BiScO3-PbTiO3 Piezoelectric Ceramics With Bi Excess For Energy Harvesting

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 43

Journal Pre-proof

BiScO3-PbTiO3 piezoelectric ceramics with Bi excess for energy harvesting


applications under high temperature

Jae-Hoon Ji, Dong-Jin Shin, Jinhwan Kim, Jung-Hyuk Koh

PII: S0272-8842(19)32980-3
DOI: https://doi.org/10.1016/j.ceramint.2019.10.117
Reference: CERI 23185

To appear in: Ceramics International

Received Date: 3 September 2019


Revised Date: 10 October 2019
Accepted Date: 12 October 2019

Please cite this article as: J.-H. Ji, D.-J. Shin, J. Kim, J.-H. Koh, BiScO3-PbTiO3 piezoelectric ceramics
with Bi excess for energy harvesting applications under high temperature, Ceramics International (2019),
doi: https://doi.org/10.1016/j.ceramint.2019.10.117.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Ltd.


BiScO3-PbTiO3 piezoelectric ceramics with Bi excess for energy

harvesting applications under high temperature

Jae-Hoon Jia, Dong-Jin Shina,b, Jinhwan Kima and Jung-Hyuk Koh a,*
a
School of Electrical and Electronic Engineering, Chung-Ang University, Seoul,

Republic of Korea
b
Korea Electrotechnology Research Institute, Changwon, Republic of Korea

Abstract

0.36BiScO3-0.64PbTiO3 ceramic is a competitive piezoelectric material even

though it contains lead and volatile Bi contents. It contains a relatively decreased lead

content compared to that of the Pb(Zr,Ti)O3 system but it has similar piezoelectric

properties with high Curie temperature. However, due to the very volatile component of

Bi the 0.36BiScO3-0.64PbTiO3 system has Bi-deficient composition. Therefore, in

order to compensate for deficient Bi contents in the 0.36BiScO3-0.64Pb,TiO3 system,

excess 0.005, 0.01, 0.015 and 0.02 mol of Bi were added to the ceramics to enhance the

piezoelectric properties for the first time. By employing excess Bi addition, the

piezoelectric charge coefficient, electromechanical coupling factor, output open circuit

voltage, and generated output power density were improved from 417 pC/N, 51.48 %,

19.69 V and 0.28 mJ/cm3 to 452 pC/N, 52.25 %, 26.93 V and 0.48 mJ/cm3. We expect

that piezoelectric properties of 0.36BiScO3-0.64PbTiO3 ceramics were improved by

adding Bi excess.

1
Keywords: BS-PT ceramics with Bi excess; Piezoelectric properties; Energy

harvesting
*
Corresponding author. Tel: +82-2-820-5311; Fax: +82-2-825-1584

E-mail address: jhkoh@cau.ac.kr

Postal address: School of Electrical and Electronic Engineering, Chung-Ang

University, 84 Heukseok-Ro, Dong-Jak Gu, Seoul, Korea.

2
1. Introduction

Pb(Zr,Ti)O3 (PZT)-based ceramics have been widely applied in electronic

devices such as sensors, resonators, actuators and transducers, since PZT-based

ceramics exhibit excellent piezoelectric properties [1-3]. However, the easy de-poling

and poor temperature stability of PZT-based ceramics at temperatures above 200 °C can

be one of the main obstacles for certain device applications [4, 5]. Therefore,

piezoelectric ceramics with both high piezoelectric properties and Curie temperature

have been intensively researched.

Recently, the BiScO3-PbTiO3 (BS-PT) ceramics close to the MPB region have

shown excellent piezoelectric properties, such as a piezoelectric charge coefficient (d33)

of 400 pC/N and a high Curie temperature (Tc) of 450 °C [6, 7]. This indicates that BS-

PT ceramics have similar piezoelectric properties and even better high-temperature

performance than those of PZT ceramics [8]. Moreover, these BS-PT ceramics contain

lower lead content in terms of composition. For certain piezoelectric devices

applications, there are two strict regulations to follow from the industrial aspects. The

first thing is the high piezoelectric charge coefficient more than 200 pC/N and the

second thing is the high Curie temperature more than 300 °C. Moreover, lower lead

content piezoelectric ceramic materials may be considered more favorable materials for

device applications.

Meanwhile, the Bi2O3 element in the BS-PT ceramics is likely to have problems

of volatility, as indicated by the low melting point of Bi2O3 and PbO. The melting

temperature of Bi2O3 is around 817 °C and the melting point of PbO is 889 °C which is

slightly higher than that of Bi2O3 [9-10]. Considering the vapor pressures of

piezoelectric ceramics, the vapor pressure of Bi2O3 over Bi0.5K0.5TiO3 is very close to

3
the vapor pressure of PbO over PZT ceramics [11-12]. By considering the melting

temperature of Bi2O3 (817 °C) and PbO (889 °C), the volatilization degree will similar.

On the while, Pb can be considered as toxic component, PbO excess was not considered.

The easy volatilization of Bi2O3 and PbO can cause a composition change in the BS-PT

ceramics, which often results in defects, porosity, unusual grain growth and

heterogeneous microstructures during the sintering process.

Unfortunately, the Bi deficiency problem of BS-PT ceramics has not yet been

reported. However, there have been many reports on the (Na,K)NbO3 (NKN)

piezoelectric ceramics for their volatile Na and K components [13-16]. In order to solve

this deficiency problems, excess addition of volatile composition was introduced during

the sintering process. In the case of NKN, excess Na and K were introduced [17-19]. It

seems that pre-addition of the volatile components can enhance the dielectric and

piezoelectric properties with increasing the densification [13]. By adding excess Bi2O3,

it is to be expected that the amount of Bi loss during the sintering process can be

compensated, and the stoichiometry of the BS-PT ceramics can be optimized. Also, it is

necessary to optimize the excess Bi content and sintering temperature for provide

superior dielectric and piezoelectric properties of BS-PT ceramics with Bi excess. The

optimized composition and process condition of piezoelectric materials are important

for the development of enhanced piezoelectric energy harvesters. However, there are no

reports on improving the dielectric and piezoelectric properties by optimizing the excess

Bi content and sintering temperature of the BS-PT ceramics.

