Next Article in Journal
Supramolecular Weaving by Halogen-Bonding in Functionality-Rich Hexasubstituted Aromatic Synthons
Next Article in Special Issue
Influence of the Addition of Rare Earth Elements on the Energy Storage and Optical Properties of Bi0.5Na0.5TiO3–0.06BaTiO3 Polycrystalline Thin Films
Previous Article in Journal
CO2-Switchable Hierarchically Porous Zirconium-Based MOF-Stabilized Pickering Emulsions for Recyclable Efficient Interfacial Catalysis
Previous Article in Special Issue
Microstructural Origin of the High-Energy Storage Performance in Epitaxial Lead-Free Ba(Zr0.2Ti0.8)O3 Thick Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Piezoelectric Response and Cycling Fatigue Resistance of Low-Temperature Sintered PZT-Based Ceramics

1
School of Material Science and Engineering, Wuhan University of Technology, Wuhan 430070, China
2
Sanya Science and Education Innovation Park of Wuhan University of Technology, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(4), 1679; https://doi.org/10.3390/ma16041679
Submission received: 9 January 2023 / Revised: 15 February 2023 / Accepted: 16 February 2023 / Published: 17 February 2023
(This article belongs to the Special Issue Preparation and Properties of Novel Energy Storage Materials)

Abstract

:
The preparation of low-cost multilayer piezoelectric devices requires using cheap internal electrodes between the dielectric layers. A general strategy is to reduce the sintering temperature Ts of the ceramic layer by sintering aids which can form a liquid phase. Here, 0.2 wt% Li2CO3 was added as a sintering aid to tailor the sinterability and piezoelectricity of the commercial PZT ceramics. As verified from experiments, the piezoelectric ceramics could be densified at a sintering temperature above 940 °C, suitable for co-firing with the cheap internal electrode. The optimized sintering temperature of 980 °C can be confirmed for the 0.2 wt% Li2CO3-modified PZT ceramics due to its high piezoelectric coefficient d33 ~ 701 pC/N, planar coupling factor kp ~ 66.7%, and a low mechanical quality factor Qm ~ 71 with a transition temperature of Tc ~ 226 °C, presenting the characteristics of typical soft piezoelectric ceramics. Moreover, both the potential piezoelectric strain ~0.13% under 20 kV/cm and the good cycling fatigue characteristic (>104 cycles) of the studied piezo compositions indicates strong competitiveness in the field of multilayer piezoelectric devices.

