Next Article in Journal
Design and Fabrication of Double-Layer Crossed Si Microchannel Structure
Next Article in Special Issue
Single-Beam Acoustic Tweezer Prepared by Lead-Free KNN-Based Textured Ceramics
Previous Article in Journal
Ultrafast Deep-Ultraviolet Laser-Induced Voltage Response of Pyrite
Previous Article in Special Issue
Evolvable Acoustic Field Generated by a Transducer with 3D-Printed Fresnel Lens
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation, Structure, and Electrical Properties of Cobalt-Modified Bi(Sc3/4In1/4)O3–PbTiO3–Pb(Mg1/3Nb2/3)O3 High-Temperature Piezoelectric Ceramics

1
Institute of Materials, China Academy of Engineering Physics, Mianyang 621907, China
2
School of Microelectronics, Xidian University, Xi’an 710071, China
*
Authors to whom correspondence should be addressed.
Micromachines 2021, 12(12), 1556; https://doi.org/10.3390/mi12121556
Submission received: 6 November 2021 / Revised: 3 December 2021 / Accepted: 12 December 2021 / Published: 13 December 2021
(This article belongs to the Special Issue Advances in Piezoelectric Sensors, Transducers and Harvesters)

Abstract

:
Cobalt-modified 0.40Bi(Sc3/4In1/4)O3–0.58PbTiO3–0.02Pb(Mg1/3Nb2/3)O3 ceramics (abbreviated as BSI–PT–PMN–xCo) were produced by conventional two-step solid-state processing. The phase structure, micro structure morphology, and electrical properties of BSI–PT–PMN–xCo were systematically studied. The introduction of Co ions exerted a significant influence on the structure and electrical properties. The experiment results demonstrated that Co ions entered the B-sites of the lattice, resulting in slight lattice distortion and a smaller lattice constant. The average grain size increased from ~1.94 μm to ~2.68 μm with the increasing Co content. The optimized comprehensive electrical properties were obtained with proper Co-modified content 0.2 wt.%. The Curie temperature (Tc) was 412 °C, the piezoelectric constant (d33) was 370 pC/N, the remnant polarization (Pr) was 29.2 μC/cm2, the relatively dielectric constant (εr) was 1450, the planar electromechanical coupling coefficient (kp) was 46.5, and the dielectric loss (tanδ) was 0.051. Together with the enhanced DC resistivity of 109 Ω cm under 300 °C and good thermal stability, BSI–PT–PMN–0.2Co ceramic is a promising candidate material for high-temperature piezoelectric applications.

Graphical Abstract

1. Introduction

Lead zirconate titanate (PZT) systems represent most of the market share of piezoelectric materials because of their excellent piezoelectric performance, simple preparation process, and low cost [1,2,3]. However, their relatively low Curie temperature (Tc), which means their piezoelectric properties decrease quickly with rising environmental temperatures, restrict their high-temperature application in petrochemical, aerospace, and other industries [4,5,6].
In 2001, Eitel et al. [7] first reported a new high-temperature piezoelectric material based on (1−x)Bi(Me)O3xPbTiO3, where Me3+ refers to Sc3+, Y3+, In3+, Yb3+, etc. Preferably, (1−x)BiScO3xPbTiO3 (BS–PT) ceramics exhibit a morphotropic phase boundary (MPB) at x = 0.64, with a high Curie temperature of 450 °C and excellent piezoelectric performance (d33 = 460 pC/N), comparable to commercial soft PZT ceramics [8]. Later, extensive studies, such as the single-element doping [9,10,11,12,13,14,15], composition modification [16,17,18,19,20,21,22], and the introduction of a third component [23,24,25,26,27,28,29,30], focused on improving the piezoelectric, dielectric or mechanical properties of BS–PT ceramics were reported. For example, single-element doping refers to the substitution of Sc3+ by other +3 valence cations, such as In3+, Co3+, et al. Unfortunately, although the Tc is still over 400 °C, the piezoelectric properties decrease significantly. For example, for the In3+-modified, BS–PT based solid solutions Bi(Sc3/4In1/4)O3–PbTiO3 (BSI–PT), the Tc reaches 457 °C, whereas the d33 reduces to only 201 pC/N [16] for the MPB composition. The composition modification refers to the substitution of Sc3+ by composition ions such as (Mg1/2Ti1/2)3+, (Ni1/2Ti1/2)3+, (Zn1/2Zr1/2)3+, (Ni1/2Zr1/2)3+, etc. With proper composition ion modification, obtained improved piezoelectric properties can be obtained, together with considerable Tc. For example, BS–PT–0.025Bi(Ni1/2Zr1/2)O3 solid solutions exhibit d33 of 480 pC/N and Tc 439 °C for MPB composition [20]. Another way to modify the electrical properties of the BS–PT binary system is to introduce a third component to form BS–PT–ABO3 ternary solid solutions. The electrical properties can be regulated due to the different chemical composition of ABO3. For example, the piezoelectric properties (d33 = 555 pC/N, kp = 59%) and Curie temperature, of 408 °C, were obtained for the MPB composition in the BS–PT–Pb(Sn1/3Nb2/3)O3 system [27]. Furthermore, the BS–PT–LiNbO3 system possesses an enhanced d33 of 551 pC/N and reduced Tc of 337 °C in the vicinity of the MPB [29]. The BS–PT–BiGaO3 system exhibits an enhanced Tc of 511 °C and a reduced d33 of 102 pC/N [30].
In our previous study, we combined the first method and the third method to form a new ternary system Bi(Sc3/4In1/4)O3–PbTiO3–Pb(Mg1/3Nb2/3)O3 to modify the electrical properties of the BS–PT solid solution [31]. The optimal properties were obtained for the (0.98-x)BSI–xPT–0.02PMN ceramics at the MPB x = 0.58 of piezoelectric constant (d33) 403 pC/N, planar electromechanical coupling factor (kp) 47.2%, and remnant polarization (Pr) 36.4 μC/cm2. The Curie temperature (Tc) remained at 421°C, making the system much more suitable for high-temperature piezoelectric applications.
Furthermore, high electrical resistance is so necessary for high-temperature application that a large electric field can be applied during poling without breakdown or excessive charge leakage. As a common doping element in various piezoelectric material systems, cobalt is often used to improve the electrical properties of piezoelectric materials [32,33,34,35,36,37]. Co ion is commonly used as acceptor-type dopant to replace the B-site ion in perovskite materials. Generally, in the case of B-site substitution, the improvement of piezoelectric properties is not as good as with A-site substitution, but it can improve the high temperature resistivity greatly while maintaining the Curie temperature. In this study, we attempt to select a Co ion substitution for 0.40Bi(Sc3/4In1/4)O3–0.58PbTiO3–0.02Pb(Mg1/3Nb2/3)O3 ternary ceramics, aiming to generate good electrical performance while still providing a relatively high Curie temperature.

