• Tidak ada hasil yang ditemukan

Journal of Alloys and Compounds - CSDL Khoa học

N/A
N/A
Nguyễn Gia Hào

Academic year: 2023

Membagikan "Journal of Alloys and Compounds - CSDL Khoa học"

Copied!
10
0
0

Teks penuh

(1)

Enhancement in dielectric, ferroelectric, and piezoelectric properties of BaTiO 3 - modi fi ed Bi 0.5 (Na 0.4 K 0.1 )TiO 3 lead-free ceramics

Le Dai Vuong

a,*

, Phan Dinh Gio

b

aHue Industrial College, Viet Nam

bUniversity of Sciences, Hue University, Viet Nam

a r t i c l e i n f o

Article history:

Received 30 July 2019 Received in revised form 15 October 2019 Accepted 23 October 2019 Available online 24 October 2019

Keywords:

Ceramics Ferroelectrics Piezoelectricity Energy storage materials Crystal structure

a b s t r a c t

In this paper, the effect of BaTiO3on dielectric, ferroelectric, and piezoelectric properties was observed in (1-x)[Bi0.5(Na0.4K0.1)TiO3]exBaTiO3ceramics (BNKT-BT, withx¼0; 0.01; 0.02; 0.03 and 0.04). The phase formation and microstructure of BNKT-BT ceramics were examined by X-ray diffraction (XRD), the Raman scattering spectra, and scanning electron microscopy (SEM). A pure perovskite phase was observed in the ceramic samples, proving the dissolution of BaTiO3into the lattice structure of the BNKT to form a homogeneous solid solution (x0.04). Atx¼0.02, the best physical properties of the BNKT-BT ceramics, such as density,r¼5.86 g/cm3(98.5% of the theoretical density); electromechanical coupling factors (kp), 0.30; (kt), 0.31; remanent polarization (Pr), 17.2mC cm2; dielectric constant (εr), 1354; and highest dielectric constant (εmax), 4037, were obtained. Further, the energy storage properties of 0.98Bi0.5(Na0.4K0.1)TiO3e0.02BaTiO3ceramics were studied and compared with the textured ceramics (the same composition), which were formed due to the high degree of orientation obtained by the template grain growth method. The textured ceramics showed lower remnant polarization, while the piezoelectric coefficient (d31) and electromechanical coupling coefficients (kpandkt) of the textured ceramics were found to increase by about 8%, 13.3%, and 19.4%, respectively.

©2019 Elsevier B.V. All rights reserved.

1. Introduction

In recent years, there has been a considerable increase in the interest on the application of Pb(Zr(1-x)Tix)O3(PZT)-based ceramics owing to their excellent piezoelectric properties and applications in piezoelectric actuators and transformers [1e7]. However, the use of the lead-based ceramics poses serious environmental concerns owing to the high toxicity of lead oxide. Therefore, it is necessary to develop lead-free ceramics with good ferroelectric and piezoelec- tric properties to replace the lead-based ceramics [8e10]. None- theless, lead-free materials still cannot completely replace lead- based materials in all applications especially high-pressure piezo- electric materials, piezoelectric transformers, or ultrasonic motors [3,6,11]. Thus it became highly intriguing to develop new lead-free materials with improved properties that are suitable for the desired industrial applications [11]. Among the various lead-free materials available today, Bi0.5Na0.5TiO3(BNT) [12e17], BaTiO3[11,18e20] and K0.5Na0.5NbO3 (KNN) [8,19,21] based ceramics have been

considered as most promising candidates owing to their excellent electromechanical properties near the morphotropic phase boundary (MPB) [18].Kamnoy et al. [11] added BaTiO3as starting materials in Bi0.5(Na0.8K0.2)0.5TiO3ceramic by using solid-state re- action method. BaTiO3, at a content of 8 mol%, resulted in the highest value of the dielectric constantεmaxof 6700. However, in this traditional ceramic material, the orientation of the particles is random, and the exhibited properties are the average values of all particles in the material. Therefore, texture engineering, the tech- nology used to develop crystallographically textured materials, seems to be promising in the fabrication of ceramic materials with desired properties. Dong et al. [18] reported a high degree of orientation (96%) in the textured 0.88Na0.5Bi0.5TiO3- 0.08K0.5Bi0.5TiO3-0.04BaTiO3ceramics with optimal properties, at 4 wt% of NaNbO3 templates. The optimum properties, such as piezoelectric coefficient (d33) of 185 pC/N, high remanent polari- zation (Pr) of 26.4mC/cm2, low coercivefield (Ec) of 31 kV/cm, and the dielectric constant (εmax) of 3950, were obtained. Recently, Fan et al. [16] prepared (1-x)Bi1/2(Na0.82K0.18)1/2TiO3 e xBi4Ti3O12ce- ramics using conventional solid-state reaction method and found that the normal ferroelectric-to-ergodic relaxor (FE-to-ER)

*Corresponding author.

E-mail address:[email protected](L.D. Vuong).

Contents lists available atScienceDirect

Journal of Alloys and Compounds

j o u rn a l h o m e p a g e :h t t p : / / w w w . e l s e v i e r . c o m / l o c a t e / j a l c o m

https://doi.org/10.1016/j.jallcom.2019.152790 0925-8388/©2019 Elsevier B.V. All rights reserved.

(2)

transition occurs at a low temperature (TF-Rz0C). At a Bi4Ti3O12

content ofx¼9 wt%, ceramics exhibited best electrical properties such as a large signal piezoelectric coefficient,d*33of 485 pm/V, and a small hysteresis,hof 23%.

In this study, the synthesis of (1-x)[Bi0.5(Na0.4K0.1)TiO3]exBa- TiO3(BNKT-BT) lead-free textured ceramics was carried out by the conventional solid-state reaction method and template grain growth method with pure-phase Bi4Ti3O12templates. The textured samples with preferred orientation showed an enhanced electrical property. The goal of texturing was to form BNKT-BT lead-free piezoelectric ceramic with improved physical properties. As such, this study will provide two important information: (i) thefirst, the studied results indicate that BaTiO3enhances the dielectric, ferro- electric, and piezoelectric properties in BNKT-based ceramics; (ii) the second, the BNKT-BT ceramics with the optimized content of BaTiO3 will be selected for fabrication by template grain growth method. Besides, based on the obtained results, (1-x) [Bi0.5(Na0.4K0.1)TiO3]exBaTiO3ceramics are promising candidates for lead-free ceramics for dielectric, ferroelectric, piezoelectric and energy storage applications.

2. Experimental

The general formula of the studied material was (1-x) [Bi0.5(Na0.4K0.1)TiO3]exBaTiO3, wherex¼0.0, 0.01, 0.02, 0.03, and 0.04. The following reagents were used: Bi2O3, TiO2, Na2CO3, K2CO3

and BaCO3(Merck, purity99.5%). In particular, Bi0.5(Na0.4K0.1)TiO3

ceramics based on our previous research formula [12,14].

2.1. Synthesis of BNKT-BT ceramics by the conventional solid-state reaction method

The process of fabrication technology for Random ceramics, as shown inFig. 1. Starting raw materials of Bi2O3, Na2CO3, K2CO3, TiO2 and BaCO3(Fig. 1(a)) were weighed and milled with ultrasound treatment with an electric power of 100 W andfixed frequency of 28 kHz in ethanol medium for 1 h (Fig. 1(b)) [12].

