05) and y (x = 0.30). It can be seen that their tan[thin space (1/6-em)]δ reduced gradually when the measurement temperatures increased from room temperature to 250 °C, and their tan[thin space (1/6-em)]δ values ascended higher and higher with the further increase of temperatures due to the increased conductivity.18,22,25


image file: c6ra10264f-f5.tif
Fig. 5 The temperature dependence of: (a, c) εr, and (b, d) tan δ of the (1 − x)Bi1+yFeO3+3y/2xBaTiO3 ceramics as a function of x (y = 0.05) and y (x = 0.30) content. TC against x and y was plotted in the inset of (a) and (c), respectively.

Fig. 6(a) and (b) show the composition dependence of polarization–electric field (PE) loops of the ceramics, measured at room temperature and f = 10 Hz. Compared with ions-modified BFO ceramics, it is easier to attain the saturated PE loops in BaTiO3-modified ceramics. Their PE loops almost remained unchanged with the variations of BT contents [Fig. 6(a)], while the ceramics with too excess Bi2O3 contents (actual y content: y = 0.22) exhibited a slim PE curve [Fig. 6(b)]. In order to distinctly reveal the composition dependence of ferroelectricity, we plotted the remnant polarization (Pr) and coercive field (EC) of the ceramics against x and y, as shown in Fig. 6(c) and (d). According to recent literature, Pr almost has no changes in the composition range of 0.30 ≤ x ≤ 0.40 for (1 − x)BFO − xBTO ceramics, and a much poorer Pr value was observed in the ceramics with x = 0.20 and 0.50.27 Therefore, in this work, we conjectured that the week composition dependence of ferroelectricity was ascribed to the quenched process. The quenched process can partly compensate the ferroelectricity for the compositions deviated from phase boundaries region. Appropriate excessive Bi contents (y = 0.05) can enhance the ferroelectricity of BFO–BTO ceramics due to the dense microstructure and improved electrical insulation. The weakened ferroelectricity was observed in the ceramics with severe loss and excess Bi content (actual y content: y = −0.02 and y = 0.22). It can be interpreted that the weakened ferroelectricity is due to the difficult mobility of domain walls – induced by the increased amount of defects.


image file: c6ra10264f-f6.tif
Fig. 6 (a) and (b): PE loops of (1 − x)Bi1+yFeO3+3y/2xBaTiO3 ceramics with different x (y = 0.05) and y (x = 0.30) contents. (c) and (d): Pr and EC values as a function of x (y = 0.03) and y (x = 0.30) contents.

The composition dependence of d33, εr, and tan[thin space (1/6-em)]δ of the ceramics was presented in Fig. 7(a) and (b). It can be seen that their εr has weak composition dependence at room temperature. Here, the relatively high tan[thin space (1/6-em)]δ (∼0.1) may be attributed to the involvement of dielectric loss peak around room temperature. Some researchers have reported a strong composition dependence of dielectric properties at room temperature.39,40 However, there is a week composition dependence of dielectric properties for BFO–BTO ceramics with a quenched process reported by Song et al.27 Therefore, the week composition dependence of dielectric properties may be attributed to the quenched technique. In the past, Yu et al. studied the dielectric behavior of BFO–BTO–BZT ceramics with MnO2 addition and pointed out that high tan[thin space (1/6-em)]δ can be observed in the undoped ceramics due to the space charges, while the addition of MnO2 can greatly decrease its tan[thin space (1/6-em)]δ.24 In addition, the piezoelectricity of BFO–BTO ceramics can be greatly affected by the variations of x and y contents. As shown in Fig. 7(a), d33 increased from 84 pC/N to 180 pC/N with the increase of x (0.25–0.30), and then decreased to 130 pC/N as x further increased to 0.35. In addition, the observed d33 increased rapidly from 40 pC/N to 165 pC/N as y increased, and then slightly increased and fluctuated in the vicinity of 180 pC/N for y = −0.05–0.15 [Fig. 7(b)]. This variation of d33 with different y contents are similar to those of BNT or KNN-based ceramics with the addition of Bi2O3.28–30 As discussed above, the excessive deficiency or excess of Bi (actual y content: y = −0.02 or 0.22) can lead to the formation of a large number of VBi3− or VBa2−. It was thought that the pinning of domain walls resulted in the degraded piezoelectricity of a material.41 As a result, the pinning of domain walls caused by defects, the pseudocubic phase, and loose microstructure are totally responsible for the serious reduction of d33 in the ceramics with y = −0.1 (actual y content: y = −0.02) and 0.25 (actual y content: y=0.22). In addition, the R–C phase boundary and dense microstructure can totally contribute to the enhancement of piezoelectricity in BFO–BTO ceramics with −0.05 ≤ y ≤ 0.15 (x = 0.30).


image file: c6ra10264f-f7.tif
Fig. 7 d33, εr and tan[thin space (1/6-em)]δ as a function of (a) x (y = 0.05) and (b) y (x = 0.30) contents.