2. Experimental Procedure

The 0.40Bi(Sc3/4In1/4)O3–0.58PbTiO3–0.02Pb(Mg1/3Nb2/3)O3xCo (x = 0, 0.2, 0.4, 0.6, 0.8 wt.%) ceramics were produced through conventional two-step solid state processing. The reagent-grade materials of Bi2O3 (99.8%), Sc2O3 (99.99%), In2O3 (99.99%), PbO (99%), TiO2 (99.5%), MgO (99%), Nb2O5 (99.9%), and Co2O3 (99.9%) were utilized as the raw materials and weighted by the stoichiometric amount. First, the powder MgO and Nb2O5 were mixed and calcined at 1100 °C for 4 h to synthesize the columbite precursor MgNb2O6. Next, the precursor MgNb2O6 and other raw materials were mixed together for 12 h by ball milling with alcohol. The powders were then dried, ground, granulated by milling with polyvinyl alcohol (PVA) as the binder, and then pressed into tablets with a diameter of 12.5 mm and a thickness of 1.5 mm under 120 MPa of pressure. Before sintering, a cold isostatic pressing process of 200 MPa was carried out on the tablets for 20 min. After burning out PVA at 650 °C, the tablets were sintered at 1120 °C for 3 h in a muffle furnace.
The X-ray diffraction (XRD) patterns of the BSI–PT–PMN samples were obtained by using an X-ray diffractometer (Cu Kα, λ = 1.5406, D/Max 2500, Rigaku, Tokushima, Japan). The fresh fracture surface microstructure of the samples was observed by using scanning electron microscopy (SEM) (S–4800, Hitachi, Tokyo, Japan). After polishing and ultrasonic cleaning, the ceramic samples were evenly coated with silver slurry and then fired at 600 °C. After cooling to room temperature, the ceramic samples were poled at 120 °C in a silicone oil bath under a DC electric field of 4.5 kV/mm for 20 min. The dielectric behavior as a function of temperature was analyzed using an impedance analyzer (4294A, Agilent, Palo Alto, California, USA). The piezoelectric coefficient d33 was measured using a quasi-static type meter (ZJ–3A, Institute of Acoustics, Beijing, China). The P-E hysteresis loops of the samples were characterized using a ferroelectric analyzer (TF1000, aixACCT, Aachen, Germany). The temperature dependence of the DC resistivity was measured by a Digit Multimeter (34410A 6½, Agilent, Palo Alto, CA, USA). When it was necessary to measure the thermal stability, the samples were heated to the set temperature and held for an hour. The piezoelectric coefficient d33 was measured when the samples cooled down to room temperature.