After that, the powders were dried and calcined at 850C for 2 h to synthesize BNKT-BT compounds (Fig. 1(c)). The calcined BNKT-BT powder milled for 20 h (Fig. 1(d)). The ground materials were pressed into disk 12 mm in diameter and 1.5 mm in thick under 100 MPa (Fig. 1(e)). The BNKT-BT samples were sintered at tem- perature 1100C for 2 h (Fig. 1(f)). Then, the sample is treated, covered with electrodes and examined for properties (Fig. 1(g and h)). The BNKT-BT ceramics with the optimized content of BaTiO3

will be selected for fabrication by template grain growth method.

2.2. Synthesis of BNKT-BT ceramics by template grain growth method

As thefirst step, Bi2O3and TiO2were weighed and milled under ultrasound (electric power of 100 W and a fixed frequency of 28 kHz) in ethanol for 1 h. The as-prepared powders were then dried at 100C for 2 h and subsequently calcined at 500C for 2 h in order to obtain amorphous Bi4Ti3O12precursor powder which was subsequently mixed with Na2CO3/K2CO3 eutectic mixture in a sealed alumina crucible and was sintered at 1050C for 2 h char- acteristics of Bi4Ti3O12 template as demonstrated in Fig. 2. The microstructure study revealed a characteristic plate-like shape for the Bi4Ti3O12template after sintering (Fig. 2(a)). The revelation of particle size length in the range of 4e18mm and approximate width of 0.5e1mm (Fig. 2(b)), which originated due to the high anisotropy of Bi4Ti3O12crystal structure (Fig. 2(c)), suggested that the molten salt method was useful not only in producing particles with anisotropy but also for reducing the aggregation of the particles [22].

As the last step, according to the abovementioned stoichio- metric ratio for the reaction, a predetermined amount of Bi4Ti3O12 powder (Fig. 2(a)) and calcined BNKT-BT powder (Fig. 3(a)) were mixed together and milled in a planetary milling machine (PM400/

2-MA-Type) for 2 h (Fig. 3(b)). Then, this mixture was continuously mixed with a binder solution and was subsequently tape cast to form thin sheets with a thickness of about 0.1 mm (Fig. 3(c)) in which the plate-like Bi4Ti3O12particles were arranged such that the faces of the plates were parallel to the sheet surface (Fig. 3(d)). As the next step, the sheets were cut into disks of 12 mm (Fig. 3(e)), stacked in 30e40 layers (Fig. 3(f)), and pressed at room tempera- ture under a pressure of 100 MPa for 3 min to form green compacts with a thickness of about 2 mm, as shown inFig. 3(g). At the end of this process, it reached textured microstructure as shown in Fig. 3(h).

The properties of the samples were studied using various analytical methods. In elaboration, X-ray diffraction (XRD) analysis (Rigaku RINT 2000) at room temperature was employed to explore the crystallinity, whilefield emission scanning electron microscopy (Nova NanoSEM 450-FEI-HUS-VNU) was used to examine the morphology [23]. Furthermore, the density and ferroelectric loops, used to identify ferroelectricity, of the samples, were measured by the Archimedes and Sawyer-Tower method, respectively [24].

Moreover, the grain size was determined by using a mean linear intercept method [3]. The measurement of dielectric properties was achieved by measuring the temperature dependencies of capaci- tance and phase angle (HIOKI 3532). Lastly, in other to study piezoelectric responses, the pellets were poled in a 120C silicone oil bath by applying a DC electricfield of 35 kV/cm for 20 min then cooling down to room temperature. The pellets were aged for 24 h prior to testing. The coupling factors (kp, kt) and piezoelectric constant (d31) were determined using a resonance method (HIOKI 3532, HP4193A) and calculated by the IEEE standard [24].

3. Results and discussion

Fig. 4 shows the variations in the density of (1-x)BNKT-xBT samples at different concentrations of BaTiO3. All the sintered samples exhibit high relative density more than 96%, indicating that dense microstructures are obtained. With the increase in the BaTiO3content up to 0.02 mol, the density of BNKT ceramics was increased. It achieved the maximum value (r¼5.86 g/cm3, 98.5% of the theoretical density) at a BaTiO3content of 0.02 mol and then decreased. This could be explained by the fact that when BaTiO3 content in the ceramic system was<0.02 mol, a large number of pores were present, indicating insufficient densification of the Fig. 1.The process of fabrication technology for random ceramics.

(3)

samples (Fig. 5). As the BaTiO3content was increased, the ceramics became denser, and the sample was almost completely dense at a BaTiO3content of 0.02 mol. In general, it has been observed that the addition of BaTiO3 enhances the density of BNKT ceramics. This result is consistent with that observed by Manotham et al. [25]. The effect of BaTiO3concentration on the dielectric constant (εr) and dielectric loss (tand) of the BNKT ceramics at 1 kHz is also illustrated inFig. 4. When the concentration of BaTiO3was increased from 0.0

to 0.02 mol, the values ofεrwere also increased proportionally and reached a maximum of 1354 at 0.02 mol of BaTiO3. A further in- crease inxresulted in a rapid decrease inεr. On the other hand, tand was decreased with increasing BaTiO3content. The minimum tand of 0.041 was obtained atx¼0.02, which increased further by the addition of BaTiO3. It could be explained on the basis of the large and homogeneous grain size [1,6] and dense composition of 0.98Bi0.5(Na0.8K0.2)0.5TiO3e0.02BaTiO3ceramic.

Fig. 5shows SEM micrographs and grain size distribution of the BNKT-BT ceramics at different BaTiO3contents. The grains in all the samples exhibited rectangular shapes with a clear grain boundary.

The grain size was increased with the increasing BaTiO3 content and reached to the maximum value of 1.28mm at 0.02 mol of BaTiO3

and then decreased. The smallest average grain size of 0.83mm was observed for the sample without BaTiO3(x¼0.0). This suggests that the microstructure can be improved by a BaTiO3additive. According to the research results of Khanal et al. [26] the conventionally prepared BaTiO3 ceramics sintered at 1350C for 5 h exhibited a grain size of 15.9mm. The large grains observed with increasing concentration of seed content may have resulted from the large particles (BT) diffusing into the BNKT ceramics, thus increasing the grain growth. This relationship can be used to empirically deter- mineGsusing Eq.(1):

Gs¼(1-x)(0.83mm)þx(15.9mm) (1)

Fig. 2.Characteristics of Bi4Ti3O12template: a) Microstructure; b) The distribution of particle size length; c) Crystal structure.

Fig. 3.The process of fabrication technology for textured ceramics.

Fig. 4.The values ofr,εr, and tandobserved for the BNKT-BT ceramics as a function of BaTiO3contents.

(4)

It is evident from the data that Eq. (1) gives a reasonable approximation of the average grain size with x¼0 to 0.02. How- ever, when the addition of BaTiO3is higher than 0.02 mol, there appears a large difference between the experimental data and theoretical calculations (Fig. 5). This can be explained by the fact that the diffusivity of solute atoms is normally different from that of host atoms. This result indicates that the addition of BaTiO3

exceeding the solubility limit in BNKT ceramics inhibits grain growth due to the solute drag effect.