In this research, the effect of excess Bi content and sintering temperature in the

BS-PT ceramics will be discussed. By adding excess Bi to BS-PT ceramics, we

expected that the BS-PT ceramics show improved the dielectric and piezoelectric

4
properties for energy harvester applications. Also, the microstructure, dielectric and

piezoelectric properties of the BS-PT ceramics were systematically investigated.

2. Experimental Procedure

The 0.36BS–0.64PT ceramics were prepared by employing a conventional

sintering process. Bi2O3 (99.9 %), Sc2O3 (99.9 %), PbO (99.0 %) and TiO2 (99.9 %)

were used as the starting materials. To prepare 0.36BS–0.64PT ceramics with Bi excess,

0 – 0.02 mol of Bi1O1.5 (To avoid confusing mol calculation, Bi1O1.5 was employed,

instead of Bi2O3) were added in the starting powder. To add Bi, Bi2O3 powder was

employed instead of Bi. The powders were dried in an oven at 120 °C for 12 h. The

powders were weighed according to the chemical formula and milled for 24 h with

zirconia balls, dried, and calcined at 780 °C for 2 h. To make disk shape specimen, 3

metric ton was applied to the 12 mm diameter, it means pressure of 2,654 kgf/cm2 was

applied to the specimen, this value converted to the 265 MPa. This disk was sintered at

temperatures ranging from 1125 to 1225 °C for 2 h. The obtained samples were cut and

polished, and silver paste was fired at 700 °C for 10 min. The samples were poled in 80

°C silicon oil by applying a DC electric field of 4.0 kV/mm for 30 min. In order to

analyze the crystalline structures of the samples, X-ray diffraction (XRD, Bruker-AXS;

New D8-Advance) was performed. Field emission scanning electron microscopy (FE-

SEM, Carl Zeiss; SIGMA) was used to study the morphologies and microstructures of

the samples. The dielectric constant (ɛr) of the samples were measured using an

impedance analyzer from 1kHz to 1 MHz (Agilent 4294A precision). The piezoelectric

charge coefficient, d33, was measured using a Berlincourt quasi-static d33 meter. The

electromechanical coupling factor, kp, was calculated using the resonance-antiresonance

5
frequency. The generated open circuit voltage and power density were measured using

an oscilloscope (Agilent Technologies, DSO-X2002A) and a Femto/Picoammeter

(Agilent Technologies, B2981A), respectively.

3. Results and Discussion

Figure 1 shows the X-ray diffraction patterns and peak ratio of (00l) of BS-PT

ceramics with 0 – 0.02 mol Bi excess sintered at 1200 °C, (a), (c): without poling and

(b), (d): with poling. As shown in the figure 1 (a), all specimens exhibit a single

perovskite structure without any secondary phase. Figure 1 (b) indicate the XRD

patterns of BS-PT ceramics with Bi excess after poling process. The samples of BS-PT

ceramics were poled in 80 °C silicon oil under an electric field of 4.0 kV/mm for 30

min. Also, the Ag electrode was removed before analysis. As shown in the figure, all

peak intensities of BS-PT ceramics with 0 – 0.02 mol Bi excess were increased after

poling process. BS-PT ceramics with Bi excess have tetragonal structure with (00l)

orientation. Meanwhile, growing of (00l) orientation is closely related to piezoelectric

properties [20]. The peak ratio at (00l) of the BS-PT ceramics with Bi excess was

obtained as shown in the figure 1 (c) and (d). The peak ratio of (00l) is calculated by the

following equation.


Peak ratio 00 = ∑

(1)

As shown in figure 1 (c), the peak ratios of (00l) of unpoled BS-PT ceramics with 0 –

0.02 mol Bi excess at 2 theta range of 20° to 70° were 16.8, 16.9, 17.4, 16.8 and 16.4 %,

respectively. The peak ratio (00l) was increased with excess Bi content up to 0.01 mol,

and then decreased. In the figure 1 (d), BS-PT ceramics with 0.01 mol Bi excess have

6
the highest peak ratio of 21.9 %. By comparison with unpoled specimens of figure 1 (c),

the peak ratio of poled BS-PT ceramics with Bi excess was higher than those of unpoled

BS-PT with Bi excess. The peak ratio of (00l) orientation means relative peak

intensities of (00l) compared with those of other orientations. This high (00l) peak ratio

represents high c-axis orientation, which is related to higher polarization or

piezoelectric properties. We believe tetragonal phase was increased by adding Bi up to

0.01 mol, this can be seen in the figure 1 (b) and (d) until we doped Bi up to 0.01 mol,

the peak ratio of (00l) was increased up to 17.4 %, then peak ratio was decreased. By

considering figure 5, the density of BS-PT together, tetragonal phase was increased

during the densification process. Since BS-PT ceramics have mixture phase of

tetragonal and monoclinic structure, increased tetragonal structure can be expressed

increased ratio of (00l) peak ratio. In all compositions, both (001) and (100) peaks were

observed, also (102) and (210) peaks were observed. This is the evidence of mixture

phase of tetragonal and monoclinic structure. Therefore, we believe that BS-PT

ceramics with Bi excess have growing of (00l) orientation up to 0.01 mol. This indicates

that BS-PT ceramics with Bi excess has improved density and tetragonal phase to

enhance piezoelectric and dielectric properties.

Figure 2 shows the Rietveld refinement results of stoichiometric BS-PT ceramic

and BS-PT ceramic with 0.01 mol Bi excess using the High Score Plus program.

According to figure 1, BS-PT ceramics have a mixture phase of tetragonal and

monoclinic structure. Therefore, a two-phase refinement method (P4mm + Cm) was

used. Non-centrosymmetric space group P4mm and the polar type space group Cm

belongs to tetragonal structure and monoclinic structure, respectively. As a result, both

of observed data and Rietveld refinement patterns were well coincided with each other.

7
The simulated parameters through Rietveld refinements for BS-PT with 0 – 0.02 mol Bi

excess ceramics are listed in Table 1. The simulated parameters were profile residual

factor Rp, weighted profile residual factor Rwp, and goodness of fit χ2. As shown in the

table 1, goodness of fit is around 1.2. This low goodness of fit represents the simulation

values are reliable properties. In the Table 1, as the Bi excess concentration increase

from 0 to 0.02 mol, the phase fraction of the tetragonal (P4mm) in BS-PT ceramics

increased from 28.5 to 42.6 %. By varying Bi contents, the phase ratio between the

tetragonal and monoclinic has been changed. This refinement results coincide well with

the peak ratio in figure 1 (c) and (d). Therefore, we believe that the addition of 0.01 mol

Bi excess to BS-PT ceramics can improve polarization or piezoelectric properties as

increasing c-axis orientation.