Fig. 8(a) and (b) exhibit the variations of d33 with annealing temperatures (Ta) of the ceramics. The conventional thermal depolarization theory illustrated that the switching back of the domains to the initial state leads to the loss of piezoelectricity with the increase of Ta.42 For all ceramics except for y = −0.1 (actual y content: y = −0.02), d33 has a slight increase and can maintain a relatively high value with the increase of Ta and then drops greatly with the further increase of Ta. For example, the ceramics with x = 0.30 have a much better thermal stability for Ta = 400 °C, while the ceramics with x = 0.25, 0.33, and 0.35 respectively show a poor thermal stability of d33 for a lower Ta of 300 °C, 350 °C, and 300 °C [Fig. 8(a)]. In addition, all ceramics except y = −0.1 (actual y content: y = −0.02) have a similar changed trend with the increase of Ta, whose Ta is 400 °C. Especially, the ceramics with y = 0.05 have a much larger d33 value (∼150 pC/N) as compared with other samples even if Ta increases to 450 °C, indicating that doping with optimum Bi can enhance the thermal stability of d33 for BFO–BTO ceramics. An enhanced thermal stability was also reported in lead free high temperature BF–BT ceramics with excess Bi content.43 As a result, large piezoelectricity (d33 ∼ 180 pC/N) and its good thermal stability can be simultaneously achieved in the ceramics with appropriate excess Bi content.


image file: c6ra10264f-f8.tif
Fig. 8 Thermal stability of (1 − x)Bi1+yFeO3+3y/2xBaTiO3 ceramics with different (a) x (y = 0.05) and (b) y (x = 0.30) contents.

4. Conclusion

The phase structure, microstructure, and electrical properties of BFO were sensitive to Bi nonstoichiometry and the addition of BaTiO3. An obvious refinement of grain sizes, serious diffused εrT curves, and degraded electrical properties appeared in the ceramics with severe Bi deficiency. However, the improvement of piezoelectricity (d33 ∼ 180 pC/N), ferroelectricity (Pr ∼ 25 μC cm−2), Curie temperature (TC ≥ 500 °C), and thermal stability can be observed in the ceramics with excess Bi because of the existence of phase boundary and dense microstructure as well as the increase of electrical resistivity. As a result, it is necessary to improve the electrical properties of BFO ceramics by introducing both Bi nonstoichiometry and BaTiO3.

Acknowledgements

Authors gratefully acknowledge the supports of the National Science Foundation of China (NSFC No. 51102173 and 51472169). We also thank Ms. Hui Wang for measuring the SEM patterns.