Fig. 6(a) depicts the XRD patterns of BaTiO3-modified BNKT ceramics. All samples exhibited a pure perovskite phase, and no trace of secondary phase was detected in the investigated region. In order to illustrate the effect of BaTiO3on the structure of the ma- terials, a magnified view of the peaks in the region from 39to 41, and 46e47are shown inFig. 6(b) and(c), respectively. It can be seen that the pure BNKT sample possesses a rhombohedral sym- metry characterized by a pure (200) peak [12]. However, the large size Ba2þ(1.61 Å) ions diffuse into the BNKT lattice to replace Bi3þ (1.17 Å), Naþ(1.39 Å), and Kþ(1.64 Å) [12,27,28] resulting in the enlargement of lattice constant and a shift in the XRD peak position toward a lower diffraction angle direction. Though, a significant shift in the peak position of the reflection towards a higher angle side has been observed for the samples for samples containing

BaTiO3content higher than 0.03 mol. Similar kinds of behavior have been observed by V. Pal et al. [29] in the XRD patterns of the (1- x)(Bi0.96La0.04)0.5Na0.5TiO3)x(Ba0.90Ca0.10TiO3) ceramics, where the system changes its symmetry from rhombohedral R3c to tetragonal P4bm through an MPB between those two phases, similar to the study of Sowmya et al. [30] for Bi0.5(Na0.4K0.1)TiO3 ceramics. It proves that Ba diffuses into the BNKT lattice to change the structure of the material. Ghosh et al. [27] observed that the structural changes occurred gradually from a rhombohedral phase in BNKT ceramic to tetragonal in (1-x)BNKT-xBaTiO3ceramics at x¼0.05. In the study of Kamnoy et al. [11], (1-x)BNKT-xBaTiO3 ceramic with 0 and 2 mol% of BaTiO3 exhibited rhombohedral phase, which changed to tetragonal when the BaTiO3content of the samples exceeded 2 mol%. Recently, Manotham et al. [25] reported that Ba2þwas successfully diffused into the 0.92(Bi0.5Na0.42K0.08) TiO3-0.08(BaNb0.01Ti0.99)O3. The sample also showed a tetragonal perovskite (P4mm) structure in the ceramic. The phase diagram of the binary system (1-x)Bi0.5Na0.5TiO3exBaTiO3[31] showed mor- photropic phase boundary regions between the ferroelectric rhombohedral and tetragonal phases in thexrange of 0.06e0.08.

According to Kwei et al. [32] The long-range structure of BaTiO3is can be described by the P4mm space group with lattice parameters of a¼3.99095 Å and c¼4.0352 Å with a c/a ratio of 1.011. With Fig. 5.Typical SEM images of the BNKT-BT ceramics at different contents of BaTiO3.

Fig. 6.The XRD patterns of BaTiO3-modified BNKT ceramics with 2qranging from (a) 20e70, (b) 39e41and (c) 46e47.

(5)

increasing molar fraction of BaTiO3, the crystal symmetry of the (1- x)BNKT-xBaTiO3 ceramics should change due to the tetragonal distortions of BaTiO3. Therefore, in order to understand the degree of disorder and phase transformation behavior, the tolerance factor (t) [14,33] was calculated for these ceramics based on the ABO3 structure as shown inFig. 7(a).

tABO3¼ RAþRO ffiffiffi2

p ðRBþROÞ (2)

Therefore, in the pure Bi0.5(Na0.4K0.1)TiO3structure,

tBNKT¼0:5RBiþ ð0:4RNaþ0:1RKÞ þRO ffiffiffi2

p ðRTiþROÞ (3)

and in the pure BaTiO3(BT),

tBT¼ RBaþRO ffiffiffi2

p ðRTiþROÞ (4)

From Eqs.(3) and (4), the tolerance factors of BNKT (t¼0.9492), which is lower than the tolerance factor of BT (t¼1.0554), was calculated by considering the ionic radii of RBa¼1.61 Å, RBi3þ¼1.17 Å,RNaþ ¼1.39 Å,RKþ¼1.64 Å,RTi4þ¼0.61 Å andRO2¼1.42 Å [27,33]. It can be shown that the incorporation of BT into BNKT slightly increased the tolerance factor in the investigated region (Fig. 7(b)), corresponding to the decrease in lattice distortion and a trend toward slightly stable structure of tetragonal symmetry, similar to the study of Ghosh et al. [27] for Bi0.5(Na0.8K0.2)0.5TiO3

ceramics.

In order to provide convincing evidence for the phase structure evolution of BNKT-BT ceramics at different BaTiO3contents, Raman spectroscopy in the range of 80e1000 cm1 was carried out (Fig. 8(a)). It was applied to further characterize the surface groups and chemical bond states of materials [29,34e36]. As shown in Fig. 8(a), all of the bands were relatively broad, which is typical for relaxor based perovskite ferroelectrics [37]. The regions can be split into four categories according to the wavelength as above 700 cm1, between 400 and 700 cm1, between 200 and 400 cm1 and below 200 cm1. In the present study, the spectral deconvo- lution was performed according to six Gaussian-Lorentzian modes by using a best-fitting algorithm to illustrate the effect of compo- sition on the changes in peak characteristics [29,38].

In thefirst band region, the peak centered between 100 and 129 cm1was assigned to A1(TO) mode, which could be associated with the A site vibrations of the BieO, NaeO, KeO, and BaeO bonds [16,39] and is sensitive to phase changes in the crystal structure of

the perovskite [39]. It is interesting to note that the A-site mode exhibited a slight decrease in the wavenumber and intensity with the increasing BaTiO3contents (Fig. 8(a)). It can be explained by the fact that the BaTiO3 could dissolve into the BNKT lattice thus affecting the vibrations of the perovskite A-site and may induce a phase transition [40,41]. As mentioned above, the radius of Bi3þion (~1.61 Å) is similar to those of the Bi(~1.17 Å), Naþ(~1.39 Å) and Kþ(~1.64 Å) ions, suggesting that Ba2þions having a smaller ionic radius enter the A site of the perovskite unit cell to replace Bi3þ, Naþ and Kþions, which changes the lattice structure. Similar observa- tions have been reported by Manotham et al. [41] in other (1-x) Bi0.5(Na0.84K0.16)0.5TiO3-xBa(Nb0.01Ti0.99)O3ceramics.

The second region is located in the range of 200e400 cm1and is attributed to the vibration of the TieO bond. It is localized at around 266 cm1(x¼0) and associated with the tetragonal phase in the perovskite structure [37,42,43]. However, with the increasing BaTiO3 contents, the peak splits into two bands at ~254 and

~306 cm1. This proves that the (1-x)BNKT-xBaTiO3ceramics with 0e2 mol% of BaTiO3 exhibited the rhombohedral phase, which changed to tetragonal, as the BaTiO3content was increased further.

Similar kinds of behavior have been observed by Kreisel et al. [44]

in the Raman spectra of the Bi0.5(Na1xKx)0.5TiO3ceramics, where the system changes its symmetry from rhombohedral R3c to tetragonal P4bm through an MPB between those two phases.

Hayashi et al. [45] suggested that the bands around 305 cm1, which is characteristic of the tetragonal BaTiO3phase, is assigned to the B1mode.

Meanwhile, the two overlapping bands associated with the vi- brations of the TiO6octahedra within 450e650 cm1range become increasingly distinct with the increasingxcontent demonstrating the emergence of phonon behavior in the structure [46]. The modes of the TiO6octahedra localized at around 601 cm1exhibited an increase in the wavenumber with the increasing contents of BaTiO3

(Fig. 8(b)). It can be seen that the peaks shifted slightly, suggesting a distortion in the TiO6 octahedral and the presence of internal stresses that could possibly stabilize dielectric and ferroelectric properties at the room temperature after BaTiO3addition [41]. The broad band observed in the fourth region could be assigned to the A1(LO) and E(LO) overlapping bands, while the presence of broader peaks is related to the presence of oxygen vacancy [43].

The temperature dependence of dielectric characteristics of BaTiO3-modified BNKT ceramics measured at 1 kHz from room temperature to 330C is presented inFig. 9(a). The ceramics un- dergo two phase transitions in the measured temperature range.