3. Results and Discussion

Figure 1 presents the X-ray diffraction patterns of BSI–PT–PMN–xCo (x = 0, 0.2 0.4, 0.6, 0.8 wt.%). From the figure, one can clearly observe that all the Co-doped ceramic samples presented the perovskite phase with a secondary phase of Co2O3 (Index PDF#02-0770), indicating that the Co modification did not change the crystal structure of BSI–PT–PMN ceramics significantly. Furthermore the characteristic peak moved slightly, to a high angle, as the Co content increased, which can be seen in Figure 1b. This phenomenon indicates the decrease in the lattice constant in this system. According to the Bragg equation (2dsinθ = ), the lattice constant was calculated and listed in Table 1. It can be seen that the lattice parameter a changed slightly and the lattice parameter c decreased with the increasing Co content x. It has been reported that Co ions exist in the valence state as Co2+ if the sintered temperature exceeds 1080 °C [38]. On account of the fact that the samples were sintered at 1120 °C in this study, the Co3+ and Co2+ were coexistent for Co2O3 used as dopants in this work. Furthermore, the radii of Co3+ and Co2+ were 0.63 Å and 0.72 Å, respectively, smaller than the radius of Ti4+ of 0.86 Å; Co3+ or Co2+ may have entered the lattice to replace Ti4+ in the B-sites for the BSI–PT–PMN ceramics and resulted in the decrease in the lattice constant. In addition, the c/a ratio decreased from 1.022 to 1.017 with the increasing Co content x, which exerted a significant effect on the electric properties.
Figure 2 presents the microstructure of the faces for BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8 wt %) sintered at 1120 °C. All the samples exhibited a dense structure with no obvious porosity, in accordance with the high relative density listed in Table 1. The average grain sizes of the BSI–PT–PMN–xCo samples were statistically analyzed and listed in in Table 1 using Nano Measurer software (V1.2, developed by Visual Basic 6.0 (Microsoft, Redmond, WA, USA), free from the Internet), with the number of the grains over 50. It was found that the average grain size increased from ~1.94 μm to ~2.68 μm as the Co content increased, suggesting that the doping of Co ions can reduce the sintering temperature and facilitate grain growth, which has been reported in other Co-doped piezoelectric ceramics [39].
Figure 3 depicts the curves of the temperature dependence of the dielectric permittivity (ε) for the BSI–PT–PMN–xCo ceramics. It can be seen that the dielectric peak of the BSI–PT–PMN–xCo ceramics became suppressed and broader when the measured frequency increased, and there was no frequency dispersion in any of the samples. The spectral line of the dielectric permittivity featured only one peak, which corresponded to the Curie temperature.
The Curie temperature ranged from 421 °C to 398 °C with increasing Co ion-doped content, which can be attributed to the enlarged tolerance factor ‘t’ caused by the B-site substitution. For the perovskite ferroelectric structure, the tolerance factor ‘t’ can be described as
t = r A + r O 2 ( r B + r O )
where rA, rB, and rO refer to the ionic radii of the A-site, B-site, and oxygen atom, respectively. As is already known, for the perovskite ferroelectric structure, there is a consistent relationship between the Curie temperature and the tolerance factor ‘t’: the lower tolerance factor ‘t’, the higher the Curie temperature. As discussed above, the introduced Co3+ and Co2+ replaced Ti4+ and entered into B-sites for BSI–PT–PMN ceramics. As a result, the tolerance factor ‘t’ enlarged and the Curie temperature declined due to the reduced B-sites effective ionic radius. Furthermore, typical dielectric diffuse phenomena can be found in this system, which can be described by the modified Curie–Weiss law as
1 ε 1 ε m = ( T T m ) γ C
where εm is the maximum value of the dielectric permittivity ε, Tm is the temperature of εm, C is the Curie–Weiss constant, and γ is the diffuseness coefficient value between 1 to 2, respectively. The values of γ of all the BSI–PT–PMN–xCo ceramics were calculated and presented in Figure 3. In this system, the dielectric diffuseness may be mainly attributed to the polar nanodomains generated by the relaxor PMN third component (0.02Pb(Mg1/3Nb2/3)O3 ), due to the slight change in the values of γ in line with the increasing Co content.
Figure 4 presents the PE hysteresis loop behavior for the BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8) ceramics at room temperature under 1 Hz. The samples featured better saturated PE loops when the Co ion-doped content increased. For the undoped BSI–PT–PMN ceramic, Pr was 36.4 μC/cm2 and Ec was 24.7 kV/cm. By doping the Co ions, Pr decreased and reached a minimum value of 25.1 μC/cm2, while Ec increased and reached a maximum value of 32.3 kV/cm at x = 0.8. This phenomenon may have been due to the Co ions entering the lattice to replace the Ti ions in the B-sites, resulting in oxygen vacancy, which plays a role in domain wall pinning.
Figure 5 presents the strain (S) versus the electric field (E) loop by applying an electric field of 40 kV/cm and 1 Hz of BSI–PT–PMN–xCo ceramics. The strain under a unipolar electric field of 40 kV/cm was 0.222, 0.205, 0.198, 0.188, and 0.182, with 0, 0.2%, 0.4%, 0.6% and 0.8% Co ion doping, respectively.
Furthermore, the relevant large-signal d33* and strain hysteresis h were calculated and listed in Table 2, where the strain hysteresis h can be described as
h = Δ S S max × 100 %
where ΔS is the strain difference at half the maximal driven electric field in a cycle and the Smax is the strain maximum. It was found the large-signal d33* were over 450 pm/V for all these BSI–PT–PMN–xCo ceramics and the strain curve exhibited a good linearity with small hysteresis h, which is advantageous for piezoelectric actuator applications.
The electrical properties of the Co ion-doped BSI–PT–PMN ceramics at room temperature are listed in Table 2. It can be seen that the Curie temperature (Tc), relatively dielectric constant (εr), piezoelectric constant (d33), remnant polarization (Pr), planar electromechanical coupling coefficient (kp), and dielectric loss (tanδ) monotonically decreased, while the coercive electric field (Ec) monotonically increased, when the Co ion content increased from 0 to 0.8 wt.%.
Figure 6 presents the Arrhenius-type plots of conductivity as a function of temperature for BSI–PT–PMN with undoped and 0.2% doped Co ion. The activation energy Eac was calculated according to the Arrhenius law: σ = σ0exp(-Eac/kT), where σ, Eac, k, and T are conductivity, activation energy, Boltzmann constant, and temperature, respectively. The activation energy was 0.3 eV at 25 < T < 220 °C and 1.5 eV at 220 < T < 400 °C for the undoped BSI–PT–PMN, while the activation energy was 0.33 eV at 25 < T < 220 °C and 1.43 eV at 220 < T < 400 °C for 0.2% Co-doped BSI–PT–PMN. The value of 0.2% Co ion-doped BSI–PT–PMN ceramic was found to be slightly larger than the value of the undoped BSI–PT–PMN ceramic at high temperature.
It has been reported that for the BS–PT system, the activation energy Eac for ionic conduction is to be ≥1 eV and for electronic conduction around 0.1 eV [40,41,42]. This indicates that ionic conduction plays a more important role than electronic conduction when T > 220 °C. In this study, Co ions were introduced into the BSI–PT–PMN system. The employment of Co ions may reduce the Pb leakage loss during calcination and sintering by trapping the bonded electrons and increase the resistivity. As a result, the 0.2% Co ion-doped BSI–PT–PMN ceramic exhibited a lower activation energy at 220 < T < 400 °C.
In addition, as can be seen in the inset of Figure 6, it was observed that the 0.2% Co ion-doped BSI–PT–PMN ceramic exhibited a higher resistivity value than the undoped BSI–PT–PMN ceramic (109 Ω magnitude to108 Ω magnitude) at 300 °C, which is advantageous for high-temperature actuator applications.
Figure 7 depicts the piezoelectric constant d33 as a function of temperature for BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8). All the piezoelectric properties were measured at room temperature after annealing the poled specimens at different temperatures for 1 h. All the samples displayed the same trend and the piezoelectric constant d33 remained almost unchanged from room temperature to 250 °C; it began to drop when the annealing temperature exceeded 250 °C, dropped quickly when the annealing temperature exceeded 300 °C, and finally dropped to zero near the Curie temperature point. For the ceramic with 0.2 wt.% Co ion doping, the value of the piezoelectric constant d33 at 325 °C could still reach 80% of room temperature. The results reveal that BSI–PT–PMN–0.2Co ceramics possesses better thermal stability than ceramics without Co ion doping. The harder domain switching [43] may explain this phenomenon.