1. Introduction

Piezoceramics come from the general design of ferroelectrics due to the piezoelectric effect via electric-field-driven polarization treatment of polycrystals. These piezoelectric effects can generate an electro-response or micro-displacement response, which can be used for electronic components and devices for displacement and electrical-signal sensitivity [1,2,3]. Compared to electromagnetic actuators, piezoelectric ceramics mainly feature compact size, low driving voltage, quick response, high efficiency, accurate position, etc. [4,5,6,7].
Chemical doping dominates the structural design of piezoelectric materials to form donor-type (soft) or acceptor-type (hard) characteristics. Doping high-valence ions (e.g., La3+, Nd3+, Sm3+, Nb5+ and Ta5+) always cause lead vacancies, which is beneficial to domain rotation during electric field polarization, and further improve dielectric, electromechanical, and piezoelectric properties, similar to softening ceramics [8,9,10,11]. Soft piezoelectric ceramics are characterized by high direct piezoelectric constant d33, strain rate S, significant remnant polarization Pr, and low coercive field Ec. Typical soft Pb-based compositions by La doping, Pb1−xLax(ZryTi1−y)1−x/4O3 (abbreviated as PLZT) ceramics have been extensively studied due to their high electromechanical and electro-optical properties. La/Zr/Ti ratio can make the PLZT ceramics own various phase structures such as ferroelectric, relaxor ferroelectric, antiferroelectric, and paraelectric phases for diverse applications. Wu et al. reported Pb0.91La0.06(Zr0.58Ti0.42)0.975Nb0.02O3 prepared via the ceramic injection molding (CIM) technique with ultra-high d33 (731 pC/N) [12]. Kumar et al. reported a high effective piezoelectric constant d33* ~ 632 pm/V and a high unipolar strain rate S ~ 0.25% in Pb0.92La0.08(Zr0.60Ti0.40)O3, accompanied by a high Pr ~ 33 µC/cm2 and a low Ec ~ 10 kV [13]. As reported, Nb doping can compensate for the vacancy at the B site for ABO3-type perovskites caused by La doping. Nb-doping also has been proven to contribute to the improvement of piezoelectric responses and fatigue behavior, then the decrease in the coercive field [14,15]. In addition, Nb oxide is helpful to densify the Pb(Zr1−xTix)O3 (abbreviated as PZT)-based piezoceramics and decrease the sintering temperature slightly [14].
A multilayered design can be instrumental in the increase in micrometric displacement and the reduction in the driving field of PZT-based piezoelectric devices [16]. Multilayer ceramics are designed to co-sinter the inner electrodes and tape-cast ceramic sheet layers together. Developing the multilayered design is essential to lower the cost of products via co-sintering ceramic layers with cheap electrode paste, such as pure Ag or high Ag/low Pd paste instead of expensive platinum (Pt)/palladium (Pd) electrode paste. However, the low-temperature stability of cheap metal electrode paste limits the co-sintering of traditional PZT-based systems due to high sintering temperatures above 1200 °C [17,18]. There are various strategies to lower the sintering temperature of PZT-based piezoceramics, such as nanosized powder sintering [19], liquid phase assisted sintering [20], cold sintering process [21], high-energy ball milling [22], etc. Among these, the most straightforward and economical method is to form a liquid phase by sintering aid. Metal oxides/carbonates, such as B2O3, CuO, Bi2O3, ZnO, and Li2CO3, are the most used sintering additives. Moreover, the complex sintering aids can make low-temperature sintering effective [23]. Zhang et al. confirmed that the combined use of LiBiO2 and CuO has an excellent enhancement effect on d33* and optimal sinterability compared to pure LiBiO2 [24]. Choi et al. investigated that the mixed 4PbO-B2O3 liquid frits could optimize the sintering temperature of 0.4Pb(Zn1/3Nb2/3)O3–0.6Pb(Zr0.47Ti0.53)O3 piezoceramics, providing a liquid phase at below 800 °C [25].
The mechanism of sintering aids is that it can form a liquid transitional phase at a relatively low temperature, which can facilitate the dissolution and migration of substances, accelerate grain growth, and finally be absorbed into the ceramic to form a solid solution in the lattice and impurity at the grain boundary. Donnelly et al. reported that Li2CO3 reacted with excess PbO to form an intermediate liquid phase, the Li2PbO3 phase, which subsequently melted at 836 °C, resulting in the promotion of particle rearrangement [26]. Siddiqui et al. studied the effect of Li2CO3 on Pb0.93La0.02Sr0.05(Zr0.52Ti0.48)O3 (abbreviated as PLSZT) ceramics sintered at 850 °C. They found that the relative permittivity εr ~ 1270 and the planar electromechanical couple factor kp ~ 46% were enhanced compared to those without Li2CO3 sintered at 1150 °C (εr ~ 1145, kp ~ 33%) [27]. However, d33 decreased from 315 pC/N for pure PLSZT to 259 pC/N for the modified one. Kim et al. reported that CuO-doped 0.69Pb(Zr0.47Ti0.53)O3-0.31[Pb(Zn0.4Ni0.6)1/3Nb2/3]O3 ceramics exhibited a higher strain of 0.165% than that 0.145% for undoped ones [28]. Zhang et al. revealed that Li+ could segregate at the grain boundaries and suppress grain growth of 15Pb(Sc1/2Nb1/2)O3-52Pb(Mg1/3Nb2/3)O3-33PbTiO3 (abbreviated as PSN-PMN-PT) piezoceramics when excessive Li2CO3 was employed when sintered at 850 °C [29]. Both d33 and kp increased from 465 pC/N and 60% for pure PSN-PMN-PT ceramics to 520 pC/N and 62% for 0.5 wt% Li2CO3-modified one, respectively. Some other examples of low-temperature piezoceramics are listed in Table 1. It has been reported that Li2CO3 is a typical sintering aid usually used in PZT-based ceramics. The effect of the content of sintering aids on lead-based piezoelectric ceramics has been experimentally and theoretically studied. Commonly, the addition of sintering aids can always decrease the sintering temperature at the expense of reducing piezoelectricity. In this work, an appropriate amount of 0.2 wt% Li2CO3 was chosen as the sintering aid of commercial soft PZT-based piezoceramics to develop the low-temperature cofired piezoelectric ceramic system. The modified ceramics can be cofired with cheap metal electrodes while retaining excellent piezoelectric response.