Figure 3 shows (a) the X-ray diffraction patterns of BS-PT ceramics with 0.01

mol Bi excess sintered from 1125 to 1225 °C, (b) magnified peak position of (110) peak

and (c) magnified peak position of (111) peak. Similar to figure 1, all specimens exhibit

a single perovskite structure without any secondary phase. Also, it seems that all

specimens have mixture of tetragonal and monoclinic phase. Mainly tetragonal

structure. The peak position of (110) was moved to a higher angle by increasing the

sintering temperature up to 1200 °C. The (110) peak positions were 31.32, 31.34, 31.40

and 31.42° at sintering temperatures of 1125, 1150, 1175 and 1200 °C, respectively.

However, as increasing the sintering temperature above 1200 °C, the (110) peak

positions were decreased from 31.42 to 31.36°. The shift of (110) peak position to

higher angles implies that the lattice parameter of the BS-PT ceramics with Bi excess

was decreased with increasing sintering temperature up to 1200 °C, and then increased.

This indicates that BS-PT ceramics with Bi excess sintered at 1225 °C has increased

8
lattice parameter of a and b. Therefore, we believe lattice parameter of c was decreased

by considering the Poisson’s ratio. Due to the decreased lattice parameter of c, BS-PT

ceramics with Bi excess sintered at 1225 °C has decreased piezoelectric properties. Up

to sintering temperature of 1200 °C, lattice parameter a and b were decreased. Lower

lattice parameter of a and b means increased lattice parameter of c due to Poisson’s

ratio, therefore, increased lattice parameter c can improve piezoelectric properties. As a

result, we can expect the BS-PT ceramics with Bi excess sintered at 1200 °C have the

longest lattice parameter c, which are related to piezoelectric properties.

Figure 4 shows FE-SEM (Field emission - scanning electron microscopy)

images of the specimens. The FE-SEM surface images of the BS-PT ceramics with 0 –

0.02 mol Bi excess were measured just after the sintered sample without any polishing

but blowing process was carried out to remove the particles. The specimens were

sintered at the temperature from 1125 to 1225 °C. The grain size of the specimens was

measured and calculated one by one from the FE-SEM images. Average grain size and

standard deviation was calculated and estimated for each compositions and sintering

conditions. To further increase statistically accuracy, we have measured several times

for each particle. The calculated average grain size and standard derivation of BS-PT

ceramics with Bi excess for the sintering temperature is displayed in the table 2. As

shown in the figure, the grain size of the BS-PT ceramics with Bi excess was gradually

increased in all compositions with increasing sintering temperature. The sintering

temperature-dependent grain size can be expressed by the following equation [21]:

= ᆞ ! −#/%& (2)

Here, G is the average grain size (µm), t is the sintering time (h), n is the kinetic grain

growth exponent, K0 is the constant related with grain size and sintering time, Q is the

9
apparent activation energy (kJ/mol), R is the gas constant (8.314 J/mol·K) and T is the

absolute temperature (K) for sintering process, respectively. To calculate the value of n,

the equation (2) can be expressed in the form:

log = 1/* log + 1/* [log − 0.434 #/%& ] (3)

Here, approximate number 0.434 is log(e) rounded to 3 decimals. From the slope (1/n)

of logG versus logt in the equation (3), the kinetic grain growth exponent can be

determined. Generally, the slopes are equal to 1/3 at the high sintering temperature,

which the kinetic grain growth exponent is 3 [21]. Also, this exponent of 3 was

confirmed our previous research, which based on the sintering time dependent grain size

effects, in the experiment, the grain size can be calculated and estimated. The simulated

exponent value is around 2.9 [22]. In order to determine the activation energy with

different sintering temperature, the following equation (4) can be obtained:


1
log / = log − 0.434 #/%& (4)

From the slope of −0.434 ∙ Q/R versus 1/T equation, the activation energy can be

determined in the figure 5. The calculated activation energies Q of BS-PT ceramics with

0 – 0.02 mol Bi excess were around 90.5, 64.9, 59.0, 58.1, and 66.9 kJ/mol, respectively.

Also, from the ordinate in equation (4), logK0 of BS-PT ceramics with 0 – 0.02 mol Bi

excess can be calculated. The calculated logK0 were 32.9, 23.6, 21.3, 21.0 and 23.8,

respectively. The activation energies of the stoichiometric BS-PT ceramics were higher

than those of the BS-PT ceramics with Bi excess. This indicates that required energy for

the sintering process was decreased by employing the pre-addition of Bi.

According the equation (2) and from our experimental data, the grain size was

increased as increasing the sintering temperature. At starting sintering temperature of

1125 °C, BS-PT ceramics with Bi excess have grain size of 0.4 ~ 0.8 µm. However, as

10
increasing sintering temperature up to 1200 °C, BS-PT ceramics with Bi excess have

dense microstructure and uniform grain size. We believe these uniform grain size and

dense microstructures are co-related with the lattice parameters which were discussed in

the figure 3. Due to increased density and uniform grain size, the lattice parameter was

decreased. According to figure 6, the bulk density of BS-PT ceramics with 0.01 mol Bi

excess was higher than that of the stoichiometric BS-PT ceramics. Therefore, excess Bi

addition for the BS-PT ceramics have a higher bulk density due to the dense

microstructure compared with those of the stoichiometric BS-PT ceramics. It indicates

that pre-addition of Bi2O3 can compensate evaporation loss of Bi and improve the

densification in BS-PT ceramics.

Figure 6 displays three-dimensional plots of bulk density of BS-PT ceramics

with 0 – 0.02 mol Bi excess sintered from 1125 to 1225 °C. In all compositions, the

densification was increased with increasing the sintering temperature up to 1200 °C,

and then decreased. The decreasing of bulk density at sintering temperature above 1200

°C results from the increased lattice parameters. By observing the sintering temperature

dependent XRD data, the (111) peak position of 1225 °C sintered specimen was shifted

to the lower angle compared with those of sintered other temperatures, it means lattice

parameters can be increased as increasing the sintering temperature up to 1225 °C.