4. Conclusions

High-temperature piezoelectric ceramic 0.40Bi(Sc3/4In1/4)O3–0.58PbTiO3–0.02Pb(Mg1/3Nb2/3)O3 (abbreviated as BSI–PT–PMN) doped with different amounts of Co ions (x = 0, 0.2, 0.4, 0.6, 0.8 wt.%) were prepared. The influences of Co ion doping on the phase transition, micro structure morphology, and electrical properties of the prepared BSI–PT–PMN–xCo ceramics were studied. It was found that Co ions entered the B-sites of the lattice, resulting in slight lattice distortion and a smaller lattice constant. With the increase in the doping of the Co ions, the grain size became larger; Co ions can promote the grain growth. The high temperature resistivity and temperature stability of the BSI–PT–PMN ceramics can be improved through an appropriate amount of Co ion doping. When the Co ion-doped content was 0.2 wt.%, the BSI–PT–PMN exhibited optimized comprehensive electrical properties: the Curie temperature (Tc) was 412 °C, the piezoelectric constant (d33) was 370 pC/N, the remnant polarization (Pr) was 29.2 μC/cm2, the relatively dielectric constant (εr) was 1450, the planar electromechanical coupling coefficient (kp) was 46.5, and the dielectric loss (tanδ) was 0.051. BSI–PT–PMN–0.2Co ceramics feature the merits of relatively high Curie temperature, high resistivity at high temperature, large strain, good strain linearity, and good thermal stability. BSI–PT–PMN–0.2Co ceramics are suitable for high-temperature actuator applications.