2. Materials and Methods

Commercial soft PZT-based (PZT-5H) piezoelectric powders (Xi’an Konghong New Material Technology Co., Ltd., Xi’an, China) were selected as the base materials, and the Li2CO3 (99% purity, Shanghai Aladdin Reagent Co., Ltd., Shanghai, China) was selected as the sintering aid. The base materials were designed to form the 0.2 wt% Li2CO3-modified compositions for fabricating low-temperature cofired ceramic devices. The mixed powders were milled with ethanol for one day. The calcination process was conducted at 850 °C for 2 h. The resultant powders were granulated with a 2.0 wt.% polyvinyl alcohol (PVA) aqueous solution. Pressing was carried out to form the disc samples 12 mm in diameter under 180 MPa. The binders were fired at 600 °C for 2 h, and then the ceramics were sintered from 900 °C to 1100 °C for 2 h.
The ceramics were ground and polished up to a thickness of 0.5 mm on sandpaper. Density measurement was conducted using the Archimedes method. Silver pastes with conducting compositions of 80% Ag and 3% Pd were coated and fired for 10 min at 580 °C on both sides of the samples. The poling of the samples was conducted in silicon oil at 120 °C for 30 min under 25 kV/cm. The ceramics were aged for one day before piezoelectric measurements. The crystal structure was determined by X-ray diffraction (XRD) with CuKα = 1.54Å. The microstructures were observed through a scanning electron microscope (SEM, JSM-6700F). The piezoelectric d33 was measured by a ZJ-3A quasi-static piezoelectric meter 1 pC/N, and 0.2 N. Dielectric property-temperature spectra were characterized by an E4980A LCR meter connected with the temperature control system (Partulab, DMS-1000) in the temperature range from RT to 500 °C at a heating rate of 2 °C/min. The S-E relations can be evaluated by a ferroelectric workstation (aixACCT, TF analyzer 2000) under 20 kV/cm electric field strength. The polarization-filed (P-E) hysteresis loops were determined by the ferroelectric analyzer (PolyK, PK-CPE1701). The impedance |Z| and phase angle θ were collected by the impedance analyzer (Agilent, HP4294A, Santa Clara, CA, USA). The kp value was calculated according to Equation (1) [37]:
k p = 1 0.395   f r f a f r + 0.574
where fr and fa represent resonance and antiresonance frequencies, respectively, the mechanical quality factor Qm was calculated by Equation (2):
Q m = f a 2 2   π   f r   R   c 0 + c 1   f a 2 f r 2
where R, c1, and c0 are the minimum impedance at resonance, the capacitance of the ceramics at resonance, and the capacitance in an electrostatic field, respectively [37].

3. Results and Discussion

3.1. XRD Patterns

Figure 1 presents the phase structure of the sintered 0.2 wt% Li2CO3-modified bulk ceramics. All of the specimens show the pure perovskite single phase without any impurity composition under the limit of X-ray detection. The broadening of the (200) peak is observed in the samples sintered at 900~1000 °C at the 2θ = 44~46° region, which is evidence of the coexistence of the tetragonal and rhombohedral phases. In other words, near the morphotropic phase boundary (MPB). It is generally believed that piezoelectric ceramics have the optimum piezoelectric properties near the MPB. When the sintering temperature increases, a distinct splitting of the (002)T and (200)T peaks can be observed, which implies that the coexistence of the rhombohedral-tetragonal phase near MPB tends to transform into the tetragonal-rich phase. As reported, this transformation can be attributed to the stress relaxation caused by the grain growth of piezoelectric ceramics [38,39]. The ferroelectric ceramics will form an inner stress conducive to the stability of the rhombohedral phase, as reported. However, with the increase of the grain size, the internal stress is challenging to continue to concentrate and finally turns into cracks, releasing this stress and resulting in the rise of tetragonal phases [40].