Therefore, we believe the density was decreased due to increased lattice parameters.

Also, when the stoichiometric BS-PT ceramics was sintered at 1200 °C, the bulk

density reached 7.41 (g/cm3). As increasing the Bi excess content up to 0.01 mol, the

bulk density was increased, and then decreased. The BS-PT ceramics with 0.01 mol Bi

excess have the highest bulk density of 7.61 (g/cm3) when sintered at 1200 °C. The bulk

density of the BS-PT ceramics with 0.01 mol Bi excess was higher than that of the

11
stoichiometric BS-PT ceramics. This higher densification of ceramics can improve their

structural stability and decrease the occurrence of defects and unstable phases. As a

result, we expected that the BS-PT ceramics with Bi excess show improved the

dielectric and piezoelectric properties.

Figure 7 shows three-dimensional plots of dielectric constant of the BS-PT

ceramics with 0 – 0.02 mol Bi excess sintered from 1125 to 1225 °C at 1 kHz. The

figures show that the dielectric constant was increased with increases in sintering

temperature up to 1200 °C, then decreased in all compositions. In addition, as

increasing the Bi excess content up to 0.01 mol, the dielectric constant was increased,

and then decreased. The BS-PT ceramics with 0.01 mol Bi excess have the highest

dielectric constant of 1455 when it was sintered at 1200 °C. The dielectric constant of

the BS-PT ceramics with 0.01 mol Bi excess was higher than that of the stoichiometric

BS-PT ceramics. The increased dielectric constant may be attributed to the increase of

densification, which compensates for the Bi evaporation loss of the BS-PT ceramics.

Figure 8 shows the dielectric constants and dielectric losses of BS-PT ceramics

with 0 – 0.02 mol Bi excess sintered at 1200 °C depending on the frequency range. As

mentioned in figure 7, the dielectric constant was increased, and then decreased as

increasing the Bi excess content up to 0.01 mol. The dielectric constant of BS-PT

ceramics with 0.01 mol Bi excess is higher than that of other compositions, which have

a maximum value of 1455. Also, the dielectric losses of BS-PT ceramics with 0 – 0.02

mol Bi excess were approximately 0.039, 0.039, 0.032, 0.039 and 0.043 at 1kHz,

respectively. We expected that the lowest dielectric loss of BS-PT ceramics with 0.01

mol Bi excess result from effectively compensation for the Bi loss.

12
Figure 9 shows the temperature dependence of the dielectric constant of BS-PT

ceramics with 0 – 0.02 mol Bi excess sintered at 1200 °C. The Curie temperature, Tc,

for BS-PT with 0.005 – 0.02 mol Bi excess are 397.5, 400.2, 394.6, and 389.6 while

that of the stoichiometric BS-PT ceramics has 393.2 °C, respectively. As increasing the

Bi excess content up to 0.01 mol, the maximum dielectric constant and Curie

temperature of BS-PT ceramics were increased, and then decreased. By introducing

excess Bi addition, BS-PT ceramics have increased Curie temperature compared with

those of the stoichiometric BS-PT ceramics.

Figure 10 displays three-dimensional plots of piezoelectric charge coefficient

(d33) of BS-PT ceramics with 0 – 0.02 mol Bi excess sintered from 1125 to 1225 °C.

The piezoelectric charge coefficient of BS-PT ceramics with 0 – 0.02 mol Bi excess

sintered at 1200 °C were 417, 432, 452, 427 and 416, respectively. As shown in figure,

BS-PT ceramics with 0.01 mol Bi excess showed the highest piezoelectric charge

coefficient of 452 pC/N. The variation in the piezoelectric charge coefficient was

similar to that in the densification data, which was discussed in figure 6. Therefore, the

increased piezoelectric charge coefficient, d33, results from the increase of densification

and the elimination of the defects, which enhance the poling condition and reduce the

leakage current. By adding excess 0.01 mol Bi to BS-PT ceramics, the piezoelectric

charge coefficient was enhanced from 417 to 452 pC/N.

Figure 11 shows three-dimensional plots of electromechanical coupling factor

(kp) of BS-PT ceramics with 0 – 0.02 mol Bi excess sintered from 1125 to 1225 °C. As

shown in the figure, the highest value of electromechanical coupling factor, kp, was

52.25 %, which is higher than that of the stoichiometric BS-PT ceramics. By

introducing excess Bi addition of BS-PT ceramics, the electromechanical coupling

13
factor, kp, was improved from 51.48 to 52.25 %. These results of piezoelectric

properties mean that BS-PT ceramics with Bi excess can be another candidate material

for piezoelectric devices including energy harvesters. Even though kp is not so high

around 52.25 %, but BS-PT ceramics with Bi excess has Curie temperature reach up to

450 °C and high de-poling temperature of 350 °C [23].

In order to research the effect of the excess Bi addition on the ferroelectric

properties of the BS-PT ceramics, the polarization versus electric field hysteresis loops

were measured at room temperature. Figure 12 shows P-E hysteresis loops and

remanent polarization (Pr) of BS-PT ceramics with 0 – 0.02 mol Bi excess sintered at

1200 °C. According to figure 12, the hysteresis loops were clearly observed with a

rectangular shape for all specimens. Additionally, the behavior of the remanent

polarization was similar to that of the piezoelectric charge coefficient, which was

discussed in figure 10. In increasing the Bi excess content up to 0.01 mol, the remanent

polarization, Pr, was increased from 31.4 to 37.5 μC/cm2, and then decreased.

Figure 13 shows the open circuit voltages of BS-PT ceramics with 0 – 0.02 mol

Bi excess sintered at 1200 °C for the piezoelectric energy harvesters. The output

voltages were measured by oscilloscope when the mechanical force was applied to the

BS-PT ceramic with Bi excess. An external mechanical force of 350 N was applied at a

speed of 350 mm/min. The sample has diameter of 10.0 mm. And the thickness of all

specimens is around 1 mm. Therefore, by considering the applied external mechanical

force, the applied stress for the specimen is around 445.9 N/cm2, which corresponds to

4.5 MPa. The stoichiometric BS-PT ceramics showed the output circuit voltage of 19.69

V. Also, the measured open circuit voltage of BS-PT ceramics with 0.01 mol Bi excess

was 26.93 V, which was higher than that of the stoichiometric BS-PT ceramics.