Thefirst is the normal ferroelectric-to-ergodic relaxor (FE-to-ER) transition occurring at a low temperature (TF-R), and the second is

Fig. 7.a) The perovskite structure illustrated for ABO3; b) The tolerance factors of BNKT-BT ceramics.

(6)

the ferroelectriceparaelectric transition located at a higher tem- perature (Curie temperature, Tm). This exhibited relaxor charac- teristics by the phase transition occurring within a broad temperature range. This result is in a good agreement with the literature [30,41,47]. When the content of BaTiO3 changes from x¼0e0.02 mol, the maximum value of the dielectric constant (εmax) increases from 3296 to 4037 (Fig. 9(b)). On the other hand, εmaxdecreases with the increasing BaTiO3concentration beyond 0.02 mol concomitant with the increase in the grain size as shown inFig. 5.

The diffuseness (g) was evaluated by plotting ln(1/ε - 1/εmax) versus ln(T - Tm) [14] at 1 kHz and temperatures greater thanTm, as demonstrated inFig. 9(c). The increase ingin this range is higher than the increase in the case when the content of BaTiO3was from 0.0 (g¼1.60) to 0.02 mol (g¼1.87) (Fig. 9(b)). Then, it rapidly de- creases with increasingx. Considering from a structural viewpoint, we can explain that the BaTiO3get dissolved into the BNKT lattice, which affects the vibrations of the perovskite A-site and may induce a phase transition due to the difference in radius ions as discussed above. In other words, when the BaTiO3addition is above the limit, the excessive Ba ions would segregate at the grain boundary, inhibit the grain growth, and lead to the heterogeneity of the structure causing g to decrease. These results are in good agreement with a study of Kamnoy et al. [11] on lead-free BNKT-

based ceramics. Thus, it can be concluded that the addition of BaTiO3into BNKT ceramics increases the value ofg, which means that the paraelectric-ferroelectric phase transition is of diffuse type, and the ceramics exhibit high disorder [14]. This is also the reason for the decrease ofTF-Rfrom 120C to 85C and the decrease ofTm

from 269C to 231C when increasing BaTiO3content is gradual (Fig. 9(d)). These results are consistent with the literature [27].

According to Ghosh et al. [27], the decrease inTF-Ris related to non- cubic distortion with the incorporation of BaTiO3 into BNKT ce- ramics, consequently destabilizing the long range ordering state.

According to Lu et al. [48], the temperatureTF-Rcorresponds to the ferroelectricerelaxor transition, TF-R should be lower than the temperature range of energy storage application to obtain a large energy density. Recent investigations by Liu et al. [49] have shown that the Raman modes vary nonmonotonically with distinctive abrupt changes aroundTF-R, suggesting a thermal transition from thefield-forced ferroelectric regime to disordered ergodic relaxor phase. As can be observed inFig. 7(c), the incorporation of BaTiO3

into BNKT decreasesTmin the investigated region. There is a dif- ference between the phase transformation temperatures of BNKT (Tm~ 269C, at x¼0) and BaTiO3 (Tm~ 120C) [33], thus it is important to determine the dependence of phase transition tem- perature of the (1-x)Bi0.5(Na0.8K0.2)0.5TiO3 exBaTiO3ceramics on BaTiO3 content (Fig. 9(d)). There is a good linear relationship Fig. 8.Raman spectra of BNKT-BT ceramics.

Fig. 9.Temperature dependence of dielectric characteristics of BNKT-BT ceramics.

(7)

betweenTmandx, indicating that this ceramic is a well behaved complete solid solution. This relationship can be used to empiri- cally determineTmusing Eq.(5)[50]:

Tm¼(1-x)(269C)þx(120C) (5)

The variation in the calculatedTmas a function of compositionx is shown inFig. 9(d). It is evident from the data that Eq.(5)gives a reasonable approximation of the transition temperatureTmwith x¼0.0 to 0.02. This result suggests that Tm of (1-x) Bi0.5(Na0.8K0.2)0.5TiO3exBaTiO3ceramics can be varied over a wide range from 120C to 269C by controlling the content of BaTiO3in the composition. However, when the addition of BaTiO3is higher than 0.02 mol, there appears a large difference between the experimental data and theoretical calculations. This can be explained by the solubility limit of Ba2þions in Bi0.5(Na0.8K0.2)0.5-

TiO3ceramics near 0.02 mol. However, if the addition of BaTiO3is above 0.02 mol, excess Baions accumulate at the grain bound- aries, which has little influence on the average energy of the oxygen octahedron; thus, the temperatureTmwill be abnormally influ- enced (not according to rules) by the excess Baions.

Fig. 10shows the changes in electromechanical coupling factors in planar mode (kp), thickness mode (kt), the piezoelectric constant (d31), and mechanical quality factorQmas a function of the BaTiO3

content in BNKT ceramics sintered at 1100C for 2 h. When the content of BaTiO3is0.02 mol, the values ofkp,kt,d31, andQm

increase rapidly with an increase in the BaTiO3 content, then decrease. The optimum values forkpof 0.30, ktof 0.31,d31of 99 pC/

N andQmof 189 were obtained atx¼0.02, which is near the so- lution limit. In other words, the improvement in the piezoelectric properties of the samples can be partially ascribed to the enhancement in the density and the increasing grain size effect.

Similar observations were made for Ba-doped Bi0.5(Na0.85K0.15)0.5-

TiO3 ceramics [51]. It can be found in Fig. 4that all the relative densities are larger than 95%, indicating that the effect of pores on the piezoelectric properties is negligible [52]. Therefore, the improvement of piezoelectric properties of ceramics can only be explained by the particle size increase effect and density.

Qaiser et al. [53] show that the dense structure not only im- proves the piezoelectric response but also increases theQmvalue by reducing the dissipated energy. As a result ofFig. 5presents the ceramic samples have surfaces with very few pores and exhibit good density with sufficient grain growth. In which, the 0.98Bi0.5(Na0.8K0.2)0.5TiO3 - 0.02BaTiO3 samples have the best modification of microstructure with great density. It is believed to be associated with the larger grain comparing to other samples, where far fewer grain boundaries will make the ferroelastic domain wall reversal easier [18]. On the other hand, according to Hayati et al. [54] the improvement of the dielectric and piezoelectric properties are assumed to be related to the formation of a

homogenous microstructure with optimum grain size in piezoceramics.

TheP-Ehysteresis loops of the BaTiO3doped BNKT sintered at 1100C were measured at room temperature and are shown in Fig. 11(aee); the remanent polarization (Pr), coercivefield (Ec) and the squareness of the hysteresis loop (Rsq) were also determined and are shown inFig. 11(g). Based on the increase ofPrand decrease of Ec, there is improvement of ferroelectric properties of BaTiO3 doped BNKT ceramics by doping BaTiO3up to 0.02 mol. However, when the BaTiO3 content of the samples exceeded 2 mol%, the BNKT-BT ceramics have a small grain size, which results in large grain surface area, and internal stress which will reduce the orientation of polarization along the electric field, and strongly influence ferroelectric properties of ceramics [25]. On the other hand, the 0.98Bi0.5(Na0.4K0.1)TiO3 - 0.02BaTiO3 ceramics shows good ferroelectric properties, with high remanent polarization of Pr¼17.3mC/cm2 and rather low coercive electric field of Ec¼27.2 kV/cm. Cao et al. [55] derived an empirical equation to measure not only the deviation in the polarization axis but also that in the electricfield axis like this:

Rsq¼Pr

Psþ P1:1EC

Pr (6)

whereRsqis the squareness of the hysteresis loop,Pris the remnant polarization,Psis the saturation polarization, andP1:1Ecis the po- larization at an electric field equal to 1.1 times of coercivefield.