Author Contributions

Conceptualization, Z.C. and T.Z.; methodology, Z.Y. and J.Z.; software, T.Z.; validation, Z.Y., and J.Z.; formal analysis, X.S.; investigation, Z.C. and K.S.; resources, T.Z.; data curation, N.L.; writing—original draft preparation, Z.C.; writing—review and editing, T.Z.; visualization, N.L. and J.Z; supervision, X.S. and B.G.; project administration, Z.C. and N.L.; funding acquisition, B.G. and T.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundations of China (Grant No. 51802242), the China Postdoctoral Science Foundation (Grant No. 2019M663927XB), and the Foundation from Institute of Materials, China Academy of Engineering Physics (Grant No. TP02201612).

Data Availability Statement

The data used to support the findings of this study are available from the corresponding author upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jaffe, B.; Roth, R.S.; Marzullo, S. Piezoelectric properties of lead zirconate-lead titanate solid-solution ceramics. J. Appl. Phys. 1954, 25, 809–810. [Google Scholar] [CrossRef]
  2. Tressler, J.F.; Alkoy, S.; Newnham, R.E. Piezoelectric sensors and sensor materials. J. Electroceram. 1998, 2, 257–272. [Google Scholar] [CrossRef]
  3. Dong, S.X. Review on piezoelectric, ultrasonic, and magnetoelectric actuators. J. Adv. Dielect. 2012, 2, 1230001. [Google Scholar] [CrossRef]
  4. Gubinyi, Z.; Batur, C.; Sayir, A.; Dynys, F. Electrical properties of PZT piezoelectric ceramics at high temperature. J. Electroceram. 2008, 20, 95–105. [Google Scholar] [CrossRef]
  5. Qureshi, A.H.; Shabbir, G.; Hall, D.A. On the synthesis and dielectric studies of (1−x)Bi(Mg1/2Zr1/2)O3xPbTiO3 piezoelectric ceramic system. Mater. Lett. 2007, 61, 4482–4484. [Google Scholar] [CrossRef]
  6. Zhang, S.J.; Yu, F.P. Piezoelectric materials for high temperature sensors. J. Am. Ceram. Soc. 2011, 94, 3153–3170. [Google Scholar] [CrossRef]
  7. Eitel, R.E.; Randall, C.A.; Shrout, T.R.; Rehrig, P.W.; Hackenberger, W.; Park, S.E. New high temperature morphotropic phase boundary piezoelectrics based on Bi(Me)O3-PbTiO3 ceramics. Jpn. J. Appl. Phys. 2001, 40, 5999–6002. [Google Scholar] [CrossRef]
  8. Eitel, R.E.; Randall, C.A.; Shrout, T.R.; Park, S.E. Preparation and characterization of high temperature perovskite ferroelectrics in the solid-solution (1−x)BiScO3xPbTiO3. Jpn. J. Appl. Phys. 2002, 41, 2099. [Google Scholar] [CrossRef]
  9. Iniguez, J.; Vanderbilt, D.; Bellaiche, L. First-principles study of (BiScO3)1−x-(PbTiO3)x piezoelectric alloys. Phys. Rev. B 2003, 67, 224107. [Google Scholar] [CrossRef] [Green Version]
  10. Inaguma, Y.; Miyaguchi, A.; Yoshida, M.; Katsumata, T.; Shimojo, Y.; Wang, R.; Sekiya, T. High-pressure synthesis and ferroelectric properties in perovskite-type BiScO3-PbTiO3 solid solution. J. Appl. Phys. 2004, 95, 231–235. [Google Scholar] [CrossRef]
  11. Winotai, P.; Udomkan, N.; Meejoo, S. Piezoelectric properties of Fe2O3-doped (1−x)BiScO3xPbTiO3 ceramics. Sens. Actuators A 2005, 122, 257–263. [Google Scholar] [CrossRef]
  12. Chen, J.G.; Shi, H.D.; Liu, G.X.; Cheng, J.R.; Dong, S.X. Temperature dependence of dielectric, piezoelectric and elastic properties of BiScO3-PbTiO3 high temperature ceramics with morphotropic phase boundary (MPB) composition. J. Alloy. Compd. 2012, 537, 280–285. [Google Scholar] [CrossRef]
  13. Chen, J.G.; Cheng, J.R.; Dong, S.X. Review on high temperature piezoelectric ceramics and actuators based on BiScO3-PbTiO3 solid solutions. J. Adv. Dielect. 2014, 4, 1430002. [Google Scholar] [CrossRef] [Green Version]
  14. Meng, X.F.; Chen, Q.; Fu, H.; Liu, H.; Zhu, J.G. Influence of PNN on the structure and electronic properties of BSPT ceramics. J. Mater. Sci. Mater. Electron. 2018, 29, 12785–12794. [Google Scholar] [CrossRef]
