Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Ceramics International xxx (xxxx) xxx

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

Solution-based fabrication of high-entropy Ba(Ti,Hf,Zr,Fe,Sn)O3 films on


fluorine-doped tin oxide substrates and their piezoelectric responses
Yi-Wei Chen, Jr-Jeng Ruan, Jyh-Ming Ting, Yen-Hsun Su, Kao-Shuo Chang *
Department of Materials Science & Engineering, National Cheng Kung University, No.1, University Road, Tainan City, 70101, Taiwan

A R T I C L E I N F O A B S T R A C T

Keywords: This paper reports the first fabrication of high-entropy Ba(Ti,Hf,Zr,Fe,Sn)O3 films on fluorine-doped tin oxide
Ba(Ti,Hf,Zr,Fe,Sn)O3 film substrates through spin coating and hydrothermal synthesis. A homogeneous distribution of all constituent el­
High-entropy oxide ements was observed through energy-dispersive x-ray spectroscopy mapping. High distortion (approximately
Hydrothermal synthesis
1.7%) of the lattice structure was caused by the underlying substrate and partial substitution of Ti with other
Piezoelectric coefficient
Ferroelectricity
larger elements of Hf, Zr, Fe, and Sn on B sites. Two types of Ba(Ti,Hf,Zr,Fe,Sn)O3 films were studied: high-
density and slightly textured polycrystalline particles (Sample A), and randomly distributed [110]-oriented
nanorods with marginal crystallinity (Sample B). In the piezoelectric experiment, Samples A and B substan­
tially outperformed the medium-entropy oxide (MEO) samples and single-crystal ZnO bulk. Sample A also
exhibited the highest piezoelectric coefficient (d33 ~ 60 pm/V) among all studied samples. Its superior perfor­
mance was attributable to a combination of high crystallinity, strong out-of-plane distortion, slightly textured
polycrystalline structures, and cocktail effects. Furthermore, enhanced piezoelectricity was achieved after poling
for the high-entropy oxide (HEO) and MEO samples, indicating their ferroelectric effect. The d33 of Sample A
more than doubled (≈126.7 pm∙V− 1), and the increase was substantially higher than that of Sample B. Our HEO
samples have promise for use in piezoelectric- and ferroelectric-related applications.

1. Introduction to viable applications for engineering.


Among high-entropy ceramics, high-entropy oxides (HEOs) are a
Typical piezoelectric structures include wurtzite (e.g., AlN, CdS, and particularly promising option because of the ready availability of
ZnO), perovskite [e.g., Pb(Zr,Ti)O3 (PZT) and BaTiO3 (BTO)], and manufacturing facilities and the reactive oxygen–related sources used to
quartz (e.g., GeO2, GaAsO4, and AlPO4) [1,2]. Among them, perovskite produce them. In 2015, Rost et al. first studied entropy-stabilized oxides,
PZT has been the most studied and applied since World War II because of whose properties were driven by configurational disorder [7]. They used
its exceptional spontaneous polarization without external stress. How­ a solid-state reaction to synthesize (Co,Cu,Mg,Ni,Zn)O by mixing equi­
ever, toxic Pb has negative impacts on the environment and human molar CoO, CuO, MgO, NiO, and ZnO. The resulting powders were then
health; thus, the search for alternatives has persisted. fired in air (700 ◦ C− 1100 ◦ C) and quenched. A single-phase quinary
High-entropy alloys and equiatomic multicomponent alloys were rock salt (Co,Cu,Mg,Ni,Zn)O was formed at 850 ◦ C− 900 ◦ C, and the
independently developed by Yeh et al. [3] and Cantor et al. [4] in 2004, formation of a multiphase mixture of rock salt and tenorite was observed
respectively. They comprise at least five principal elements with 5–35 at. below that range. The entropy-stabilized (reversible) phase evolution as
% each. According to Yeh’s theory [3], high-entropy alloys have a a function of temperature indicated the increased (or decreased)
configurational entropy (ΔSconf) greater than 1.5 R (R: gas constant) at a contribution of configurational entropy at high (or low) temperatures.
random state, whereas medium-entropy alloys have ΔSconf between 1 Furthermore, the removal of any component oxide resulted in multi­
and 1.5 R [5]. The high-entropy effect has been extended to ceramics phase states, implying that the decrease in ΔSconf reduced the stability
[6], which covers a wide range of classes, including oxide, carbide, (increased the Gibbs free energy) of associated quaternary compounds.
nitride, silicide, diboride, and sulfide. The broad compositional space The same HEO system has also been investigated by other groups.
provides a diversity of unexplored functional properties that could lead Berardan et al. [8] examined dopant-doped (e.g., Li+, In3+, Ga3+, and