When the BaTiO3content varied from 0 to 0.04 mol, the values of Rsq changed to 0.96, 0.99, 1.04, 0.89 and 0.73, respectively (Fig. 11(g)). Similar observations were made for Ba(Nb0.01Ti0.99)O3- doped Bi0.5(Na0.84K0.16)0.5TiO3 ceramics [41]. Thus, the samples showed relaxor dielectric relations and this observation is consis- tent with the dielectric analysis (Fig. (9)).

Fig. 11(f) showsPeEhysteresis loops at room temperature for textured 0.98BNKT-0.02BT ceramics with 10 wt% of Bi4Ti3O12tem- plates using the process of fabrication technology for textured ce- ramics (Fig. 3). The Bi4Ti3O12templates were well-aligned parallel to the casting direction and seeded the formation of the BNKT phase around the oriented template particles (Fig. 12). Almost a single layer of BNKT particles was formed on the (100) surface of Bi4Ti3O12 templates, similar to the study of Chang et al. for Sr0.61Ba0.39Nb2O6ceramics [56]. As can be seen that the textured sample exhibited low remnant polarization (Pr¼15.2mC/cm2) and coercivefield (Ec¼22.6 kV/cm) compared with the random sample (Fig. 11(c),Pr¼17.3mC/cm2,Ec¼27.2 kV/cm) due to the crystal-like characteristics of the textured ceramic). It is known that the coer- cive field of a single crystal is usually lower than that of poly- crystalline ceramic with the same composition [19]. Textured BNKT-BT ceramic exhibitedRsqof 1.01, which is lower than that of the random samples (Rsq¼1.04).

The energy storage densityW1(Fig. 11(f), marked in blue area) was obtained by integrating the area between the polarization axis and the discharge curve of the unipolarPeEhysteresis loops using Eq.(7)[48,57]:

W1¼ ð

Pmax

Pr

EdP (7)

The energy loss density (W2) (Fig. 11(f), marked in the green area) caused by the domain reorientation was obtained by inte- grating the area between the charge and the discharge curve. The energy storage efficiency (h) of the material can be calculated by Eq.

(7)[48,57]:

Fig. 10.kp,kt,d31, andQmas a function of BaTiO3content (x) in BNKT ceramic.

(8)

h

¼ W1

W1þ W2 (8)

The varying trend in the energy storage density (W1) of the ceramics at different BaTiO3contents, is similar to the trend of the energy storage efficiency (h), as shown inFig. 11(h). BothW1andh increase almost linearly with the increment of BaTiO3content at x<0.03 and reaches a maximum value of 0.30 J/cm3 and 24.6%, respectively at x¼0.04. While, as the BaTiO3 content increased from 0 to 0.04 mol, the values ofW2of samples increased from 1.0 to 1.16 J/cm3 reaching the highest value of 1.16 J/cm3 atx¼0.02, upon which it then decreased. At this content, the textured 0.98BNKT-0.02BT ceramic exhibited W1 of 0.14 J/cm3 and W2 of 0.88 J/cm3, which are lower than that of the random samples (W1¼0.15 J/cm3, W2¼1.16 J/cm3) at 42.3 kV/cm as illustrated in Fig. 11(h). These values are comparable to the previously reported of BNKT-based bulk ceramics studied by Manotham et al. [41]

which showed an energy density value of ~0.09 J/cm3and an en- ergy storage efficiency value of ~9% (measured at 25C). However, in the same composition (0.98BNKT-0.02BT), the energy storage efficiency (h) of textured BNKT is higher than that of the random ceramics, i.e., 13.5% and 11.5%, respectively (Table 1). Although the random 0.98BNKT-0.02BT ceramic possesses higherPmvalues, it also bears largerPr,ECprobably because of the larger FE hysteresis and higher leakage current at high electricfields. As a result, their energy densities are substantially limited, as also observed by Pan et al. [58]. In other words, the largeRsqalso causes a remarkable decline inhof random ceramics.

The ferroelectric and piezoelectric properties of the textured and random ceramics are compared inTable 1. As can be noticed that the textured ceramic shows a lower remnant polarization, while the piezoelectric coefficient (d31) and electromechanical coupling coefficients (kp and kt) of the textured ceramics were found to increase by about 8%, 13.3%, and 19.4%, respectively. This finding is in a good agreement with the degree of orientation of the ceramic samples (Fig. 12). The derivations of the plate-like square- shaped particles with a greater length of 3mm and thickness of 1.5mm from the template growth at the expense of the matrix powders after sintering can be seen inFig. 12. This affirms the effectiveness of the template grain growth method in aligning the template particles. In short, the BaTiO3-modified BNKT ceramics show improved electrical properties, which are maximum at a BaTiO3 content of 0.02 mol. Besides, the electrical properties of BNKT - BT ceramics can be improved by the process of fabrication technology for textured ceramics (Fig. 3). As shown the inset in Fig. 12, the non-split (100) peak at the 2qof around 22and the (200) peak at the 2qof around 46proved that textured 0.98BNKT- 0.02BT ceramic has a rhombohedral structure [14]. The intensity of the (110) reflections was found to decrease with an increase in the intensity of the (100) and (200) reflections at 2qangles of 22.76 and 46.61, respectively, which indicates the development of texture in the samples [59].

To confirm the achievements of the present study, the charac- teristic parameters of the BNKT-BT ceramics sintered at 1100C as such: the sintering temperature (Ts), Curie temperature, (Tm), electromechanical coupling coefficient (kp), the piezoelectric coef- ficient (d31), the coercivefield (Ec), and the remnant polarization Fig. 11.(aef)PeEhysteresis loops and (g, h) the calculated the ferroelectric parameters of ceramics.

Fig. 12.Microstructure of Textured 0.98BNKT-0.02BT ceramics.

(9)

(Pr) were extracted and compared with those of other Bi-based leadefree ceramics [16e19,60,61] as listed inTable 2. Our results indicated that the BaTiO3-modified BNKT ceramics shows a lower sintering temperature, while the electrical properties of the ma- terial ceramics are well maintained.

4. Conclusions

In this work, the BaTiO3 doped Bi0.5(Na0.4K0.1)TiO3 ceramics were successfully synthesized by a solid-state mixed oxide method.

The results indicate that BaTiO3enhances the dielectric, ferroelec- tric, and piezoelectric properties of BNKT-based ceramics. The best physical properties of the BNKT-BT ceramics, such as density, r¼5.86 g/cm3, 98.5% of the theoretical density); electromechanical coupling factors (kp), 0.30; (kt), 0.31; remanent polarization (Pr), 17.2mC cm2; dielectric constant (εr), 1354; and highest dielectric constant (εmax), 4037, were obtained at 0.02 mol of BaTiO3. The dielectric curve exhibited broad transition peaks aroundTF-Rand Tm, which showed the characteristics of a diffuse phase transition.

In addition, the ferroelectric, piezoelectric properties and the energy storage efficiency in textured 0.98Bi0.5(Na0.4K0.1)TiO3 e 0.02BaTiO3ceramics using the process of fabrication technology for textured ceramics were studied and compared with the random ceramics.