References

  1. M. Mahesh Kumar, A. Srinivas and S. V. Suryanarayana, Structure property relations in BiFeO3/BaTiO3 solid solutions, J. Appl. Phys., 2000, 87, 855 CrossRef.
  2. S. O. Leontsevw and R. E. Eitel, Dielectric and Piezoelectric Properties in Mn-Modified (1-x)BiFeO3-xBaTiO3 Ceramics, J. Am. Ceram. Soc., 2009, 92(12), 2957–2961 CrossRef.
  3. H. B. Yang, C. R. Zhou, X. Y. Liu, Q. Zhou, G. H. Chen, W. Z. Li and H. Wang, Piezoelectric properties and temperature stabilities of Mn- and Cu-modified BiFeO3-BaTiO3 high temperature ceramics, J. Eur. Ceram. Soc., 2013, 33, 1177–1183 CrossRef CAS.
  4. Y. Wei, X. T. Wang, J. T. Zhu, X. L. Wang and J. J. Jia, Dielectric, Ferroelectric, and Piezoelectric Properties of BiFeO3-BaTiO3 Ceramics, J. Am. Ceram. Soc., 2013, 96(10), 3163–3168 CAS.
  5. Z. Y. Cen, C. R. Zhou, H. B. Yang, Q. Zhou, W. Z. Li and C. L. Yuan, Structural, ferroelectric and piezoelectric properties of Mn-modified BiFeO3-BaTiO3 high-temperature ceramics, J. Mater. Sci.: Mater. Electron., 2013, 24, 3952–3957 CrossRef CAS.
  6. Y. X. Wei, X. T. Wang, J. J. Jia and X. L. Wang, Multiferroic and piezoelectric properties of 0.65BiFeO3-0.35BaTiO3 ceramic with pseudo-cubic symmetry, Ceram. Int., 2012, 38, 3499–3502 CrossRef CAS.
  7. H. B. Yang, C. R. Zhou, X. Y. Liu, Q. Zhou, G. H. Chen, H. Wang and W. Z. Li, Structural, microstructural and electrical properties of BiFeO3-BaTiO3 ceramics with high thermal stability, Mater. Res. Bull., 2012, 47, 4233–4239 CrossRef CAS.
  8. T. Rojac, M. Kosec, B. Budic, N. Setter and D. Damjanovic, Strong ferroelectric domain-wall pinning in BiFeO3 ceramics, J. Appl. Phys., 2010, 108, 074107 CrossRef.
  9. S. T. Zhang, M. H. Lu, D. Wu, Y. F. Chen and N. B. Ming, Larger polarization and weak ferromagnetism in quenched BiFeO3 ceramics with a distorted rhombohedral crystal structure, Appl. Phys. Lett., 2005, 87, 262907 CrossRef.
  10. J. Walker, P. Bryant, V. Kurusingal, C. Sorrell, D. Kuscer, G. Drazic, A. Bencan, V. Nagarajan and T. Rojac, Synthesis-phase-composition relationship and high electric-field-induced electromechanical behavior of samarium-modified BiFeO3 ceramics, Acta Mater., 2015, 83, 149–159 CrossRef CAS.
  11. S. X. Zhang, L. Wang, Y. Chen, D. L. Wang, Y. B. Yao and Y. W. Ma, Observation of room temperature saturated ferroelectric polarization in Dy substituted BiFeO3 ceramics, J. Appl. Phys., 2012, 111, 074105 CrossRef.
  12. X. X. Shi, X. Q. Liu and X. M. Chen, Structure evolution and piezoelectric properties across the morphotropic phase boundary of Sm-substituted BiFeO3 ceramics, J. Appl. Phys., 2016, 119, 064104 CrossRef.
  13. G. L. Yuan, K. Z. Baba-Kishi, J. M. Liu, S. W. Or, Y. P. Wang and Z. G. Liu, Multiferroic Properties of Single-Phase Bi0.85La0.15FeO3 Lead-Free Ceramics, J. Am. Ceram. Soc., 2006, 89(10), 3136–3139 CrossRef CAS.
  14. T. Zheng and J. Wu, Quenched bismuch ferrite-barium titanate lead -free piezoelectric ceramics, J. Alloys Compd., 2016, 676, 505–512 CrossRef CAS.
  15. C. R. Zhou, H. B. Yang, Q. Zhou, Z. Y. Cen, W. Z. Li, C. L. Yuan and H. Wang, Dielectric, ferroelectric and piezoelectric properties of La-substituted BiFeO3-BaTiO3 ceramics, Ceram. Int., 2013, 39, 4307–4311 CrossRef CAS.