  15. Wu, J.G.; Gao, X.Y.; Yu, Y.; Yang, J.K.; Chu, Z.Q.; Bokov, A.A.; Ye, Z.G.; Dong, S.X. Quantitative studies of domain evolution in tetragonal BS-PT ceramics in electric poling and thermal depoling processes. J. Mater. Chem. C 2019, 7, 4517. [Google Scholar] [CrossRef]
  16. Hu, Z.Q.; Chen, J.G.; Li, M.Y.; Li, X.T.; Liu, G.X.; Dong, S.X. Morphotropic phase boundary and high temperature dielectric, piezoelectric, and ferroelectric properties of (1−x)Bi(Sc3/4In1/4)O3xPbTiO3 ceramics. J. Appl. Phys. 2011, 110, 064102. [Google Scholar] [CrossRef]
  17. Ansell, T.Y.; Cann, D.P. High temperature piezoelectric ceramics based on (1−x)[BiScO3+Bi(Ni1/2Ti1/2)O3]−xPbTiO3. Mater. Lett. 2012, 80, 87–90. [Google Scholar] [CrossRef]
  18. Sehirlioglu, A.; Sayir, A.; Dynys, F. Microstructure-property relationships in liquid phase-sintered high-temperature bismuth scandium oxide-lead titanate piezoceramics. J. Am. Ceram. Soc. 2008, 91, 2910–2916. [Google Scholar] [CrossRef]
  19. Zhang, S.J.; Eitel, R.E.; Randall, C.A.; Shrout, T.R.; Alberta, E.F. Manganese-modified BiScO3-PbTiO3 piezoelectric ceramic for high-temperature shear mode sensor. Appl. Phys. Lett. 2005, 86, 262904. [Google Scholar] [CrossRef]
  20. Zhao, T.-L.; Fei, C.; Pu, K.; Dai, X.; Song, J.; Wang, C.-M.; Dong, S. Structure evolution and enhanced electrical performance for BiScO3–Bi(Ni1/2Zr1/2)O3–PbTiO3 solid solutions near the morphotropic phase boundary. J. Alloy. Compd. 2021, 873, 159844. [Google Scholar] [CrossRef]
  21. Cheng, J.R.; Eitel, R.E.; Li, N.; Cross, L.E. Structural and electrical properties of (1−x)Bi(Ga1/4Sc3/4)O3xPbTiO3 piezoelectric ceramics. J. Appl. Phys. 2003, 94, 605–609. [Google Scholar] [CrossRef]
  22. Zhao, T.L.; Wang, C.M.; Wang, C.L.; Wang, Y.M.; Dong, S.X. Enhanced piezoelectric properties and excellent thermal stabilities of cobalt-modified Aurivillius-type calcium bismuth titanate (CaBi4Ti4O15). Mater. Sci. Eng. B 2015, 201, 51–56. [Google Scholar] [CrossRef]
  23. Ryu, J.; Priya, S.; Sakaki, C.; Uchino, K. High power piezoelectric characteristics of BiScO3-PbTiO3-Pb(Mn1/3Nb2/3)O3. Jpn. J. Appl. Phys. 2002, 41, 6040–6044. [Google Scholar] [CrossRef]
  24. Stringer, C.J.; Randall, C.A. In situ TEM investigations of the high-temperature relaxor ferroelectric BiScO3-Pb(Mg1/3Nb2/3)O3-PbTiO3 ternary solid solution. J. Am. Ceram. Soc. 2007, 90, 1802–1807. [Google Scholar] [CrossRef]
  25. Yao, Z.H.; Liu, H.X.; Hao, H.; Cao, M.H. Structure, electrical properties, and depoling mechanism of BiScO3-PbTiO3-Pb(Zn1/3Nb2/3)O3 high-temperature piezoelectric ceramics. J. Appl. Phys. 2011, 109, 014105. [Google Scholar] [CrossRef]
  26. Chen, J.G.; Zhao, T.L.; Cheng, J.R.; Dong, S.X. Enhanced piezoelectric performance of (0.98-x)Bi(Sc3/4In1/4)O3xPbTiO3-0.02Pb(Zn1/3Nb2/3)O3 ternary high temperature piezoelectric ceramics. J. Appl. Phys. 2013, 113, 144102. [Google Scholar] [CrossRef]
  27. Zhao, T.L.; Bokov, A.A.; Wu, J.G.; Wang, H.L.; Wang, C.M.; Yu, Y.; Wang, C.L.; Zeng, K.Y.; Ye, Z.G.; Dong, S.X. Giant piezoelectricity of ternary perovskite ceramics at high temperature. Adv. Funct. Mater. 2019, 29, 1807920. [Google Scholar] [CrossRef]
  28. Kang, H.J.; Chen, J.; Liu, L.J.; Fang, L.; Xing, X.R. Temperature dependences of the ferroelectric and dielectric properties of high Curie temperature PbTiO3-BiScO3-Bi(Zn1/2Zr1/2)O3. Mater. Res. Bull. 2013, 48, 2006–2009. [Google Scholar] [CrossRef]
  29. Hu, Q.R.; Wang, Y.P.; Wu, L.; Yin, J.; Chen, L.; Yuan, G.L.; Yang, Y. Effects of LiNbO3 doping on the microstructures and electrical properties of BiScO3-PbTiO3 piezoelectric system. J. Mater. Sci. Mater. Electron. 2018, 29, 18036–18044. [Google Scholar] [CrossRef]