14
Figure 14 shows the generated output power density of BS-PT ceramics with 0 –

0.02 mol Bi excess sintered at 1200 °C. The generated output power density is

calculated by integrating the multiplication of the output voltage and current during the

generation time. The generated output power densities for the BS-PT ceramics with 0 –

0.02 mol Bi excess for the piezoelectric energy harvesters were 0.28, 0.40, 0.48, 0.31

and 0.07 mW/cm3, respectively. Piezoelectric charge coefficient d33 for BS-PT with 0 –

0.02 mol Bi excess are 417, 432, 452, 427 and 416, respectively. By considering

relative dielectric permittivity εr, BS-PT with 0 – 0.02 mol Bi excess have 1384, 1416,

1455, 1411 and 1380, respectively. Therefore, FOM value (FOM=d33×g33) of BS-PT

specimen with 0.02 mol Bi excess has minimum value among the specimens. Since the

measuring system has threshold energy to operate and measure the data, certain amount

of power should be used for offset energy. Therefore, even though BS-PT ceramic with

0.02 mol Bi excess have just little lower d33 values, the obtained energy can be lower

level compared with other specimens. The increased open circuit voltage and generated

output power density may come from the increase of piezoelectric and dielectric

properties, including the piezoelectric charge coefficient and dielectric permittivity, due

to the complementation of Bi vacancies by introducing pre-addition of volatile

components to the BS-PT system. The BS-PT ceramics have a high Curie temperature

of 450 °C. Near the Curie temperature region, the material cannot be employed. Only

less than 50 % of Curie temperature region can be employed to operate the devices. In

this point of view, high Curie temperature can be unique characteristics for devices to

apply in high quality and military application. To know the merit of BS-PT ceramics

with Bi excess, other electrical properties were collected and compared in the table 2.

The various properties of lead-based piezoelectric energy harvesters based on different

15
piezoelectric materials are investigated and sorted as table 2 [24-31]. As shown in the

table 2, BS-PT piezoelectric ceramics with Bi excess have relatively high voltage of

26.9 V and power density of 0.48 mW/cm3 by comparing the output voltage and power

density of different piezoelectric ceramics. And moreover, BS-PT with Bi excess has

Curie temperature reach up to 400 °C and high de-poling temperature of 350 °C. Due to

this high Curie temperature and high de-poling temperature, BS-PT ceramics with Bi

excess have very versatile applications.

16
4. Conclusions

In this study, energy harvesters based on the BS-PT ceramics with Bi excess

were successfully fabricated by employing the conventional ceramic process. The

dielectric and piezoelectric properties of the BS-PT ceramics with Bi excess were

analyzed and discussed. By optimizing the sintering temperature and excess Bi content

from the BS-PT ceramics, the highest piezoelectric charge coefficient of 452 pC/N,

electromechanical coupling factor of 52.25 %, generated open circuit voltage of 26.93

V, and generated power density of 0.48 mJ/cm3 were obtained in the BS-PT ceramics

with 0.01 mol Bi excess sintered at 1200 °C. The increased piezoelectric charge

coefficient may be attributed to the increase of densification and tetragonal properties

due to the compensation of Bi which was the loss of BS-PT ceramics during the

sintering process. The Rietveld refinement was performed to confirm that the addition

Bi excess in BS-PT ceramics increased the tetragonal phase, which means growth of the

c-axis orientation. The (00l) peak ratio of BS-PT ceramics with Bi excess was higher

than that of the stoichiometric BS-PT ceramics. As a result, the BS-PT ceramics with

0.01 mol Bi excess show higher generated power density than that of the stoichiometric

BS-PT ceramics. We demonstrate that excess Bi addition can enhance the piezoelectric

properties of BS-PT ceramics by compensating for the evaporation loss of Bi. The BS-

PT ceramics with 0.01 mol Bi excess are promising candidate materials for piezoelectric

energy harvesting applications.

17
Acknowledgements

This research was supported by Basic Science Research Program through the National

Research Foundation of Korea (NRF) funded by the Ministry of Education, Science

and Technology(grant number 2019R1F1A1060546)and supported by the Human

Resources Development (No.20184030202070) of the Korea Institute of Energy

Technology Evaluation and Planning (KETEP) grant funded by the Korea government

Ministry of Trade, Industry and Energy.

18
References

[1] B. Jaffe, R. S. Roth, S. Marzullo, “Piezoelectric Properties of Lead Zirconate‐Lead

Titanate Solid‐Solution Ceramics”, J. Appl. Phys. (1954) 25, 809-810.

[2] T. R. Shrout, S. J. Zhang, Lead-free piezoelectric ceramics: Alternatives for PZT?, J.

Electroceram. 19 (2007) 111-124.

[3] E. Cross, “Materials science: Lead-free at last”, Nature. (2004) 432, 24-25.

[4] A. Sehirlioglu, A. Sayir and F. Dynys, “Doping of BiScO3–PbTiO3 Ceramics for

Enhanced Properties”, J. Am. Ceram. Soc. (2010) 93 [6], 1718-1724.

[5] Z. Liu, C. Zhao, J. F. Li, K. Wang and J. Wu, “Large strain and temperature-

insensitive piezoelectric effect in high-temperature piezoelectric ceramics”, J. Mater.

Chem. C. (2018) 6, 456-463.

[6] J. H. Ahn, D. J. Shin and J. H. Koh, “Comparative study on the thickness dependent

output energy for (Bi,Sc)O3-(Pb,Ti)O3 multilayered structure”, Ceram. Int. (2017) 43,

S643-S648.

[7] Z. Liu, C. Zhao, R. Xie and J. Wu, “Tailored electrical properties in ternary BiScO3-

PbTiO3 ceramics by composition modification”, Ceram. Int. (2018) 44, 8057-8063.

[8] Z. Liu, B. Wu and J. Wu, “Reduced dielectric loss and high piezoelectric constant in

Ce and Mn co-doped BiScO3-PbCexTi1-xO3-Bi(Zn0.5Ti0.5)O3 ceramics”, Ceram. Int.

(2018) 44, 16483-16488.

[9] D. H. Kang and Y. H. Kang, “Dielectric and Pyroelectric Properties of Lead-Free

Sodium Bismuth Titanate Thin Films Due to Excess Sodium and Bismuth Addition”, J.