* Corresponding author.
E-mail address: kschang@mail.ncku.edu.tw (K.-S. Chang).

https://doi.org/10.1016/j.ceramint.2020.12.272
Received 30 November 2020; Received in revised form 18 December 2020; Accepted 29 December 2020
Available online 5 January 2021
0272-8842/© 2020 Elsevier Ltd and Techna Group S.r.l. All rights reserved.

Please cite this article as: Yi-Wei Chen, Ceramics International, https://doi.org/10.1016/j.ceramint.2020.12.272
Y.-W. Chen et al. Ceramics International xxx (xxxx) xxx

Ti4+) (Co,Cu,Mg,Ni,Zn)O dielectrics. Only Li+ doping could stabilize a been limited to exploring films fabricated on Si or MgO substrates
single-phase structure because of charge compensation mechanisms that through physical vapor deposition techniques. Furthermore, studies on
could occur through the oxidation of Co+2 (or Ni+2) to Co+3 (or Ni+3) or morphology adjustment and piezoelectricity are still lacking. In this
through the formation of oxygen vacancies. Furthermore, Sarkar et al. study, a facile spin-coating and hydrothermal approach was developed
[9] employed nebulized spray pyrolisis, flame spray pyrolysis, and to fabricate various Ba(Zr,Hf,Sn,Fe,Ti)O3 films on fluorine-doped tin
reverse coprecipitation to obtain nanocrystalline (Co,Cu,Mg,Ni,Zn)O. oxide (FTO) substrates. The influence of their morphology, film orien­
Other fascinating single-phase HEOs have also been studied, tation, and crystallinity on the resulting piezoelectricity was investi­
including perovskite [10–14], rock salt [15–19], spinel [20,21], fluorite gated to gain insight into their ability to optimize piezoelectric devices.
[22,23], and pyrochlore [24,25], produced through various fabrication
approaches, such as solid-state reactions [16,17,21–26], physical vapor 2. Experimental procedure
deposition [12,27], solution combustion synthesis [15,18,28], hydro­
thermal synthesis [29,30], and a sonochemical-based method [10]. FTO/glass substrates were ultrasonically cleaned using acetone,
Certain representative studies are discussed as follows. Braun et al. [19] ethanol, isopropyl alcohol, and deionized (DI) water in sequence for 5
reported that the thermal conductivity of rock salt (Mg0.2Ni0.2Cu0.2 min each to remove both organic and inorganic contaminants before the
Co0.2Zn0.2)O was doubly reduced by the local disorder of ionic charges experiment. The samples were then prepared through spin coating,
without weakening its mechanical stiffness when another element of followed by hydrothermal synthesis. To prepare the spin-coating solu­
equal molarity (e.g., Sc, Sb, Sn, Cr, and Ge) was added to the system. tion, 0.122 g of BaCl2 2H2O, 0.129 mL of TiCl3, 0.032 g of HfCl4, 0.05 g
Sharma et al. [12] observed extremely low thermal conductivity in of ZrOCl2 8H2O, 0.027 g of FeCl3 6H2O, and 0.035 g of SnCl4 5H2O (Ba:
single-crystal epitaxial thin films of Ba(Zr0.2Sn0.2Ti0.2Hf0.2Nb0.2)O3 Ti:Hf:Zr:Fe:Sn = 5:1:1:1:1:1 in molar ratio) were dissolved in 10 mL of
grown on MgO and SrTiO3 substrates. Their respective conductivities ethanol and stirred for 30 min. Next, 0.25 mL of the solution was spin
[approximately 0.58 and 0.54 W⋅(m⋅K)− 1] were close to the theoretical coated twice onto a cleaned FTO/glass substrate at 500 rpm for 5 s,
limit for cubic-phase BaTiO3 in an amorphous state [approximately 0.48 followed by 2000 rpm for 20 s. The substrate was then dried at 150 ◦ C
W⋅(m⋅K)− 1], which suggests that they may have potential for use in for 3 min in air and naturally cooled to room temperature. The entire
thermoelectric power generation applications and thermal barrier process was performed 10 times before the substrate was annealed at
coatings. Berardan et al. [17] studied the ionic conductivity of (Co,Cu, 400 ◦ C for 2 h on a hot plate to obtain HEO-seed-layer–coated substrates.
Mg,Ni,Zn)1-xAxO (A = Li, Na, and K) and (Co,Cu,Mg,Ni,Zn)1-2xLixGaxO. The same process was also applied to obtain medium-entropy oxide
Their results revealed that (Co,Cu,Mg,Ni,Zn)1-xLixO outperformed a (MEO)-seed-layer–coated substrates, in which each of the constituent
state-of-the-art lithium phosphorous oxynitride solid electrolyte by elements on the B sites of Ba(Ti,Hf,Zr,Fe,Sn)O3 was excluded through
more than 2 orders of magnitude at room temperature. The excellent the removal of associated precursor sources. Hereafter, MEO− Ti and
ionic conductivity (>10− 3 S cm− 1) resulted from oxygen vacancies MEO− Fe indicate Ba(Hf,Zr,Fe,Sn)O3 and Ba(Ti,Hf,Zr,Sn)O3,
induced by Li+ doping because of charge compensation. respectively.
Further-enhancement of ionic conductivity could be obtained through During hydrothermal synthesis, 0.732 g of BaCl2 2H2O, 0.26 mL of