  30. Jiang, Y.H.; Qin, B.Q.; Zhao, Y.; Jiang, Y.Z.; Shi, W.; Li, Q.S.; Xiao, D.Q.; Zhu, J.G. Phase transition, piezoelectric properties, and thermal stability of (1−xy)BiScO3yBiGaO3xPbTiO3 ceramics. J. Am. Ceram. Soc. 2008, 91, 2943–2946. [Google Scholar] [CrossRef]
  31. Zhao, T.L.; Wang, C.M.; Chen, J.G.; Wang, C.L.; Dong, S.X. Phase transitional behavior and enhanced electrical properties of Bi(Sc3/4In1/4)O3-PbTiO3 by small content Pb(Mg1/3Nb2/3)O3 modification. J. Mater. Sci. Mater. Electron. 2016, 27, 606–612. [Google Scholar] [CrossRef]
  32. Jaimeewong, P.; Sittinon, S.; Buntham, S.; Bomlai, P.; Namsar, O.; Pojprapai, S.; Watcharapasorn, A. Ferroelectric, piezoelectric and dielectric behaviors of CoO- and Fe2O3-doped BCZT ceramics. Phys. Status Solidi A 2018, 215, 1701023. [Google Scholar] [CrossRef]
  33. Huang, X.Y.; Gao, C.H.; Huang, M.S.; Yue, Z.X. Influence of Co ion doped amount on property of BCTZ piezoelectric ceramics sintered at low temperature. Ferroelectrics 2015, 481, 146–154. [Google Scholar] [CrossRef]
  34. Shi, S.F.; Hashimoto, H.; Sekino, T. Enhancing piezoelectric properties of Ba0.88Ca0.12Zr0.12Ti0.88O3 lead-free ceramics by doping Co ions. Ceram. Int. 2021, 47, 3272–3278. [Google Scholar] [CrossRef]
  35. Koronska, R.B.; Vasylechko, L.; Markiewicz, E.; Nalecz, D.M.; Kalvane, A. X-ray and dielectric characterization of Co doped tetragonal BaTiO3 ceramics. Phase Transit. 2017, 90, 78–85. [Google Scholar] [CrossRef]
  36. Ma, Y.N.; Chen, H.M.; Pan, F.C.; Chen, Z.P.; Ma, Z.; Lin, X.L.; Zheng, F.; Ma, X.B. Electronic structures and optical properties of Fe/Co–doped cubic BaTiO3 ceramics. Ceram. Int. 2019, 45, 6303–6311. [Google Scholar] [CrossRef]
  37. Peng, Z.H.; Zheng, D.Y.; Zhou, T.; Yang, L.; Zhang, N.; Fang, C. Effects of Co2O3 doping on electrical properties and dielectric relaxation of PMS-PNN-PZT ceramics. J. Mater. Sci. Mater. Electron. 2018, 29, 5961–5968. [Google Scholar] [CrossRef]
  38. Guo, Z.L.; Wang, C.M.; Zhao, T.L.; Yu, S.L.; Cao, Z.P. Piezoelectric properties and thermal stabilities of cobalt-modified potassium bismuth titanate. Mater. Chem. Phys. 2013, 140, 260–265. [Google Scholar] [CrossRef]
  39. Wang, C.-M.; Wang, J.-F.; Zheng, L.-M.; Zhao, M.-L.; Wang, C.-L. Enhancement of the piezoelectric properties of sodium lanthanum bismuth titanate (Na0.5La0.5Bi4Ti4O15) through modification with cobalt. Mat. Sci. Eng. B 2010, 171, 79–85. [Google Scholar] [CrossRef]
  40. Zhang, S.J.; Randall, C.A.; Shrout, T.R. High Curie temperature piezocrystals in the BiScO3-PbTiO3 perovskite system. Appl. Phys. Lett. 2003, 83, 3150–3152. [Google Scholar] [CrossRef]
  41. Kobor, D.; Guiffard, B.; Lebrun, L.; Hajjaji, A.; Guyomar, D. Oxygen vacancies effect on ionic conductivity and relaxation phenomenon in undoped and Mn doped PZN-4.5PT single crystals. J. Phys. D Appl. Phys. 2007, 40, 2920–2926. [Google Scholar] [CrossRef]
  42. James, A.R.; Priya, S.; Uchino, K.; Srinivas, K.; Kiran, V.V. Investigations of intrinsic defect structure in 0.91Pb(Zn1/3Nb2/3)O3-0.09PbTiO3 single crystal through AC conductivity. Jpn. J. Appl. Phys. 2002, 41, 5272–5276. [Google Scholar] [CrossRef]
  43. Randall, C.A.; Kim, N.; Kucera, J.P.; Cao, W.W.; Shrout, T.R. Intrinsic and extrinsic size effects in fine-grained morphotropic-phae-boundary lead zirconate titanate ceramics. J. Am. Ceram. Soc. 1998, 81, 677. [Google Scholar] [CrossRef]
Figure 1. (a)X-ray diffraction pattern of BSI–PT–PMN–xCo Ceramics (Cobolt modified 0.40Bi(Sc3/4In1/4)O3–0.58PbTiO3–0.02Pb(Mg1/3Nb2/3)O3 ), (b) The variation trend of the selected peak (1 1 0). Note: α refers to Co2O3.
Figure 1. (a)X-ray diffraction pattern of BSI–PT–PMN–xCo Ceramics (Cobolt modified 0.40Bi(Sc3/4In1/4)O3–0.58PbTiO3–0.02Pb(Mg1/3Nb2/3)O3 ), (b) The variation trend of the selected peak (1 1 0). Note: α refers to Co2O3.