Microelectron. Packag. Soc. (2013) 20, 25-30.

19
[10] B. T. Anna and B. Monika, “Comparative studies in subsolidus areas of ternary

oxide systems PbO–V2O5–In2O3 and PbO–V2O5–Fe2O3”, J. Therm. Anal. Calorim.

(2013) 113, 137-145.

[11] J. Konig, M. Spreitzer, B. Jancar, D. Suvorov, Z. Samardzija and A. Popovic, “The

thermal decomposition of K0.5Bi0.5TiO3 ceramics”, J. Eur. Ceram. Soc. (2009) 29, 1695-

1701.

[12] A. Popovic, L. Bencze, J. Koruza and B. Malic, “Vapour pressure and mixing

thermodynamic properties of the KNbO3–NaNbO3 system”, RSC Adv. (2015) 5, 76249-

76256.

[13] J. H. Ji, J. Kim and J. H. Koh, “Improved dielectric and piezoelectric properties of

K/Na excessed (Na,K)NbO3 lead-free ceramics by the two step sintering process”, J.

Alloys Compd. (2017) 698, 938-943.

[14] Y. Lee, J. Cho, B. Kim and D. Choi, Piezoelectric properties and densification

based

on control of volatile mass of potassium and sodium in (K0.5Na0.5)NbO3 ceramics, Jpn.

J. Appl. Phys. (2008) 47, 4620-4622.

[15] J. Fang, X. Wang, Z. Tian, C. Zhong and L. Li, “Two-Step Sintering: An Approach

to Broaden the Sintering Temperature Range of Alkaline Niobate-Based Lead-Free

Piezoceramics”, J. Am. Ceram. Soc. (2010) 93 [11], 3552-3555.

[16] A. Popovic, L. Bencze, J. Koruza, B. Malic and M. Kosec, “Knudsen effusion

mass spectrometric approach to the thermodynamics of Na2O–Nb2O5 system”, Int. J.

Mass spectrom. (2012) 309, 70-78.

20
[17] S. Sasikumar, R. Saravanan, K. Aravinth, “Piezoelectric and ferroelectric

properties of lead-free (1-x)(Na1−yKy)(Nb1−zSbz)O3-xBaTiO3 solid solution”, Physica B.

(2017) 512, 58-67.

[18] I. H. Chan, C. T. Sun, M. P. Houng and S. Y. Chu, “Sb doping effects on the

piezoelectric and ferroelectric characteristics of lead-free Na0.5K0.5Nb1-x SbxO3

piezoelectric ceramics”, Ceram. Int. (2011) 37, 2061-2068.

[19] G. Lee, J. H. Ji and J. H. Koh, “Enhanced piezoelectric properties of (Bi,Na)TiO3–

(Bi,K)TiO3 ceramics prepared by two-step sintering process”, Int. J. Appl. Ceram.

Technol. (2018) 15, 531-537.

[20] J. H. Ahn and J. H. Koh, “Enhanced piezoelectric properties of (Bi,Sc)O3-

(Pb,Ti)O3 ceramics by optimized calcination process”, J. Alloys Compd. (2016) 689,

138-144.

[21] T. Senda and R. Bradt, “Grain Growth in Sintered ZnO and ZnO-Bi2O3 Ceramics”,

J. Am. Ceram. Soc. (1990) 73 [1], 106-114.

[22] J. H. Ji, D. H. Kim, B. G. Choi, K. Lee, C. W. Kim, S. K. Lee, Y. H. Ko, K. H.

Cho and J. H. Koh, “Improved Piezoelectric Properties of (Bi,Sc)O3–(Pb,Ti)O3

Ceramics Based on the Two-Step Sintering Process”, J. Nanoelectron. Optoelectron.

(2019) 14, 741-745.

[23] T. Stevenson, D. G. Martin, P. I. Cowin, A. Blumfield, A. J. Bell, T. P. Comyn,

and P. M. Weaver, “Piezoelectric materials for high temperature transducers and

actuators”, J Mater Sci: Mater Electron (2015) 26, 9256-9267.

[24] D. A. Berlincourt, C. Cmolik and H. Jaffe, “Piezoelectric Properties of

Polycrystalline Lead Titanate Zirconate Compositions”, Proc. IEEE (1960), 48, 220-229.

21
[25] J. Xie, X. P. Mane, C. W. Green, K. M. Mossi and KAM K. Leang, “Performance

of Thin Piezoelectric Materials for Pyroelectric Energy Harvesting”, J. Intell. Mater.

Syst. Struct. (2010) 21, 243-249.

[26] D. F. Berdy, P. Srisungsitthisunti, B. Jung, X. Xu, J. F. Rhoads and D. Peroulis,

“Low-Frequency Meandering Piezoelectric Vibration Energy Harvester”, IEEE Trans.

Ultrason. Eng. (2012) 59, 846-858.

[27] C. N. Xu, M. Akiyama, K. Nonaka and T. Watanabe, “Electrical Power Generation

Characteristics of PZT Piezoelectric Ceramics”, IEEE Trans. Ultrason. Eng. (1998) 45,

1065-1070.

[28] S. B. Seo, S. H. Lee, C. B. Yoon, G. T. Park and H. E. Kim, “Low-Tempcraturc

Sintering and Piezoelectric Properties of 0.6Pb(Zr0.47Ti0.53)O3.0.4Pb(Zn1/3Nb2/3)O3

Ceramics”, J. Am. Ceram. Soc. (2004) 87[7], 1238-1243.

[29] J. Wu, H. Shi, T. Zhao, Y. Yu and S. Dong, “High-Temperature BiScO3-PbTiO3

Piezoelectric Vibration Energy Harvester”, Adv. Funct. Mater. (2016) 26, 7186-7194.

[30] M. Pham-Thi, C. Augier, H. Dammak and P. Gaucher, “Fine grains ceramics of

PIN–PT, PIN–PMN–PT and PMN–PT systems: Drift of the dielectric constant under

high electric field”, Ultrasonics. (2006) 44, e627-e631.

[31] N, Zhong, X. L. Dong, D. Sun, P. Xiang, H. Du, “Electrical properties of

Pb(Mg1/3Nb2/3)O3–PbTiO3 ceramics modified with WO3”, Mater. Res. Bull. (2004) 37,

175-184.