the optimization of oxygen-vacancy concentrations/or ordering and TiCl3, 0.128 g of HfCl4, 0.1 g of ZrOCl2 8H2O, 0.108 g of FeCl3 6H2O, and
divalent cation sizes in the system. Sarkar et al. [18] reported that en­ 0.14 g of SnCl4 5H2O (Ba:Ti:Hf:Zr:Fe:Sn = 15:1:2:1:2:2 in molar ratio)
tropy stabilization substantially improved the storage capacity retention were dissolved in 40 mL of DI water. Various complex agents, including
and cycling stability of (Co0.2Cu0.2Mg0.2Ni0.2Zn0.2)O, whereas severe polyvinylpyrrolidone (PVP), hexamethylenetetramine (HMTA), poly
degradation in capacity was observed for the samples without Zn, Cu, or (vinyl alcohol) (PVA), citric acid, and ethylenediaminetetraacetic acid
Co. Berardan et al. [8] investigated the electrical properties of (Mg,Co, disodium salt (EDTA− 2Na), were added to the precursor solution under
Ni,Cu,Zn)O, (Mg,Co,Ni,Cu,Zn)0.95Li0.05O, and (Li,Mg,Co,Ni,Cu,Zn)O. extensive stirring for 30 min. Suitable amounts of KOH tablets were used
They observed that (Mg,Co,Ni,Cu,Zn)0.95Li0.05O exhibited a high rela­ to adjust the pH to approximately 13.5. The prepared solution and a
tive permittivity of 2 × 105 at 20 Hz and 440 K, which was retained in a HEO-seed-layer–coated FTO/glass substrate were subjected to hydro­
wide frequency range (2.3 MHz–100 Hz). The dielectric and electro­ thermal reactions at 200 ◦ C for 2 h and then naturally cooled to room
caloric properties of (Na0.2Bi0.2Ba0.2Sr0.2Ca0.2)TiO3 perovskite powders temperature. The substrate was removed from the autoclave and washed
at various temperatures were also explored by Pu et al. [13] for appli­ with DI water several times to obtain the Ba(Ti,Hf,Zr,Fe,Sn)O3 films. The
cations in solid-state refrigeration and energy storage. Furthermore, residuals in the autoclave were also collected through centrifugation,
Zhou et al. [14] reported the fabrication of Ba(Zr0.2Ti0.2Sn0.2Hf0.2Me0.2) thoroughly washed with DI water and ethanol, and subsequently dried
O3 (Me = Y3+, Nb5+, Ta5+, V5+, Mo6+, and W6+) perovskite powders, in at 90 ◦ C for 4 h to obtain the Ba(Ti,Hf,Zr,Fe,Sn)O3 powders. In addition,
which single-phase Ba(Zr0.2Ti0.2Sn0.2Hf0.2Ta0.2)O3 and Ba to enhance the crystallinity of the EDTA− 2Na involved Ba(Ti,Hf,Zr,Fe,
(Zr0.2Ti0.2Sn0.2Hf0.2Nb0.2)O3 respectively exhibited high dielectric con­ Sn)O3 nanorod films, the samples were further annealed at 600 ◦ C for 3 h
stants of approximately 49 and 60 at 2 MHz. Mao et al. [20] studied the after the hydrothermal process. The same process was followed to obtain
room-temperature magnetic properties of spinel (Cr0.2Fe0.2Mn0.2Co0.2 the MEO films and powders, excluding the annealing process at 600 ◦ C
Ni0.2)3O4, (Cr0.2Fe0.2Mn0.2Ni0.2Zn0.2)3O4, and (Cr0.2Fe0.2Mn0.2 because they already displayed adequate crystallinity.
Co0.2Zn0.2)3O4. When Zn2+ was substituted for either Co2+or Ni2+, the X-ray diffraction (XRD) and two-dimentional (2D) microdiffraction
saturation and remanence magnetization and coercivity decreased (D8 DISCOVER with GADDS, Bruker AXS Gmbh, Karlsruhe, Germany),
because the nonmagnetic Zn2+ ions reduced the number of equipped with a Cu (Kα) radiation source (λ = 0.154 nm), and scanning
super-exchange interactions among their magnetic moments. Further­ electron microscopy (SEM) were employed to characterize the crystal­
more, Meisenheimer et al. [31] reported the existence of linity, phases, textures, and morphology of the samples. The micro­
entropy-stabilized antiferromagnetic permalloy− (Mg0.25(1-x)CoxNi0.25 structures and crystallinity were further examined using high-resolution
(1-x)Cu0.25(1-x)Zn0.25(1-x))O heterostructures, as confirmed by the obser­ transmission electron microscopy (HRTEM) and selective area electron
vation of a critical blocking temperature and anisotropic magnetic diffraction (SAED). Energy-dispersive spectroscopy (EDS) in TEM and
exchange. inductively coupled plasma− mass spectrometry (ICP− MS) were per­
Although studies on HEOs have gained popularity, most of them formed to determine the elemental distribution and stoichiometric
have focused on studying the mechanical strength [31,36], thermal and characteristics of the samples, respectively. The piezoelectric coefficient
ionic conductivity [12,17,19], dielectric properties [13,14], and (d33) was determined using piezoelectric force microscopy (PFM), which
magnetism [20,31] of powders fabricated through solid-state reactions. was performed on the basis of detection of local vibrations induced by an
Few studies have been conducted on HEO films, and most of them have AC signal (1 kHz) applied between a conductive tip (force of