In general, two methods are considered very effective in this study: development of (1-x)Bi0.5(Na0.4K0.1)TiO3exBaTiO3ceramics through compositional modification of BaTiO3 and fabrication of textured ceramics having a uniform grain orientation. Thefirst, the enhancement of the piezoelectric and dielectric properties of the BNKT-BT ceramics by the process of fabrication technology for textured ceramics. This leads to achieving significant enhance- ments in the piezoelectric response of ceramics suitable for appli- cations to fabricate an ultrasonic sensor, ultrasonic cleaners, and hydroacoustic equipment. The second, the improved ferroelectric properties may be contributed by the increase in the BaTiO3addi- tion in BNKT ceramics. This leads to achieving significant en- hancements in the energy storage properties of the material for

dielectric capacitor applications. Note that high efficiency is also crucial for dielectrics because it means less waste heat, better reliability and longer lifetime of capacitors in practical applications [58].

Acknowledgement

This research is funded by Vietnam National Foundation for Science and Technology Development (NAFOSTED) under grant number 103.02-2017.308.

References

[1] L.D. Vuong, P.D. Gio, Structure and electrical properties of Fe2O3-doped PZT- PZN-PMnN ceramics, J. Mod. Phys. 5 (14) (2014) 1258e1263.

[2] L.D. Vuong, P.D. Gio, N.T. Tho, T.V. Chuong, Relaxor ferroelectric properties of PZT-PZN-PMnN Ceramics, Indian J. Eng. Mater. Sci. 20 (2013) 555e560.

[3] L.D. Vuong, P.D. Gio, N.D.V. Quang, T. Dai Hieu, T.P. Nam, Development of 0.8 Pb (Zr 0.48 Ti 0.52) O 3e0.2 Pb [(Zn 1/3 Nb 2/3) 0.625 (Mn 1/3 Nb 2/3) 0.375]

O 3 ceramics for high-intensity ultrasound applications, J. Electron. Mater. 47 (10) (2018) 5944e5951.

[4] L.D. Vuong, N.D.T. Luan, Temperature dependence of ferroelectric properties of Li2CO3 doped PZTePZNePMnN ceramics, J. Mechatron. 3 (3) (2015) 193e196.

[5] L.D. Vuong, P.D. Gio, T.V. Chuong, D.T.H. Trang, D.V. Hung, N.T. Duong, Effect of Zr/Ti ratio content on some physical properties of low temperature sintering PZTPZNPMnN ceramics, Int. J. Mater. Chem. 3 (2) (2013) 39e43.

[6] N.D.T. Luan, L. Vuong, B. Chanh, Microstructure, ferroelectric and piezoelectric properties of PZTePMnSbN ceramics, Int. J. Mater. Chem. 3 (3) (2013) 51e58.

[7] N. Dinh Tung Luan, L.D. Vuong, T. Van Chuong, N. Truong Tho, Structure and physical properties of PZT-PMnN-PSN ceramics near the morphological phase boundary, Adv. Mater. Sci. Eng. 2014 (2014). Article ID 821404, 8 pages.

[8] P.D. Gio, H.Q. Viet, L.D. Vuong, Low-temperature sintering of 0.96 (K0. 5Na0. 5) NbO3-0.04 LiNbO3 lead-free piezoelectric ceramics modified with CuO, Int. J.

Mater. Res. 109 (11) (2018) 1071e1076.

[9] D.A. Tuan, V.T. Tung, L.D. Vuong, N.H. Yen, L.T.U. Tu, Investigation of phase formation and poling conditions of lead-free 0.48 Ba (Zr 0.2 Ti 0.8) O 3e0.52 (Ba 0.7 Ca 0.3) TiO 3 ceramic, J. Electron. Mater. 47 (10) (2018) 6297e6301.

[10] D.A. Tuan, L.D. Vuong, V.T. Tung, N.N. Tuan, N.T. Duong, Dielectric and ferro- electric characteristics of doped BZT-BCT ceramics sintered at low tempera- ture, J. Ceram. Process. Res. 19 (1) (2018) 32e36.

[11] M. Kamnoy, K. Sutjarittangtham, P. Parjansri, U. Intatha, S. Eitssayeam, N. Raengthon, Effects of BaTiO3 doped on structural and dielectric properties of Bi0. 5 (Na0. 80K0. 20) 0.5 TiO3 ceramic, Ferroelectrics 511 (1) (2017) Table 1

The ferroelectric and piezoelectric properties of the textured and random ceramics.

Samples Textured/Random kp kt d31

pC/N Pr

mC/cm2 Ps

mC/cm2

EC(kV/cm) W1

J/cm3 W2

J/cm3

h(%) Rsq

BNKT Random 0.27 0.19 63 9.1 12.5 29.1 0.08 1.00 7.4 0.96

0.99BNKT-0.01BT Random 0.29 0.21 88 12.4 16.46 28.16 0.11 1.1 9.1 0.99

0.98BNKT-0.02BT Random 0.3 0.31 99 17.3 22.7 27.2 0.15 1.16 11.5 1.04

0.97BNKT-0.03BT Random 0.24 0.28 78 16.1 23.64 18.8 0.19 1.05 15.3 0.89

0.96BNKT-0.04BT Random 0.21 0.22 74 12.7 22.4 13.9 0.3 0.92 24.6 0.73

0.98BNKT-0.02BT Textured 0.34 0.37 107 15.2 22.3 22.6 0.14 0.88 13.5 1.01

BNKT [41] Random e e e 29 42 16e42 0.09 e 9 1.25

Table 2

The comparison of the characteristic parameters of the BNKT-BT ceramics and other ceramics.

Ceramics Textured or Random TS(ºC) Tm(ºC) kp d31

pC/N

EC(kV/cm) Pr(mC/cm2) Ref.

Bi0.5(Na0.82K0.18)0.5TiO3 Random 1100 300 e e 10.5e32.4 9e30 [16]

(1x)Bi1/2Na1/2TiO3exSrTiO3 Random 1175 200 e 20e135 5e22 2e12 [17]

0.88Na0.5Bi0.5TiO3-0.08K0.5Bi0.5TiO3-0.04BaTiO3 Random 1165 290 e 150 35 35.2 [18]

0.88Na0.5Bi0.5TiO3-0.08K0.5Bi0.5TiO3-0.04BaTiO3 Textured 1165 270 e 185 31 26.4 [18]

Na0.5Bi0.5TiO3eBaTiO3 Random 1200 280 0.24 128 30 39.6 [19]

Na0.5Bi0.5TiO3eBaTiO3 Textured 1200 294 0.34 299 24.2 34.1 [19]

(1-x)Bi0.5(Na0.84K0.16)0.5TiO3-xBa(Nb0.01Ti0.99)O3 Random 1125 300 e e 16e42 20e29 [41]

(Ba0.7Ca0.3)Ti1-xCuxO3-x Random 1350 110 0.24 120 5e20 6e10 [60]

Bi3.25La0.75Ti2.97V0.03O12 Random 1350 110 0.22 17 46 11 [61]

Bi3.25La0.75Ti2.97V0.03O12 Textured 1100 110 0.28 33 46 26.2 [61]

Bi0.5(Na0.4K0.1)TiO3eBaTiO3 Random 1100 262 0.30 99 27.2 17.3 This work Bi0.5(Na0.4K0.1)TiO3eBaTiO3 Textured 1100 e 0.34 107 22.6 15.2 This work

(10)

94e103.

[12] L.D. Vuong, N.T. Tho, The sintering behavior and physical properties of Li2CO3-doped Bi0.5(Na0.8K0.2)0.5TiO3 lead-free ceramics, Int. J. Mater. Res.

108 (3) (2017) 222e227.

[13] P.D. Gio, N.V.D. Hong, L.D. Vuong, Effect of excess Bi2O3 content on the structure and dielectric, piezoelectric properties of Bi0. 5 (Na0. 8 K0. 2) 0.5 TiO3 lead free ceramics, Adv. Porous Mater. 3 (1) (2015) 29e32.