3.2. SEM Images

Figure 2 shows the fractured cross-sectional SEM images and grain size distribution diagrams of the specimens at different sintering temperatures. Grain grows gradually, consistent with the sintering temperature. The average grain size of ceramic increases from 1.86 μm at 900 °C to 3.45 μm at 1100 °C. It can be observed that grain growth remains constant above 980 °C, with an average grain size of around 3.30 μm, similar to the previous report [31].
Figure 3 depicts dielectric property-temperature spectra of the 0.2 wt% Li2CO3-modified ceramics with various frequencies (100 Hz, 1 kHz, 10 kHz, and 100 kHz). With the increase in testing frequency, no shift or broadening of the dielectric peak was observed. Dielectric maximum peaks show sharp phase transitions indicative of a typical ferroelectric characteristic without any relaxor behavior. It can be concluded that the Curie temperature of the 0.2 wt% Li2CO3-modified ceramics is around 226 °C. It can be deduced that piezoelectric devices based on these compositions can be used in temperatures above 110 °C. For that, the poled piezoelectric ceramics often begin to become unstable at a temperature somewhat above half the Curie temperature.
The physical parameters of the studied ceramic samples, including kp, Qm, d33, d33*, Ec, and Pr, are summarized in Table 2. The piezoceramics sintered at 980 °C have the optimal piezoelectric characteristics with a d33 of 701 pC/N and a kp of 66.7%. Qm values of all the samples remain constant at about 70, independent of sintering temperatures. Lead vacancies which are widely thought will result in low Qm, making the ceramics “soft”. Lead vacancies are propitious to domain switching. Thus, soft piezoelectric ceramics feature quick response speed, high piezoelectric constant, and low mechanical quality factor value. As reported, it is hard to ensure that the ceramics sintered at low temperatures have a high Qm [41]. At low sintering temperatures, the effect of temperature on lead/oxygen vacancies is far less than that of doping [42]. This should be responsible for the temperature independence for Qm. The impedance plots are shown in Figure 4.
Figure 5 illustrates the ferroelectric properties of the 0.2 wt% Li2CO3-modified ceramics. To achieve polarization saturation, the specimens were tested under 40 kV/cm, which is about four times Ec. The maximum remnant polarization Pr (32.60 μC/cm2) can be obtained for the ceramics sintered at 980 °C. The remnant polarization has almost the same trend as the d33 due to the relationship of εrPr.
As mentioned above, the piezoelectricity of the 0.2 wt% Li2CO3-modified ceramics exhibits slight variation with the increase in sintering temperature. Ultrahigh piezoelectric properties can be obtained under a low sintering temperature below 1000 °C, facilitating the use of cheap metal in the cofired ceramic devices.
Figure 6a depicts the unipolar strain curves (S-E) of the 0.2 wt% Li2CO3-modified ceramics under 20 kV/cm. The strain of the ceramics remains almost constant on the variation of sintering temperature in the range of 0.12–0.14%. The normalized strain large-signal longitudinal piezoelectric constant d33* can be calculated by Equation (3) and listed in Table 2.
d 33 * = S max E max
Regardless of the difference in grain size, piezoelectric strain remains constant, indicative of the maximization of reversible 90° domain switching.
However, piezoelectric ceramics are very brittle and susceptible to fracture during the electric drive. The electric field applied to the ceramic commonly causes the volume expansion/shrinkage of the unit cell of the ceramics. The variation of lattice volume usually causes internal stress, easily leading to the crack of piezoceramics. Cyclic testing of unpoled piezoelectric ceramics effectively verifies the fatigue behavior of piezoelectric devices [43]. Here, a working electric field of 25 kV/cm was selected to determine the cyclic fatigue properties by P-E loops of piezoelectric ceramics in Figure 6b. The ceramic samples were tested by 104 times at 1 Hz with a triangle input signal. It can be observed that piezoelectric ceramics sintered at low temperatures become easy to dielectric breakdown, while the piezoelectric ceramics sintered above 960 °C can withstand 104 cycles of electric excitation. It suggests that a high sintering temperature for ceramics contributes to the enhancement of fracture strength, possibly due to the improved bulk density or less porosity. It can be confirmed that the optimal Ts is near 960 °C, matching the sinterability of silver paste.

4. Conclusions

In this work, the 0.2 wt% Li2CO3-modified piezoelectric ceramics were developed to reduce sintering temperatures for low-temperature cofired applications. The influence of Ts on microstructure, dielectric, and piezoelectric properties was studied. Due to sintering aids, the piezoelectric ceramics could be densified at a low sintering temperature above 940 °C, suitable for co-firing with cheap electrodes. The piezoelectric ceramics can be optimized at a sintering temperature of 980 °C, accompanied by high piezoelectric d33 ~ 701 pC/N, kp ~ 66.7%, and a low Qm ~ 71 with a transition temperature of Tc ~ 226 °C, presenting the characteristics of typical soft piezoelectric ceramics. It had a promising strain of 0.13% under a low electric field (20 kV/cm) and was competitive for the fabrication of multilayer piezoelectric transducers.

Author Contributions

Data curation, Z.L. and Z.Z.; Project administration, Z.Y. and H.H.; Supervision, Z.Y. and H.L.; Writing—original draft, Z.L., Z.Z. and H.Z.; Writing—review & editing, Z.Y., H.Z., H.H., M.C. and H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Natural Science Foundation of China (51872213), Sanya Science and Education Innovation Park of Wuhan University of Technology (2021KF0014), Guangdong Basic and Applied Basic Research Foundation (2021A1515110060), and Guangdong Basic and Applied Basic Research Foundation (2022A1515010073).