2
­
­
­
Y.-W. Chen et al. Ceramics International xxx (xxxx) xxx

approximately 70 nN) and a bottom electrode of the samples [32]. With considerably limits their wide applicability for use in next-generation
the d33 of approximately 9.93 pm∙V− 1 of the single-crystal ZnO bulk as a sensors and piezoelectric devices. Therefore, two strategies were
reference, the d33 of the samples can be obtained as follows [Eq. (1)]: designed to improve the developed hydrothermal approach: (1) adopt­
ing a two-step hydrothermal route using various complex agents in each
d33,sample = (slopesample/slopeZnO) × 9.93 (1) step, and (2) using seed-layer spin coating first, followed by a hydro­
where d33,sample is the piezoelectric constant of the samples, and slope thermal approach (without complex agents). The first approach slightly
sample and slopeZnO denote the slopes of the samples and ZnO, respec­ improved the growth of the Ba(Ti,Hf,Zr,Fe,Sn)O3 films [Fig. S2(a)] over
tively, in a plot of amplitude against AC voltage. A slow scan rate was the single-step hydrothermal approach [red, Fig. 2(a)]; however, the
employed to reliably obtain the topography and amplitudes of the film density remained poor [Fig. S2(b)]. By contrast, the second
samples. Furthermore, to maximize the piezoelectric response of the approach considerably improved both the film crystallinity [orange,
samples, a poling process was used by applying a high DC voltage Fig. 2(a)] and density [Fig. 2(b)] (Sample A) compared with the single-
(approximately 4.2 kV) across two parallel metal plates (resulting step (Fig. 2) and two-step (Fig. S2) hydrothermal approaches. Further­
electric field of approximately 1 kV mm− 1) for 30 min. more, the resulting substantial strain (approximately 1.7%) relative to
that obtained for the powder samples was attributable to the lattice
3. Results and discussion mismatch between the film and underlying substrate, in addition to the
lattice distortion, as displayed in Fig. 1. Furthermore, the 2D micro­
The BTO powders were first obtained hydrothermally through the diffraction pattern [azimuthal angle (χ) vs 2θ] [Fig. 2(c)] revealed that
adjustment of various precursor ratios and pH values (PDF#05-0626) the Ba(Ti,Hf,Zr,Fe,Sn)O3 films had a slightly textured polycrystalline
(red, Fig. 1). The resulting strain (approximately 0.13%) was estimated structure because of a nonuniform ring associated with the (110) plane.
using a Williamson–Hall plot of (βhkl cosθ) vs (4 sinθ), where βhkl is the To adjust the morphology of Sample A, various amounts and types of
intensity of the full width at half maximum for hkl (Miller index) and θ is complex agents, including PVP, HMTA, PVA, citric acid, and
the diffraction angle [33,34]. The strain was due to size confinement, in EDTA− 2Na, were added to the hydrothermal reaction of the second
which atomic configurations in the BTO nanocrystals were modulated approach. However, only the addition of EDTA− 2Na produced dramatic
relative to those in the BTO bulk. The procedure was then modified to changes, and the others exhibited ineffective morphological tailoring for
obtain the Ba(Ti,Hf,Zr,Fe,Sn)O3 powders (green, Fig. 1), whose char­ the Ba(Ti,Hf,Zr,Fe,Sn)O3 films [Fig. S3(a)–(d)]. However, the addition
acteristic peaks shifted toward lower angles relative to those associated of EDTA− 2Na produced dramatic changes. After the addition of 0.185 g
with the BTO powders. This indicated a higher distortion (approxi­ of EDTA− 2Na, nanorods mixed with irregular particles [Fig. 3(a)] and
mately 0.15%) of the lattice structure because of the substitution of Ti on exhibiting decent crystallinity [blue, Fig. 3(d)] were observed. When the
B sites by the larger elements Hf, Zr, Fe, and Sn. Furthermore, the amount of EDTA− 2Na was doubled (approximately 0.37 g), high-
ICP− MS result revealed nearly equimolar ratios of the constituent ele­ density nanorods formed [Fig. 3(b)], possibly because of the resulting