[14] L.D. Vuong, N. Truong-Tho, Effect of ZnO nanoparticles on the sintering behavior and physical properties of Bi (NaK) TiO lead-free ceramics, J. Electron. Mater. 46 (11) (2017) 6395e6402.

[15] P. Fan, Y. Zhang, S.-T. Zhang, B. Xie, Y. Zhu, M.A. Marwat, W. Ma, K. Liu, L. Shu, H. Zhang, Low-temperature sintered (Na1/2Bi1/2)TiO3-based incipient pie- zoceramics for co-fired multilayer actuator application, J. Materiomics 5 (3) (2019) 480e488.

[16] P. Fan, Y. Zhang, Q. Zhang, B. Xie, Y. Zhu, M.A. Mawat, W. Ma, K. Liu, J. Xiao, H. Zhang, Large strain with low hysteresis in Bi4Ti3O12 modified Bi1/2 (Na0.

82K0. 18) 1/2TiO3 lead-free piezoceramics, J. Eur. Ceram. Soc. 38 (13) (2018) 4404e4413.

[17] T.A. Duong, H.-S. Han, Y.-H. Hong, Y.-S. Park, H.T.K. Nguyen, T.H. Dinh, J.-S. Lee, Dielectric and piezoelectric properties of Bi 1/2 Na 1/2 TiO 3eSrTiO 3 leadefree ceramics, J. Electroceram. 41 (1e4) (2018) 73e79.

[18] N. Dong, X. Gao, F. Xia, H. Liu, H. Hao, S. Zhang, Dielectric and piezoelectric properties of textured lead-free Na0. 5Bi0. 5TiO3-Based ceramics, Crystals 9 (4) (2019) 206.

[19] W. Zhao, J. Ya, Y. Xin, L.E., D. Zhao, H. Zhou, Fabrication of Na0. 5Bi0.

5TiO3eBaTiO3-textured ceramics templated by plate-like Na0. 5Bi0. 5TiO3 particles, J. Am. Ceram. Soc. 92 (7) (2009) 1607e1609.

[20] F. Fu, J. Zhai, Z. Xu, B. Shen, X. Yao, Grain growth kinetics of textured-BaTiO 3 ceramics, Bull. Mater. Sci. 37 (4) (2014) 779e787.

[21] P. Li, J. Zhai, B. Shen, S. Zhang, X. Li, F. Zhu, X. Zhang, Ultrahigh piezoelectric properties in textured (K, Na) NbO3-based lead-free ceramics, Adv. Mater. 30 (8) (2018) 1705171.

[22] Z. Zhao, X. Li, H. Ji, M.J.I.F. Deng, Formation Mechanism of Plate-like Bi4Ti3O12 Particles in Molten Salt Fluxes, vol. 154, 2014, pp. 154e158, 1.

[23] F. Lotgering, Topotactical reactions with ferrimagnetic oxides having hexag- onal crystal structuresdI, J. Inorg. Nucl. Chem. 9 (2) (1959) 113e123.

[24] C.B. Sawyer, C.H. Tower, Rochelle salt as a dielectric, Phys. Rev. 35 (3) (1930) 269e273.

[25] S. Manotham, P. Butnoi, P. Jaita, T. Tunkasiri, Structure and electrical prop- erties of BNKT-based ceramics, Solid State Phenom. 283 (2018) 147e153.

[26] G.P. Khanal, S. Kim, M. Kim, I. Fujii, S. Ueno, S. Wada, Grain-size dependence of piezoelectric properties in thermally annealed BaTiO3 ceramics, J. Ceram. Soc.

Jpn. 126 (7) (2018) 536e541.

[27] S.K. Ghosh, V. Chauhan, A. Hussain, S.K. Rout, Phase transition and energy storage properties of BaTiO3-modified Bi0.5(Na0.8K0.2)0.5TiO3 ceramics, Ferroelectrics 517 (1) (2017) 97e103.

[28] S.K. Rastogi, P. Divya, B. Praveenkumar, A. Kumar, Effect of Sr2þ, Ba 2þand Ta5þions on structural and electrical properties of BNKT ceramics, Mater.

Today: Proc. 2 (4e5) (2015) 2784e2788.

[29] V. Pal, O.P. Thakur, R.K. Dwivedi, Investigation of MPB region in lead free BLNT-BCT system through XRD and Raman spectroscopy, J. Phys. D Appl. Phys.

48 (5) (2015), 055301.

[30] N.S. Sowmya, P. Varade, N. Venkataramani, A.R. Kulkarni, Enhanced ferro- electric and converse piezoelectric properties of dense lead-free Na0. 4K0.

1Bi0. 5TiO3 ceramics for actuator applications, Adv. Mater. Phys. Chem. 9 (1) (2019) 1e10.

[31] K. Reichmann, A. Feteira, M. Li, Bismuth sodium titanate based materials for piezoelectric actuators, Materials 8 (12) (2015) 8467e8495.

[32] G.H. Kwei, A.C. Lawson, S.J.L. Billinge, S.W. Cheong, Structures of the ferro- electric phases of barium titanate, J. Phys. Chem. 97 (10) (1993) 2368e2377.

[33] Y. Tsur, T.D. Dunbar, C.A. Randall, Crystal and defect chemistry of rare earth cations in BaTiO3, J. Electroceram. 7 (1) (2001) 25e34.

[34] A. Feng, M. Ma, Z. Jia, M. Zhang, G. Wu, Fabrication of NiFe 2 O 4@ carbonfiber coated with phytic acid-doped polyaniline composite and its application as an electromagnetic wave absorber, RSC Adv. 9 (44) (2019) 25932e25941.

[35] Z. Jia, Z. Gao, A. Feng, Y. Zhang, C. Zhang, G. Nie, K. Wang, G. Wu, Laminated microwave absorbers of A-site cation deficiency perovskite La0. 8FeO3 doped at hybrid RGO carbon, Composites Part B: Engineering 176 (2019) 107246.

[36] X. Zhou, Z. Jia, A. Feng, X. Wang, J. Liu, M. Zhang, H. Cao, G. Wu, Synthesis of fish skin-derived 3D carbon foams with broadened bandwidth and excellent electromagnetic wave absorption performance, Carbon 152 (2019) 827e836.

[37] S. Bhandari, N. Sinha, G. Ray, B. Kumar, Flux growth of lead free (Na 0.5 Bi 0.5) TiO 3e(K 0.5 Bi 0.5) TiO 3 ferroelectric single crystals and their character- ization, CrystEngComm 16 (21) (2014) 4459e4466.

[38] T. Wang, X.-m. Chen, Y.-z. Qiu, H.-l. Lian, W.-t. Chen, Microstructure and electrical properties of (1x)[0.8Bi0.5Na0.5TiO3-0.2Bi0.5K0.5TiO3]-xBiCoO3 lead-free ceramics, Mater. Chem. Phys. 186 (2017) 407e414.

[39] D.A. Fernandez-Benavides, A.I. Gutierrez-Perez, A.M. Benitez-Castro, M.T. Ayala-Ayala, B. Moreno-Murguia, J. Mu~noz-Salda~na, Comparative study of ferroelectric and piezoelectric properties of BNT-BKT-BT ceramics near the

phase transition zone, Materials 11 (3) (2018) 361.

[40] P. Fan, Y. Zhang, Q. Zhang, B. Xie, Y. Zhu, M.A. Mawat, W. Ma, K. Liu, J. Xiao, H. Zhang, Large strain with low hysteresis in Bi4Ti3O12 modified Bi1/

2(Na0.82K0.18)1/2TiO3 lead-free piezoceramics, J. Eur. Ceram. Soc. 38 (13) (2018) 4404e4413.