Data Availability Statement

The datasets generated during and analyzed during the current study are available from the corresponding author on reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, S.J. High entropy design: A new pathway to promote the piezoelectricity and dielectric energy storage in perovskite oxides. Microstructures 2023, 3, 2023003. [Google Scholar]
  2. 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]
  3. Gao, X.Y.; Yang, J.K.; Wu, J.G.; Xin, X.D.; Li, Z.M.; Yuan, X.T.; Shen, X.Y.; Dong, S.X. Piezoelectric actuators and motors: Materials, designs, and applications. Adv. Mater. Technol. 2020, 5, 1900716. [Google Scholar] [CrossRef]
  4. Gurdal, A.E.; Tuncdemir, S.; Uchino, K.; Randall, C.A. Low temperature co-fired multilayer piezoelectric transformers for high power applications. Mater. Des. 2017, 132, 512–517. [Google Scholar] [CrossRef]
  5. Wang, Q.Z.; Li, F.Z. A low-working-field (2 kV/mm), large-strain (>0.5%) piezoelectric multilayer actuator based on periodically orthogonal poled PZT ceramics. Sens. Actuator A-Phys. 2018, 272, 212–216. [Google Scholar] [CrossRef]
  6. Li, Z.M.; Bian, L.; Yi, X.Y.; Zhu, R.Q.; Yuan, X.T.; Dong, S.X. A co-fired multilayer PMnS-PZT ceramic based linear piezoelectric ultrasonic micromotor with a high-load-density. Appl. Phys. Lett. 2022, 120, 082902. [Google Scholar] [CrossRef]
  7. Shen, Z.Y.; Qin, C.; Luo, W.Q.; Song, F.S.; Wang, Z.M.; Li, Y.M.; Zhang, S.J. Ce and W co-doped CaBi2Nb2O9 with enhanced piezoelectric constant and electrical resistivity at high temperature. J. Mater. 2020, 6, 459–466. [Google Scholar] [CrossRef]
  8. Kour, P.; Pradhan, S.K.; Kumar, P.; Sinha, S.; Kar, M. Effect of Nd doping on dielectric and impedance properties of PZT nanoceramics. J. Electron. Mater. 2018, 47, 2861–2870. [Google Scholar] [CrossRef]
  9. Samanta, S.; Sankaranarayanan, V.; Sethupathi, K.; Ramachandra Rao, M.S. Enhanced ferroelectricity in PLZT ceramic by precise La-doping, minimizing pyrochlore phase and lead loss. Vacuum 2018, 157, 514–523. [Google Scholar] [CrossRef]
  10. Promjun, T.; Funsueb, N.; Ngamjarurojana, A. Effect of Nb, Ta and Sb addition on structure and electrical properties of PZT ceramics. Mater. Today Proc. 2019, 17, 1602–1606. [Google Scholar] [CrossRef]
  11. Samanta, S.; Sankaranarayanan, V.; Sethupathi, K. Effect of successive multiple doping of La, Nb and Fe on structure and lattice vibration of MPB PZT. Mater. Today Proc. 2018, 5, 27919–27927. [Google Scholar] [CrossRef]
  12. Wu, D.; Qin, S.; Liu, H.; Cao, Z.; Chen, J.F.; Fang, B.J.; Liu, C.L. Enhancing piezoelectricity of PLZT ceramics by bismuth stearate coating via water-soluble defatted powder injection molding. Ceram. Int. 2021, 47, 17956–17963. [Google Scholar] [CrossRef]
  13. Kumar, A.; Raju, K.C.; Ryu, J.; James, A.R. Composition dependent Ferro-piezo hysteresis loops and energy density properties of mechanically activated (Pb1−xLax)(Zr0.60Ti0.40)O3 ceramics. Appl. Phys. A-Mater. Sci. Process. 2020, 126, 175. [Google Scholar] [CrossRef]
  14. Chu, S.Y.; Chen, T.Y.; Tsai, I.T. Effects of sintering temperature on the dielectric and piezoelectric properties of Nb-doped PZT ceramics and their applications. Integr. Ferroelectr. 2003, 58, 1293–1303. [Google Scholar] [CrossRef]
  15. Lee, W.I.; Lee, J.K. Dopant effects on the grain structure and electrical property of PZT thin films prepared by sol-gel process. Mater. Res. Bull. 1995, 30, 1185–1191. [Google Scholar] [CrossRef]
  16. Yan, S.; Sun, C.; Cui, Q.Y.; He, M.Z.; Willhandy; Wang, R.B.; Hao, J.J.; Chu, X.C. Dielectric, piezoelectric and dc bias characteristics of Bi-doped PZT multilayer ceramic actuator. Mater. Chem. Phys. 2020, 255, 123605. [Google Scholar] [CrossRef]