ments on B sites of the Ba(Ti,Hf,Zr,Fe,Sn)O3 powders (Fig. S1). complex species during the hydrothermal reactions. These species could
The same procedure was then applied to the fabrication of the Ba(Ti, selectively absorb or desorb onto various facets of Ba(Ti,Hf,Zr,Fe,Sn)O3
Hf,Zr,Fe,Sn)O3 films on various substrates [e.g., Si, indium tin oxide crystals to cause anisotropic crystal growth [35]. However, their crys­
(ITO) and FTO] through a single-step hydrothermal route. Our results tallinity was substantially deteriorated [green, Fig. 3(d)] compared with
confirmed that the films only grew on FTO substrates [red, Fig. 2(a)]. that of the of EDTA− 2Na (0.185 g) sample and Sample A. When greater
The characteristic peaks at approximately 30.8◦ and 44.2◦ were asso­ amounts of EDTA− 2Na (approximately 0.74 g) were involved, both the
ciated with the planes indexed at (110) and (200) of Ba(Ti,Hf,Zr,Fe,Sn) nanorod densities [Fig. 3(c)] and their crystallinity [red, Fig. 3(d)] were
O3, respectively. Compared with the Ba(Ti,Hf,Zr,Fe,Sn)O3 powders substantially reduced. Thus, crystalline Ba(Ti,Hf,Zr,Fe,Sn)O3 nanorods
(green, Fig. 1), no plane indexed at (111) (approximately 38.8◦ ) was were not obtained through simple adjustment of the amounts of
observed, indicating the growth of textured films, which is further dis­ EDTA− 2Na. The second approach was then further modified on the
cussed in the section describing PFM measurement results. This dem­ basis of the EDTA− 2Na (0.37 g) sample under various annealing con­
onstrates the first hydrothermal fabrication of HEO films on a ditions, including heat treatments (<400 ◦ C) during the first-step
transparent conducting oxide substrate, potentially creating opportu­ spin-coating process or postannealing after the second-step hydrother­
nities for use in optoelectronic engineering and devices. No films grew mal reaction. Our results revealed that Ba(Ti,Hf,Zr,Fe,Sn)O3 nanorods
on Si [blue, Fig. 2(a)] or ITO [green, Fig. 2(a)] substrates because these were unfavorably converted to particles at room temperature [Fig. S4
substrate were etched under strongly basic conditions (pH ≈ 13.5) (a)], 200 ◦ C [Fig. S4(b)], and 300 ◦ C [Fig. S4(c)] in the first-step spin
during the hydrothermal reactions. coating without postannealing; however, their crystallinity was
However, the quality of the resulting Ba(Ti,Hf,Zr,Fe,Sn)O3 film re­ enhanced, particularly at room temperature and 200 ◦ C [blue and red in
quires improvement to address its nonuniformity and low density, which Fig. S4(d), respectively]. By contrast, the sample fabricated at 400 ◦ C
during spin coating and postannealed at 600 ◦ C for 3 h (Sample B)
retained the Ba(Ti,Hf,Zr,Fe,Sn)O3 nanorods [Fig. 4(a)] and exhibited
certain, although weak, crystallinity [Fig. 4(b)], which was also
observed in the 2D microdiffraction pattern [Fig. 4(c)]. Its resulting
strain was then analyzed through TEM analysis (Fig. 6).
Because Samples A (excellent-crystallinity particle) and B (marginal-
crystallinity nanorod) exhibited distinct features, they were further
studied using TEM. The SAED pattern [Fig. 5(a)] captured from the
particle (big white circle present in the inset) exhibited multiple and
discrete diffraction spots, indicating that various orientations of Ba(Ti,
Hf,Zr,Fe,Sn)O3 nanoparticles were involved. A d-spacing of approxi­
mately 0.2844 (d110) and 0.2046 (d200) nm was associated with (110)
and (200) planes of Ba(Ti,Hf,Zr,Fe,Sn)O3, respectively. Increases of
approximately 0.7% and 2.4% were observed in comparison with the
Fig. 1. XRD results of hydrothermally fabricated BaTiO3 (red) and Ba(Ti,Hf,Sn, nominal d110 (≈0.2825 nm) and d200 (≈0.1997 nm), respectively, for
Zr,Fe)O3 (green) powders. (For interpretation of the references to colour in this the BTO bulk. Another SAED pattern [small white circle shown in the
figure legend, the reader is referred to the Web version of this article.) inset of Fig. 5(a)] illustrated strong spot arrays, which were indexed at