  16. Q. J. Zheng, L. L. Luo, K. H. Lam, N. Jiang, Y. Q. Guo and D. M. Lin, Enhanced ferroelectricity, piezoelectricity, and ferromagnetism in Nd-modified BiFeO3-BaTiO3 lead-free ceramics, J. Appl. Phys., 2014, 116, 184101 CrossRef.
  17. M. J. Tian, L. Zhou, X. Zou, Q. J. Zheng, L. L. Luo, N. Jiang and D. M. Lin, Improved ferroelectricity and ferromagnetism of Eu-modified BiFeO3-BaTiO3 lead-free multiferroic ceramics, J. Mater. Sci.: Mater. Electron., 2015, 26, 8840–8847 CrossRef CAS.
  18. Z. Y. Cen, C. R. Zhou, H. B. Yang, Q. Zhou, W. Z. Li, C. L. Yan, L. Cao, J. Song and L. Peng, Remarkably High-Temperature Stability of Bi(Fe1-xAlx)O3-BaTiO3 Solid Solution with Near-Zero Temperature Coefficient of Piezoelectric Properties, J. Am. Ceram. Soc., 2013, 96(7), 2252–2256 CrossRef CAS.
  19. Y. Wan, Y. Li, Q. Li, W. Zhou, Q. J. Zheng, X. C. Wu, C. G. Xu, B. P. Zhu and D. M. Lin, Microstructure, Ferroelectric, Piezoelectric, and Ferromagnetic Properties of Sc-Modified BiFeO3-BaTiO3 Multiferroic Ceramics with MnO2 Addition, J. Am. Ceram. Soc., 2014, 97(6), 1809–1818 CrossRef CAS.
  20. C. R. Zhou, Z. Y. Cen, H. B. Yang, Q. Zhou, W. Z. Li, C. L. Yuan and H. Wang, Structure, electrical properties of Bi(Fe,Co)O3-BaTiO3 piezoelectric ceramics with improved Curie temperature, Phys. B, 2013, 410, 13–16 CrossRef CAS.
  21. X. H. Liu, Z. Xu, X. Y. Wei and X. Yao, Ferroelectric and Ferromagnetic Properties of 0.7BiFe1-xCrxO3-0.3BaTiO3 Solid Solutions, J. Am. Ceram. Soc., 2008, 91(11), 3731–3734 CrossRef CAS.
  22. Y. Li, N. Jiang, K. H. Lam, Y. Q. Guo, Q. J. Zheng, Q. Li, W. Zhou, Y. Wan and D. M. Lin, Structure, Ferroelectric, Piezoelectric, and Ferromagnetic Properties of BiFeO3-BaTiO3-Bi0.5Na0.5TiO3 Lead-Free Multiferroic Ceramics, J. Am. Ceram. Soc., 2014, 97(11), 3602–3608 CrossRef CAS.
  23. D. M. Lin, Q. J. Zheng, Y. Li, Y. Wan, Q. Li and W. Zhou, Microstructure, ferroelectric and piezoelectric properties of Bi0.5K0.5TiO3-modified BiFeO3-BaTiO3 lead-free ceramics with high Curie temperature, J. Eur. Ceram. Soc., 2013, 33, 3023–3036 CrossRef CAS.
  24. Y. Lin, L. L. Zhang and J. Yu, Stable piezoelectric property of modified BiFeO3-BaTiO3 lead-free piezoceramics, J. Mater. Sci.: Mater. Electron., 2015, 26, 8432–8441 CrossRef CAS.
  25. Q. Zhou, C. R. Zhou, H. B. Yang, G. H. Chen, W. Z. Li and H. Wang, Dielectric, Ferroelectric, and Piezoelectric Properties of Bi(Ni1/2Ti1/2)O3-Modified BiFeO3-BaTiO3 Ceramics with High Curie Temperature, J. Am. Ceram. Soc., 2012, 95(12), 3889–3893 CrossRef CAS.
  26. L. F. Zhu, B. P. Zhang, S. Li, L. Zhao, N. Wang and X. C. Shi, Enhanced piezoelectric properties of Bi(Mg1/2Ti1/2)O3 modified BiFeO3-BaTiO3 ceramics near the morphotropic phase boundary, J. Alloys Compd., 2016, 664, 602–608 CrossRef CAS.
  27. M. H. Lee, D. J. Kim, J. S. Park, S. W. Kim, T. K. Song, M. H. Kim, W. J. Kim, D. Do and I. K. Jeong, High-performance lead-free piezoceramics with high curie temperatures, Adv. Mater., 2015, 27, 6976–6982 CrossRef CAS PubMed.
  28. R. Z. Zuo, S. Su, Y. Wu, J. Fu, M. Wang and L. T. Li, Influence of A-site nonstoichiometry on sintering, microstructure and electrical properties of (Bi0.5Na0.5)TiO3 ceramics, Mater. Chem. Phys., 2008, 110, 311–315 CrossRef CAS.