  17. Zeng, Z.Q.; Wu, Q.C.; Hao, M.M.; Lu, W.Z.; Fan, G.F.; Ming, Y.C.; Ding, M.Y. Impedance spectroscopy and piezoelectric property of LiF-doped PZN-PZT low-temperature sintering piezoelectric ceramics. J. Mater. Sci.-Mater. Electron. 2018, 29, 8279–8286. [Google Scholar] [CrossRef]
  18. Zhang, M.H.; Wang, K.; Du, Y.J.; Dai, G.; Sun, W.; Li, G.; Hu, D.; Thong, H.C.; Zhao, C.L.; Xi, X.Q. High and temperature-insensitive piezoelectric strain in Alkali niobate lead-free perovskite. J. Am. Chem. Soc. 2017, 139, 3889–3895. [Google Scholar] [CrossRef]
  19. Ghasemifard, M.; Hosseini, S.M.; Zak, A.K.; Khorrami, G.H. Microstructural and optical characterization of PZT nano-powder prepared at low temperature. Phys. E Low-Dimens. Syst. Nanostruct. 2009, 41, 418–422. [Google Scholar] [CrossRef]
  20. Corker, D.L.; Whatmore, R.W.; Ringgaard, E.; Wolny, W.W. Liquid-phase sintering of PZT ceramics. J. Eur. Ceram. Soc. 2000, 20, 2039–2045. [Google Scholar] [CrossRef]
  21. Guo, J.; Guo, H.; Baker, A.L.; Lanagan, M.T.; Kupp, E.R.; Messing, G.L.; Randall, C.A. Cold sintering: A paradigm shift for processing and integration of ceramics. Angew. Chem. Int. Ed. 2016, 55, 11457–11461. [Google Scholar] [CrossRef] [PubMed]
  22. Siddiqui, M.; Mohamed, J.J.; Ahmad, Z.A. Structural properties of Pb(Zr0.52Ti0.38Li0.1)O3 prepared via high planetary mill. Mater. Sci. Forum. 2017, 888, 91–95. [Google Scholar]
  23. Jan, H.; Francesco, P.; Sona, H.; Vojtech, N. Sintering aids, their role, and behavior in the production of transparent ceramics. Open Ceram. 2021, 7, 100137. [Google Scholar]
  24. Zhang, W.L.; Eitel, R.E. Low-temperature sintering and properties of 0.98PZT-0.02SKN ceramics with LiBiO2 and CuO addition. J. Am. Ceram. Soc. 2011, 94, 3386–3390. [Google Scholar] [CrossRef]
  25. Choi, J.J.; Hahn, B.D.; Park, D.S.; Yoon, W.H.; Lee, J.H.; Jang, J.H.; Ko, K.H.; Park, C. Low-temperature sintering of lead zinc niobate-lead zirconate titanate ceramics using nano-sized powder prepared by the stirred media milling process. J. Am. Ceram. Soc. 2007, 90, 388–392. [Google Scholar] [CrossRef]
  26. Donnelly, N.J.; Shrout, T.R.; Randall, C.A. The role of Li2CO3 and PbO in the low-temperature sintering of Sr, K, Nb (SKN)-doped PZT. J. Am. Ceram. Soc. 2009, 92, 1203–1207. [Google Scholar] [CrossRef]
  27. Siddiqui, M.; Mohamed, J.J.; Ahmad, Z.A. Piezoelectric and dielectric properties of Pb0.93La0.02Sr0.05(Zr0.52Ti0.48)O3 ceramics doped with Li2CO3 at low sintering temperature. Ceram. Int. 2016, 43, 2644–2649. [Google Scholar] [CrossRef]
  28. Kim, B.S.; Ji, J.H.; Koh, J.H. Improved strain and transduction values of low-temperature sintered CuO-doped PZT-PZNN soft piezoelectric materials for energy harvester applications. Ceram. Int. 2021, 47, 6683–6690. [Google Scholar] [CrossRef]
  29. Zhang, S.K.; Guo, F.F.; Jia, B.; Zhou, H.Q.; Dai, Z.H.; Xi, Z.Z. Effect of Li2CO3 on the structure and electrical properties of low-temperature-sintered 15PSN-52PMN-33PT ceramics. J. Am. Ceram. Soc. 2022, 33, 2044–2051. [Google Scholar] [CrossRef]
  30. Quang, D.A.; Vuong, L.D. Enhanced piezoelectric properties of Fe2O3 and Li2CO3 co-doped Pb[(Zr0.48Ti0.52)0.8(Zn1/3Nb2/3)0.125(Mn1/3Nb2/3)0.075]O3 ceramics for ultrasound transducer application. J. Mater. Sci. Adv. Mater. Dev. 2022, 7, 100436. [Google Scholar]
  31. Gao, X.Y.; Jin, H.N.; Xin, B.J.; Wang, M.W.; Dong, S.X.; Xu, Z.; Li, F. Low-temperature sintering of Li2CO3 added Pb(Ni1/3Nb2/3)-Pb(Zr,Ti)O3 ceramics with high piezoelectric properties. J. Alloys Compd. 2022, 892, 162132. [Google Scholar] [CrossRef]