3
­
Y.-W. Chen et al. Ceramics International xxx (xxxx) xxx

Fig. 2. (a) XRD results of hydrothermally fabricated Ba(Ti,Hf,Zr,Fe,Sn)O3 films on Si (blue), ITO (green), and FTO (red) substrates through a single-step approach as
well as on FTO (orange) substrates through spin coating and a hydrothermal route (Sample A). SEM image (b) and 2D microdiffraction pattern (c) of Sample A. (For
interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

Fig. 3. Various morphologies of Ba(Ti,Hf,Zr,Fe,Sn)O3 films adjusted with various amounts of EDTA− 2Na through spin coating and a hydrothermal route. (a) 0.185 g.
(b) 0.37 g. (c) 0.74 g. (d) XRD results of the three samples.

(211), (102), and (313) planes [Fig. 5(b)], indicating the formation of (Lα1), which was indistinguishable from the Ti (Kα1) peak in the EDS
single-crystalline Ba(Ti,Hf,Zr,Fe,Sn)O3 nanoparticles. EDS mapping spectrum.
revealed homogeneous distributions of all constituent elements [Fig. 5 Fig. 6(a) displays an HRTEM image of Sample B, captured from the
(c)]. The strip-like distribution at the bottom of the Sn and O images is red circle shown in the inset. Predominant and minor fringes associated
attributable to FTO substrates. Furthermore, elemental compositions with planes indexed at (110) (d110 ≈ 0.2955 nm) and (200) (d200 ≈
[lower right, Fig. 5(c)] showed that the Ti content was substantially 0.2060 nm) of Ba(Ti,Hf,Zr,Fe,Sn)O3 were observed, respectively. Sub­
higher than the content of other elements because of the mutual con­ stantial increases of approximately 4.6% and 3.1% were observed
tributions of Ba and Ti. The Ti signal was partially contributed by Ba compared with the respective nominal d-spacing values for the BTO

4
Y.-W. Chen et al. Ceramics International xxx (xxxx) xxx

Fig. 4. Characterizations of Sample B. (a) SEM image. (b) XRD result. (c) 2D microdiffraction pattern.