  29. Y. S. Sung, J. M. Kim, J. H. Cho, T. K. Song, M. H. Kim and T. G. Park, Effects of Bi nonstoichiometry in (Bi0.5+xNa)TiO3 ceramics, Appl. Phys. Lett., 2011, 98, 012902 CrossRef.
  30. H. L. Du, D. J. Liu, F. S. Tang, D. M. Zhu, W. C. Zhou and S. B. Qu, Microstructure, Piezoelectric, and Ferroelectric Properties of Bi2O3-Added (K0.5Na0.5)NbO3 Lead-Free Ceramics, J. Am. Ceram. Soc., 2007, 90(9), 2824–2829 CrossRef CAS.
  31. C. R. Zhou, H. B. Yang, Q. Zhou, G. H. Chen, W. Z. Li and H. Wang, Effects of Bi excess on the structure and electrical properties of high-temperature BiFeO3-BaTiO3 piezoelectric ceramics, J. Mater. Sci.: Mater. Electron., 2013, 24, 1685–1689 CrossRef CAS.
  32. W. J. Wu, J. Li, D. Q. Xiao, M. Chen, Y. C. Ding and C. Q. Liu, Defect dipoles-driven ferroelectric behavior in potassium sodium niobate ceramics, Ceram. Int., 2014, 40, 13205–13210 CrossRef CAS.
  33. Z. Ai, W. Ho, S. Lee and L. Zhang, Efficient photocatalytic removal of NO in indoor air with hierarchical bismuth oxybromide nanoplate microspheres under visible light, Environ. Sci. Technol., 2009, 43(11), 4143–4150 CrossRef CAS PubMed.
  34. G. Tan, W. Liu, X. Xue and H. Hao, Relative thickness effects on multiferroic Bi0.9Gd0.1Fe0.96Mn0.04O3/NiFe2O4 bilayered thin films, J. Alloys Compd., 2014, 617, 265–270 CrossRef CAS.
  35. Y. Guo, P. Xiao, R. Wen, Y. Wan, Q. Zheng, D. Shi, K. Lam, M. Liu and D. Lin, Critical roles of Mn-ions in enhancing the insulation, piezoelectricity and multiferroicity of BiFeO3-based lead-free high temperature ceramics, J. Mater. Chem. C, 2015, 3(22), 5811–5824 RSC.
  36. W. Hu, Y. Liu, R. L. Withers, T. J. Frankcombe, L. Norén, A. Snashall, M. Kitchin, P. Smith, B. Gong, H. Chen, J. Schiemer, F. Brink and J. W. Leung, Electron-pinned defect-dipoles for high-performance colossal permittivity materials, Nat. Mater., 2013, 12(9), 821–826 CrossRef CAS PubMed.
  37. J. Lv, X. J. Lou and J. Wu, Defect Dipoles-induced Poling Characteristics and Ferroelectricity of Quenched Bismuth Ferrite-based Ceramics, J. Mater. Chem. C, 2016, 4 Search PubMed.
  38. Y. Li, M. Cao, D. Wang and J. Yuan, High-efficiency and dynamic stable electromagnetic wave attenuation for La doped bismuth ferrite at elevated temperature and gigahertz frequency, RSC Adv., 2015, 5(94), 77184–77191 RSC.
  39. S. C. Yang, A. Kumar, V. Petkov and S. Priya, Room-temperature magnetoelectric coupling in single-phase BaTiO3-BiFeO3 system, J. Appl. Phys., 2013, 113, 144101 CrossRef.
  40. H. H. Zhang, W. Jo, K. Wang and K. G. Webber, Compositional dependence of dielectric and ferroelectric properties in BiFeO3-BaTiO3 solid solutions, Ceram. Int., 2014, 40, 4759–4765 CrossRef CAS.
  41. C. A. Randall, N. Kim, J. P. Kucera, W. W. Cao and T. R. Shrout, Intrinsic and Extrinsic Size Effects in Fine-Grained Morphotropic-Phase-Boundary Lead Zirconate Titanate Ceramics, J. Am. Ceram. Soc., 1998, 81(3), 677–788 CrossRef CAS.
  42. Y. Hiruma, H. Nagata and T. Takenaka, Thermal depoling process and piezoelectric properties of bismuth sodium titanate ceramics, J. Appl. Phys., 2009, 105, 084112 CrossRef.
  43. J. G. Chen and J. R. Cheng, Enhanced thermal stability of lead-free high temperature 0.75BiFeO3-0.25BaTiO3 ceramics with excess Bi content, J. Alloys Compd., 2014, 589, 115–119 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.