Micromachines 12 01556 g001
Figure 2. Microstructure of faces for BSI–PT–PMN–xCo, (a) x = 0, (b) x = 0.2, (c) x = 0.4, (d) x = 0.6, (e) x = 0.8.
Figure 2. Microstructure of faces for BSI–PT–PMN–xCo, (a) x = 0, (b) x = 0.2, (c) x = 0.4, (d) x = 0.6, (e) x = 0.8.
Micromachines 12 01556 g002
Figure 3. Dielectric permittivity (at 1 kHz, 10 kHz, and 100 kHz) as a function of temperature for BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8).
Figure 3. Dielectric permittivity (at 1 kHz, 10 kHz, and 100 kHz) as a function of temperature for BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8).
Micromachines 12 01556 g003
Figure 4. Bipolar field induced polarization for BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8).
Figure 4. Bipolar field induced polarization for BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8).
Micromachines 12 01556 g004
Figure 5. Unipolar field induced strain for BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8).
Figure 5. Unipolar field induced strain for BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8).
Micromachines 12 01556 g005
Figure 6. Arrhenius-type plots of conductivity σ as a function of temperature for BSI–PT–PMN with undoped and 0.2% doped Co ion. Inset depicting the resistivity ρ as a function of temperature.
Figure 6. Arrhenius-type plots of conductivity σ as a function of temperature for BSI–PT–PMN with undoped and 0.2% doped Co ion. Inset depicting the resistivity ρ as a function of temperature.
Micromachines 12 01556 g006
Figure 7. Piezoelectric constant d33 as a function of temperature for BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8).
Figure 7. Piezoelectric constant d33 as a function of temperature for BSI–PT–PMN–xCo (x = 0, 0.2, 0.4, 0.6, 0.8).
Micromachines 12 01556 g007
Table 1. Structure parameter of BSI–PT–PMN–xCo ceramics.
Table 1. Structure parameter of BSI–PT–PMN–xCo ceramics.
Materiala
(Å)
c
(Å)
c/aRelative Density (%)Grain Size
(μm)
x = 0.03.9204.0051.02296.21.94
x = 0.23.9214.0041.02196.32.32
x = 0.43.9224.0031.02097.12.41
x = 0.63.9223.9981.01995.32.53
x = 0.83.9213.9891.01795.62.68
Table 2. Room temperature electrical properties of Co ion-doped BSI–PT–PMN ceramics.
Table 2. Room temperature electrical properties of Co ion-doped BSI–PT–PMN ceramics.
MaterialTc
(°C)
εrtanδd33
(pC/N)
kp
(%)
Ec
(kV/cm)
Pr
(μC/cm2)
Strain
(%)
d33*
(pm/V)
h
(%)
x = 0.042116850.05540347.224.736.40.22255517.8
x = 0.241214500.05137046.527.129.20.205512.516.0
x = 0.440814330.04836045.728.628.40.19849514.5
x = 0.640313860.04335043.830.426.30.18847015.4
x = 0.839813310.04433041.632.325.10.18245514.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, Z.; Lin, N.; Yang, Z.; Zhang, J.; Shi, K.; Sun, X.; Gao, B.; Zhao, T. Preparation, Structure, and Electrical Properties of Cobalt-Modified Bi(Sc3/4In1/4)O3–PbTiO3–Pb(Mg1/3Nb2/3)O3 High-Temperature Piezoelectric Ceramics. Micromachines 2021, 12, 1556. https://doi.org/10.3390/mi12121556

AMA Style

Chen Z, Lin N, Yang Z, Zhang J, Shi K, Sun X, Gao B, Zhao T. Preparation, Structure, and Electrical Properties of Cobalt-Modified Bi(Sc3/4In1/4)O3–PbTiO3–Pb(Mg1/3Nb2/3)O3 High-Temperature Piezoelectric Ceramics. Micromachines. 2021; 12(12):1556. https://doi.org/10.3390/mi12121556

Chicago/Turabian Style

Chen, Zhijiang, Na Lin, Zhao Yang, Juan Zhang, Kefei Shi, Xinhao Sun, Bo Gao, and Tianlong Zhao. 2021. "Preparation, Structure, and Electrical Properties of Cobalt-Modified Bi(Sc3/4In1/4)O3–PbTiO3–Pb(Mg1/3Nb2/3)O3 High-Temperature Piezoelectric Ceramics" Micromachines 12, no. 12: 1556. https://doi.org/10.3390/mi12121556

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

Article Metrics

Back to TopTop