[41] S. Manotham, P. Butnoi, P. Jaita, N. Kumar, K. Chokethawai, G. Rujijanagul, D.P. Cann, Large electricfield-induced strain and large improvement in energy density of bismuth sodium potassium titanate-based piezoelectric ceramics, J. Alloy. Comp. 739 (2018) 457e467.

[42] T.H. Kim, S. Kojima, C.W. Ahn, I.W. Kim, J.-H. Ko, Raman-and Brillouin- scattering studies on lead-free piezoelectric Bi 0.5 (Na 0.78 K 0.22) 0.5x TiO 3 ceramics with A-site vacancies, J. Korean Phys. Soc. 62 (7) (2013) 1009e1013.

[43] G. Hernandez-Cuevas, J.R. Leyva Mendoza, P.E. García-Casillas, C.A. Rodríguez Gonzalez, J.F. Hernandez-Paz, G. Herrera-Perez, L. Fuentes-Cobas, S. Díaz de la Torre, O. Raymond-Herrera, H. Camacho-Montes, Effect of the sintering technique on the ferroelectric and d33 piezoelectric coefficients of Bi0.5(Na0.84K0.16)0.5TiO3 ceramic, J. Adv. Ceram. 8 (2) (2019) 278e288.

[44] J. Kreisel, A.M. Glazer, G. Jones, P.A. Thomas, L. Abello, G. Lucazeau, An x-ray diffraction and Raman spectroscopy investigation of A-site substituted perovskite compounds: the (Na1-xKx) 0.5 Bi0. 5TiO3 (0 x1) solid solution, J. Phys. Condens. Matter 12 (14) (2000) 3267.

[45] H. Hayashi, T. Nakamura, T. Ebina, In-situ Raman spectroscopy of BaTiO3 particles for tetragonalecubic transformation, J. Phys. Chem. Solids 74 (7) (2013) 957e962.

[46] J. Hao, Z. Xu, R. Chu, W. Li, P. Fu, J. Du, Field-induced large strain in lead-free (Bi0.5Na0.5)1xBaxTi0.98 (Fe0.5Ta0.5)0.02O3 piezoelectric ceramics, J. Alloy.

Comp. 677 (2016) 96e104.

[47] L.K. Pradhan, R. Pandey, S. Kumar, S. Kumari, M. Kar, Evidence of composi- tional fluctuation induced relaxor antiferroelectric to antiferroelectric ordering in Bi 0.5 Na 0.5 TiO 3eBi 0.5 K 0.5 TiO 3 based lead free ferroelectric, J. Mater. Sci. Mater. Electron. 30 (10) (2019) 9547e9557.

[48] X. Lu, J. Xu, L. Yang, C. Zhou, Y. Zhao, C. Yuan, Q. Li, G. Chen, H. Wang, Energy storage properties of (Bi0.5Na0.5)0.93Ba0.07TiO3 lead-free ceramics modified by La and Zr co-doping, J. Materiomics 2 (1) (2016) 87e93.

[49] X. Liu, J. Zhai, B. Shen, Local phenomena in bismuth sodium titanate perov- skite studied by Raman spectroscopy, J. Am. Ceram. Soc. 101 (12) (2018) 5604e5614.

[50] N. Vittayakorn, G. Rujijanagul, X. Tan, M.A. Marquardt, D.P. Cann, The mor- photropic phase boundary and dielectric properties of the x Pb (Zr 1∕2 Ti 1∕

2) O 3-(1x) Pb (Ni 1∕3 Nb 2∕3) O 3 perovskite solid solution, J. Appl. Phys.

96 (9) (2004) 5103e5109.

[51] S.K. Rastogi, P. Divya, B. Praveenkumar, A. Kumar, Effect of Sr2þ, Ba 2þand Ta5þions on structural and electrical properties of BNKT ceramics, Mater.

Today: Proc. 2 (4) (2015) 2784e2788.

[52] Y. Huan, X. Wang, J. Fang, L. Li, Grain size effects on piezoelectric properties and domain structure of BaTiO 3 ceramics prepared by two-step sintering, J. Am. Ceram. Soc. 96 (11) (2013) 3369e3371.

[53] M.A. Qaiser, A. Hussain, Y. Xu, Y. Wang, Y. Wang, Y. Yang, G. Yuan, CuO added Pb0. 92Sr0. 06Ba0. 02 (Mg1/3Nb2/3) 0.25 (Ti0. 53Zr0. 47) 0.75 O3 ceramics sintered with Ag electrodes at 900C for multilayer piezoelectric actuator, Chin. Phys. B 26 (3) (2017), 037702.

[54] R. Hayati, A. Barzegar, Microstructure and electrical properties of lead free potassium sodium niobate piezoceramics with nano ZnO additive, Mater. Sci.

Eng., B 172 (2) (2010) 121e126.

[55] R. Cao, G. Li, J. Zeng, S. Zhao, L. Zheng, Q. Yin, The piezoelectric and dielectric properties of 0.3 Pb (Ni1/3Nb2/3) O3exPbTiO3e(0.7x) PbZrO3 ferroelectric ceramics near the morphotropic phase boundary, J. Am. Ceram. Soc. 93 (3) (2010) 737e741.

[56] Y. Chang, S. Lee, S. Poterala, C.A. Randall, G.L. Messing, A critical evaluation of reactive templated grain growth (RTGG) mechanisms in highly [001] textured Sr 0.61 Ba 0.39 Nb 2 O 6 ferroelectric-thermoelectrics, J. Mater. Res. 26 (24) (2011) 3044e3050.

[57] X. Liu, J. Shi, F. Zhu, H. Du, T. Li, X. Liu, H. Lu, Ultrahigh energy density and improved discharged efficiency in bismuth sodium titanate based relaxor ferroelectrics with A-site vacancy, J. Materiomics 4 (3) (2018) 202e207.

[58] H. Pan, J. Ma, J. Ma, Q. Zhang, X. Liu, B. Guan, L. Gu, X. Zhang, Y.-J. Zhang, L. Li, Y. Shen, Y.-H. Lin, C.-W. Nan, Giant energy density and high efficiency ach- ieved in bismuth ferrite-basedfilm capacitors via domain engineering, Nat.

Commun. 9 (1) (2018) 1813.

[59] K. Fuse, T. Kimura, Effect of particle sizes of starting materials on micro- structure development in textured Bi0. 5 (Na0. 5K0. 5) 0.5 TiO3, J. Am. Ceram.

Soc. 89 (6) (2006) 1957e1964.

[60] P. Jaiban, A. Watcharapasorn, R. Yimnirun, R. Guo, A.S. Bhalla, Dielectric, ferroelectric and piezoelectric properties of (Ba0. 7Ca0. 3) Ti1-xCuxO3-x ce- ramics, J. Alloy. Comp. 759 (2018) 120e127.

[61] C.W. Ahn, E.D. Jeong, Y.H. Kim, J.S. Lee, G.S. Chung, J.Y. Lee, I.W. Kim, Piezo- electric properties of textured Bi 3.25 La 0.75 Ti 2.97 V 0.03 O 12 ceramics fabricated by reactive templated grain growth method, J. Electroceram. 23 (2e4) (2009) 392e396.

Referensi

Dokumen terkait

Result and Discussion Based on the result of the data research analysis, it’s found that the level implementation of the entrepreneurship values by the teacher was interpreted very

IERC Research BRIEF Evidence for Scale Independent Evaluation and Research Cell IERC, BRAC International A: Kampala, Uganda | R: bracresearch.net | W: bigd.bracu.ac.bd/ierc |