22
Table 1. The refined parameters of BS-PT ceramics with 0 – 0.02 mol Bi excess

obtained from Rietveld refinements.

R factors (%)
Bi excess content (mol) Space group Phase fraction (%)
Rp Rwp χ2

0 P4mm 28.5
7.37 9.47 1.16
Cm 71.5

0.005 P4mm 31.8


7.68 9.75 1.21
Cm 68.2

0.01 P4mm 36.4


7.91 8.93 1.12
Cm 63.6

0.015 P4mm 39.1


7.84 9.79 1.16
Cm 60.9

0.02 P4mm 42.6


9.10 11.56 1.44
Cm 57.4

23
Table 2. Grain size of BS-PT ceramics with 0 – 0.02 mol Bi excess sintered from 1125

to 1225 °C.

Sinter. Bi content
Temp. 0 mol 0.005 mol 0.01 mol 0.015 mol 0.02 mol
1225 °C 3.9 ± 1.4 µm 2.8 ± 0.5 µm 2.3 ± 1.3 µm 2.3 ± 1.0 µm 1.5 ± 0.8 µm

1200 °C 3.0 ± 1.5 µm 2.0 ± 0.8 µm 1.7 ± 0.7 µm 1.7 ± 0.6 µm 1.3 ± 0.5 µm

1175 °C 1.0 ± 0.4 µm 1.4 ± 0.5 µm 1.2 ± 0.3 µm 1.2 ± 0.8 µm 1.1 ± 0.6 µm

1150 °C 0.9 ± 0.4 µm 0.9 ± 0.6 µm 0.9 ± 0.1 µm 0.8 ± 0.5 µm 0.8 ± 0.5 µm

1125 °C 0.8 ± 0.3 µm 0.8 ± 0.4 µm 0.8 ± 0.1 µm 0.8 ± 0.5 µm 0.4 ± 0.1 µm

24
Table 3. Comparison of various properties of lead-based piezoelectric energy harvesters

based on different piezoelectric materials.

Piezoelectric Output performance


d33 (pC/N) kp (%) Reference
materials Voltage (V) Power density

PZT 223 52.9 - - [22]

PZT - - 0.53 0.28 µW/cm2 [23]

PZT (cantilever) 390 - - 0.2 mW/cm3 [24]

Mn doped PZT - 36 - 0.06 mJ/cm2 [25]

PZT-PZN 460 60 - - [26]

BS-PT 450 54 12.2 0.1 µW/cm3 [27]

PMN-PT 499 52 - - [28]

PMN-PT - - 0.57 0.33 µW/cm2 [23]

W doped PMN-PT 466 56.4 - - [29]

BS-PT
452 52.3 26.9 0.48 mW/cm3 This work
with Bi excess

25
0.36BS-0.64PT 0.020 mol Bi1O1.5 excess
(a) 0.015 mol Bi1O1.5 excess
0.010 mol Bi1O1.5 excess
Intensity (arb.units)

110
0.005 mol Bi1O1.5 excess
BS-PT ceramic
001

111

112
102
002

121
200

220
210
100

20 30 40 50 60 70
2θ (deg.)

0.020 mol Bi1O1.5 excess


(b) 0.36BS-0.64PT
0.015 mol Bi1O1.5 excess
110
Intensity (arb.units)

0.010 mol Bi1O1.5 excess


0.005 mol Bi1O1.5 excess
BS-PT ceramic
001

111
100

112
002

102
200

220
121
210

20 30 40 50 60 70
2θ (deg.)

26
18.0 (c) BS-PT ceramic with Bi1O1.5 excess
Peak ratio of (00l) (%)
BS-PT ceramic

17.5

17.0

16.5

16.0
0.000 0.005 0.010 0.015 0.020
BiO1.5 content (mol)

23.0
(d) BS-PT ceramic with Bi1O1.5 excess
Peak ratio of (00l) (%)

22.5 BS-PT ceramic

22.0

21.5

21.0

20.5

20.0
0.000 0.005 0.010 0.015 0.020
BiO1.5 content (mol)

Fig 1. The X-ray diffraction patterns and peak ratio of (00l) of BiScO3-PbTiO3 ceramics

with 0 – 0.02 mol Bi excess sintered at 1200 °C. (a), (c): without poling and (b), (d):

with poling

27
Fig 2. The Rietveld refinement results of (a) stoichiometric BiScO3-PbTiO3 ceramic and

(b) BiScO3-PbTiO3 ceramic with 0.01 mol Bi excess using the High Score Plus

program.

28
(a) BS-PT with 0.01 mol Bi O1 1.5 excess 1225 °C (b) (c)
1200 °C
1175 °C
Intensity (arb.units)

110

110
1150 °C
110

1125 °C
001

111

002

112
102
100

200

220
121
210

20 30 40 50 60 70 31 32 39
2θ (deg.)

Fig. 3. (a) The X-ray diffraction patterns of BiScO3-PbTiO3 ceramics with 0.01 mol Bi

excess sintered from 1125 to 1225 °C, (b) magnified peak position of (110) peak and (c)

magnified peak position of (111) peak.

29
Fig. 4. FE-SEM (Field Emission Scanning Electron Microscopy) images of the surface

of BiScO3-PbTiO3 ceramics with 0 – 0.02 mol Bi excess sintered from 1125 to 1225 °C.

30
3 BS-PT ceramic (Q = 90.56 kJ/mol)
0.005 mol Bi1O1.5 excess (Q = 64.92 kJ/mol)
0.010 mol Bi1O1.5 excess (Q = 59.01 kJ/mol)
2 0.015 mol Bi1O1.5 excess (Q = 58.14 kJ/mol)
0.020 mol Bi1O1.5 excess (Q = 66.91 kJ/mol)
Log (G /t)
3

-1

6.6 6.7 6.8 6.9 7.0 7.1 7.2


4 -1
10 /T (K )

Fig. 5. Temperature dependent grain size of BiScO3-PbTiO3 ceramics with 0 – 0.02 mol

Bi excess sintered from 1125 to 1225 °C.