  32. Fan, G.F.; Shi, M.B.; Lu, W.Z.; Wang, Y.Q.; Liang, F. Effects of Li2CO3 and Sm2O3 additives on low-temperature sintering and piezoelectric properties of PZN-PZT ceramics. J. Eur. Ceram. Soc. 2014, 34, 23–28. [Google Scholar] [CrossRef]
  33. Zhang, Q.; Yue, Y.; Nie, R.; Liu, H.; Chen, Q.; Yu, P.; Zhu, J.G.; Xiao, D.Q. Achieving both high d33 and high TC in low-temperature sintering Pb(Ni1/3Nb2/3)O3-Pb(Mg1/2W1/2)O3-Pb(Zr0.5Ti0.5)O3 ceramics using Li2CO3. Mater. Res. Bull. 2017, 85, 96–103. [Google Scholar] [CrossRef]
  34. Chen, H.; Xing, J.; Xi, J.W.; Pu, T.; Liu, H.; Zhu, J.G. Origin of high piezoelectricity in low-temperature sintering PZT-based relaxor ferroelectric ceramics. J. Alloys Compd. 2020, 860, 157930. [Google Scholar] [CrossRef]
  35. Chen, H.; Pu, T.; Fan, S.B.; Liu, H.; Zhu, J.G.; Chen, Q. Enhanced electrical properties in low-temperature sintering PNN-PMW-PZT ceramics by Yb2O3 doping. Mater. Res. Bull. 2022, 146, 111576. [Google Scholar] [CrossRef]
  36. Nie, R.; Zhang, Q.; Yue, Y.; Liu, H.; Chen, Q.; Zhu, J.G.; Yu, P.; Xiao, D.Q. Properties of Low-Temperature Sintering PNN-PMW-PSN-PZT Piezoelectric Ceramics with Ba(Cu1/2W1/2)O3 Sintering Aids. Int. J. Appl. Ceram. Technol. 2016, 13, 1119–1124. [Google Scholar] [CrossRef]
  37. Yang, S.L.; Tsai, C.C.; Hong, C.S.; Chu, S.Y. Effects of sintering aid CuTa2O6 on piezoelectric and dielectric properties of sodium potassium niobate ceramics. Mater. Res. Bull. 2012, 47, 998–1003. [Google Scholar] [CrossRef]
  38. Chen, Y.; Xie, S.X.; Wang, Q.Y.; Fu, L.; Nie, R.; Zhu, J.G. Correlation between microstructural evolutions and electrical/mechanical behaviors in Nb/Ce co-doped Pb(Zr0.52Ti0.48)O3 ceramics at different sintering temperatures. Mater. Res. Bull. 2017, 94, 174–182. [Google Scholar] [CrossRef]
  39. Dong, Y.Z.; Zhou, Z.Y.; Liang, R.H.; Dong, X.L. Correlation between the grain size and phase structure, electrical properties in BiScO3-PbTiO3-based piezoelectric ceramics. J. Am. Ceram. Soc. 2020, 103, 4785–4793. [Google Scholar] [CrossRef]
  40. Zheng, M.P.; Hou, Y.D.; Ge, H.Y.; Zhu, M.K.; Yan, H. Effect of NiO additive on microstructure, mechanical behavior and electrical properties of 0.2PZN-0.8PZT ceramics. J. Eur. Ceram. Soc. 2013, 33, 1447–1456. [Google Scholar] [CrossRef]
  41. Du, Z.Z.; Zhao, C.L.; Thong, H.C.; Zhou, Z.; Zhou, J.J.; Wang, K.; Guan, C.L.; Liu, H.; Fang, J.Z. Effect of MnCO3 on the electrical properties of PZT-based piezoceramics sintered at low temperature. J. Alloys Compd. 2019, 801, 27–32. [Google Scholar] [CrossRef]
  42. Geng, W.P.; Chen, X.; Pan, L.; Qiao, X.J.; He, J.; Mu, J.L.; Hou, X.J.; Chou, X.J. Improved crystallization, domain, and ferroelectricity by controlling lead/oxygen vacancies in Mn-doped PZT thin films. Mater. Charact. 2021, 176, 111131. [Google Scholar] [CrossRef]
  43. Beom, H.G.; Jeong, K.M.; Park, J.Y.; Lin, S.; Kim, G.H. Electrical failure of piezoelectric ceramics with a conductive crack under electric fields. Eng. Fract. Mech. 2009, 76, 2399–2407. [Google Scholar] [CrossRef]
Figure 1. XRD patterns for the 0.2 wt% Li2CO3-modified bulk ceramics.
Figure 1. XRD patterns for the 0.2 wt% Li2CO3-modified bulk ceramics.
Materials 16 01679 g001
Figure 2. SEM images and grain size distribution for the 0.2 wt% Li2CO3-modified ceramics sintered at various temperatures: (a) 900 °C; (b) 920 °C; (c) 940 °C; (d) 960 °C; (e) 980 °C; (f) 1000 °C; (g) 1050 °C; (h) 1100 °C.
Figure 2. SEM images and grain size distribution for the 0.2 wt% Li2CO3-modified ceramics sintered at various temperatures: (a) 900 °C; (b) 920 °C; (c) 940 °C; (d) 960 °C; (e) 980 °C; (f) 1000 °C; (g) 1050 °C; (h) 1100 °C.
Materials 16 01679 g002