31
8.0

cm )
nsity (g/ 3
7.8
7.614
7.537 7.6
7.460
7.4
7.383

Bulk De
7.306 7.2
7.229
20 7.0
7.152 0.0
12
15 )
25
7.075 0 .0
(°C
Bi 1

ure
12
00

10
O 1.

6.998 0.0
at
per
11
5

75
con

05 m
0.0
Te
11
ten

50

ng
t

eri
11

00
(m

0.0 t
25

n
ol)

Si

Fig. 6. Three-dimensional plots of bulk density of BiScO3-PbTiO3 ceramics with 0 –

0.02 mol Bi excess sintered from 1125 to 1225 °C.

32
150
1455 0

t
onstan
1433 145
0
1410
14
00

tric C
1388
13
1366 50

Dielec
1343 13
00

1321 0.0
20 12
5
12 0
1298 15 ) 25
Bi 1

0.0
(°C
O 1.

12

1276
re
00

10
0.0
tu
5

11
con

era
75

05
p
ten

0.0
11

em
50
t

T
(m

11

00
ng
25

0.0 i
ol)

r
Si nte

Fig. 7. Three-dimensional plots of dielectric constant of BiScO3-PbTiO3 ceramics with

0 – 0.02 mol Bi excess sintered from 1125 to 1225 °C at 1 kHz.

33
0.8
0.020 mol Bi1O1.5 excess
1600 0.015 mol Bi1O1.5 excess
Dielectric constant (εr)

0.010 mol Bi1O1.5 excess


0.005 mol Bi1O1.5 excess 0.6

Dielectric loss
BS-PT ceramic
1400

0.4

1200

0.2

1000

0.0
1k 10k 100k 1M
Frequency (Hz)

Fig. 8. The dielectric constants and dielectric losses of BiScO3-PbTiO3 ceramics with 0

– 0.02 mol Bi excess sintered at 1200 °C depending on the frequency range.

34
40k BS-PT ceramic
0.005 mol Bi1O1.5 excess
0.010 mol Bi1O1.5excess Tm= 400.2°
Dielectric constant

30k 0.015 mol Bi1O1.5 excess


0.020 mol Bi1O1.5 excess Tm= 393.2°

Frequency: 1 kHz
20k

10k

0
100 200 300 400
Temperature (°C)

Fig. 9. The temperature dependence of the dielectric constant of BiScO3-PbTiO3

ceramics with 0 – 0.02 mol Bi excess sintered at 1200 °C.

35
480
452.0

439.8
440
427.5

/N)
415.3

d3 (pC
40
0
403.0

3
36
0
390.8 20
0.0
12
378.5
0.0
15 25
)
°C
Bi 1

(
12

re
00

10
O 1.

366.3
0.0 u
t
11

era
5

75
con

354.0 05
p
11

0.0
m
ten

50

Te
11

00
t (m

0.0 ng
25

ri
te
ol)

Si n
Fig. 10. Three-dimensional plots of piezoelectric charge coefficient (d33) of BiScO3-

PbTiO3 ceramics with 0 – 0.02 mol Bi excess sintered from 1125 to 1225 °C.

36
52.26 54
51.79
51.32 52

kp (%)
50.85
50
50.38
49.92 48
2 0
0.0
49.45 12 )
0.0
15 25
(°C
48.98
re
Bi 1

12

tu
00

10
O 1.

era
48.51 0.0
11

p
5

75
con

05 m
Te
11

0.0
ten

50

ing
t (m

r
11

00 e
0.0 t
25

Sin
ol)

Fig. 11. Three-dimensional plots of electromechanical coupling factor (kp) of BiScO3-

PbTiO3 ceramics with 0 – 0.02 mol Bi excess sintered from 1125 to 1225 °C.

37
Fig. 12. P-E hysteresis loops and remanent polarization (Pr) of BiScO3-PbTiO3 ceramics

with 0 – 0.02 mol Bi excess sintered at 1200 °C. (a) 0, (b) 0.005, (c) 0.01, (d) 0.015, (e)

0.02 and, (f) remanent polarization (Pr) at 3 kV

38
Fig. 13. The open circuit voltage for BiScO3-PbTiO3 ceramics with 0 – 0.02 mol Bi

excess sintered at 1200 °C. (a) 0, (b) 0.005, (c) 0.01, (d) 0.015, (e) 0.02 and, (f) peak

value of the open circuit voltage

39
Power density (mW/cm )
2

0.6

0.4

0.2

0.0
0.000 0.005 0.010 0.015 0.020
Bi1O1.5 content (mol)

Fig. 14. Generated output power density of BiScO3-PbTiO3 ceramics with 0 – 0.02 mol

Bi excess sintered at 1200 °C.

40
14 October, 2019

Dear editor

This letter is to confirm manuscript Re-submission to ‘Ceramics International’. This


manuscript has 40 pages including 3 tables and 14 figures.
The English expression of this manuscript was consulted by the English Language Editing
Company ‘Editages’

Title: ‘BiScO3-PbTiO3 piezoelectric ceramics with Bi excess for energy harvesting applications under
high temperature’

Declare of interest statement

We can declare that we have no interest with other research group and commercial company
groups.

In this study, we have investigated the Bi excessed BS-PT piezoelectric ceramics for energy
harvester applications. Due to volatile component of Bi, the BS-PT ceramics have Bi-deficient
composition. To solve the problem of Bi deficiency problem in the BS-PT ceramics, excess Bi
addition process was carried out during the fabrication process. This Bi excess BS-PT process has
not been yet reported. By adding excessive Bi dopants to BS-PT, the piezoelectric properties were
improved. In optimized sintering temperature and excess Bi content to BS-PT ceramics, the highest
piezoelectric charge coefficient of 452 pC/N, energy density of 0.48 mJ/cm3 were obtained. These
piezoelectric charge coefficient and energy densities are higher than those of the conventionally
fabricated BS-PT ceramics.

Author:
Jung-Hyuk Koh
School of Electrical and Electronic Engineering, Chung-Ang University
Zip-Code 156-756.
84 Heukseok-Ro, DongJak-Gu, Seoul, Korea.

1
Co-author
Jae-Hoon Ji, Jinhwan Kim
School of Electrical and Electronic Engineering, Chung-Ang University, Seoul, Korea

Dong-Jin Shin
Korea Electrotechnology Research Institute, Changwon, Korea

Looking forward to successful submission


With regards,

Corresponding author
Prof. Jung-Hyuk Koh (jhkoh@cau.ac.kr)
School of Electrical and Electronic Engineering
Chung-Ang University

You might also like