Figure 3. Dielectric property-temperature spectra of the 0.2 wt% Li2CO3-modified ceramics sintered at different temperatures: (a) 900 °C; (b) 920 °C; (c) 940 °C; (d) 960 °C; (e) 980 °C; (f) 1000 °C; (g) 1050 °C; (h) 1100 °C.
Figure 3. Dielectric property-temperature spectra of the 0.2 wt% Li2CO3-modified ceramics sintered at different temperatures: (a) 900 °C; (b) 920 °C; (c) 940 °C; (d) 960 °C; (e) 980 °C; (f) 1000 °C; (g) 1050 °C; (h) 1100 °C.
Materials 16 01679 g003
Figure 4. Impedance plots of the 0.2 wt% Li2CO3-modified ceramics: (a) 900 °C sintered; (b) 920 °C sintered; (c) 940 °C sintered; (d) 960 °C sintered; (e) 980 °C sintered; (f) 1000 °C sintered; (g) 1050 °C sintered; (h) 1100 °C sintered.
Figure 4. Impedance plots of the 0.2 wt% Li2CO3-modified ceramics: (a) 900 °C sintered; (b) 920 °C sintered; (c) 940 °C sintered; (d) 960 °C sintered; (e) 980 °C sintered; (f) 1000 °C sintered; (g) 1050 °C sintered; (h) 1100 °C sintered.
Materials 16 01679 g004
Figure 5. Ferroelectric hysteresis loop under 40 kV/cm.
Figure 5. Ferroelectric hysteresis loop under 40 kV/cm.
Materials 16 01679 g005
Figure 6. (a) Unipolar strain (S-E) curves of the 0.2 wt% Li2CO3-modified ceramics under 20 kV/cm; (b) The maximum cycle times; (c) Density of the 0.2 wt% Li2CO3-modified ceramics; (d) Cyclically P-E curve of piezoelectric ceramics under 25 kV/cm.
Figure 6. (a) Unipolar strain (S-E) curves of the 0.2 wt% Li2CO3-modified ceramics under 20 kV/cm; (b) The maximum cycle times; (c) Density of the 0.2 wt% Li2CO3-modified ceramics; (d) Cyclically P-E curve of piezoelectric ceramics under 25 kV/cm.
Materials 16 01679 g006
Table 1. Comparison of piezoelectric properties of PZT-based ceramics with sintering aids (Ts: sintering temperature; Tc: Curie temperature).
Table 1. Comparison of piezoelectric properties of PZT-based ceramics with sintering aids (Ts: sintering temperature; Tc: Curie temperature).
Sintering AidCeramic CompositionTs
(°C)
Tc
(°C)
d33
(pC/N)
Li2CO3 (0.07 wt.%) + Fe2O3 (0.25 wt.%) [30]PZT-PZN-PMnN1000 455
Li2CO3 (0.2 wt%) [31]PNN-PZT950113692
Li2CO3 (0.3 wt%) + Sm2O3 (0.3 wt%) [32]0.3PZN-0.7PZT900394483
Li2CO3 (0.2 wt%) + CaCO3 (0.3 wt%) [33]PNN-PMW-PZT900296599
Li2CO3 (0.2 wt%) + CaCO3 (0.25 wt%) + Ta2O5 (0.7 wt%) [34]PNN-PMW-PZT940640303
Yb2O3 (0.1 mol%) [35]PNN-PMW-PZT900179623
Sm2O3 (0.3 wt%) + LiF (0.01%) [17]PZN-PZT950289403
Ba(Cu1/2W1/2)O3 (0.1 wt%) [36]PNN–PMW–PSN–PZT900161551
Table 2. The specific values of electrical parameters of the 0.2 wt% Li2CO3-modified PZT-based ceramics.
Table 2. The specific values of electrical parameters of the 0.2 wt% Li2CO3-modified PZT-based ceramics.
Tsintering
(°C)
Tc
(°C)
d33
pC/N
d33*
pm/V
Ec
kV/cm
Pr
μC/cm2
kp
(%)
Qm
90022661870111.9829.1663.475
92022866564014.5430.3764.676
94022668265210.8627.6665.473
96022468362411.5327.9966.269
98022970165110.1832.6066.771
100022966164811.2827.6866.370
105023159771612.2630.2960.282
110022659172110.0330.1258.178
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lin, Z.; Zhu, Z.; Yao, Z.; Zhang, H.; Hao, H.; Cao, M.; Liu, H. Piezoelectric Response and Cycling Fatigue Resistance of Low-Temperature Sintered PZT-Based Ceramics. Materials 2023, 16, 1679. https://doi.org/10.3390/ma16041679

AMA Style

Lin Z, Zhu Z, Yao Z, Zhang H, Hao H, Cao M, Liu H. Piezoelectric Response and Cycling Fatigue Resistance of Low-Temperature Sintered PZT-Based Ceramics. Materials. 2023; 16(4):1679. https://doi.org/10.3390/ma16041679

Chicago/Turabian Style

Lin, Zirui, Zhe Zhu, Zhonghua Yao, Hao Zhang, Hua Hao, Minghe Cao, and Hanxing Liu. 2023. "Piezoelectric Response and Cycling Fatigue Resistance of Low-Temperature Sintered PZT-Based Ceramics" Materials 16, no. 4: 1679. https://doi.org/10.3390/ma16041679

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