Fig. 5. TEM results of Sample A. (a) SAED pattern taken from the big circle presented in the inset. (b) SAED pattern taken from the small circle presented in the inset
in (a). (c) EDS mapping. Elemental compositions are presented in the lower-right corner for each element.

5
Y.-W. Chen et al. Ceramics International xxx (xxxx) xxx

Fig. 6. TEM results of Sample B. (a) High-resolution image captured from a red circle marked in the inset. (b) SAED pattern taken from the same location described in
(a). (c) EDS mapping. Elemental compositions are presented in the lower-right corner for each element. (For interpretation of the references to colour in this figure
legend, the reader is referred to the Web version of this article.)

bulk, which could primarily be attributable to the one-dimensional the XRD result (Fig. 4), in which poor crystallinity was observed, highly
growth of nanorods, in addition to the lattice distortion of Ba(Ti,Hf,Zr, textured [110]-oriented nanocrystalline nanorods were fabricated. EDS
Fe,Sn)O3. Strong diffraction spots observed in the SAED [Fig. 6(b)], mapping [Fig. 6(c)] showed a uniform distribution for each constituent
taken from the same region as shown in the inset of Fig. 6(a), were also element, confirming the fabrication of high-entropy Ba(Ti,Hf,Zr,Fe,Sn)
associated with the (110) plane of Ba(Ti,Hf,Zr,Fe,Sn)O3. Combined with O3 nanorod films. The atomic ratio for each element is shown in the

Fig. 7. (a) Plot of amplitude vs AC voltage obtained through PFM of various samples, including Sample A, Sample B, MEO− Ti, MEO− Fe, single-crystal ZnO bulk,
FTO, and a piece of glass. Solid and dashed lines denote before and after poling, respectively. (b) Calculated slopes and d33. (c) Topography of Sample A.

6
Y.-W. Chen et al. Ceramics International xxx (xxxx) xxx

lower right corner. The Ti content was also substantially higher than the was considerably greater than that for Sample B, indicating a predom­
content of other elements, as discussed for Sample A. inant factor of crystallinity for enhancing its ferroelectric response. Our
PFM [32] was performed to determine the d33 of various samples, results indicate the potential of Ba(Ti,Hf,Zr,Fe,Sn)O3 films for piezo­
including Samples A and B, MEO− Ti, and MEO− Fe. The single-crystal electric- and ferroelectric-based applications.
ZnO bulk, FTO, and a piece of glass were also studied for comparison.
Details are provided in the Experimental section. Fig. 7(a) shows a plot Declaration of competing interest
of amplitude vs AC voltage for the various samples. Negligible slopes
were observed for FTO (solid pink line) and glass (solid brown line), The authors declare that they have no known competing financial
indicating their expected lack of piezoelectric response. By contrast, a interests or personal relationships that could have appeared to influence
moderate slope (solid purple line) was observed for single-crystal ZnO, the work reported in this paper.
indicating a piezoelectric response (d33 ≈ 9.9 pm/V). In general, the
piezoelectric performance of the HEO samples [Sample A (solid red line) Acknowledgments
and Sample B (solid blue line)] was substantially better than that of the
MEO samples [MEO− Ti (solid black line) and MEO− Fe (solid green This study was partially supported by the Ministry of Science and
line)]. This might be attributable to the severe lattice distortion and a Technology, Taiwan, under the grant number MOST 107-2218-E-006-
cocktail effect [3] of the HEO samples and a partial loss of non­ 047.
centrosymmetry in the MEO samples. Furthermore, all of the HEO and
MEO samples exhibited greater slopes than did ZnO, indicating their Appendix A. Supplementary data
excellent piezoelectric response. The calculated d33 for each sample is
presented in Fig. 7(b). Sample A [topography shown in Fig. 7(c)] Supplementary data to this article can be found online at https://doi.
exhibited the highest d33 (≈60 pm∙V− 1), followed by Sample B (≈35 org/10.1016/j.ceramint.2020.12.272.
pm∙V− 1), MEO− Fe (≈7.9 pm∙V− 1), and MEO− Ti (≈7.5 pm∙V− 1),
respectively. In addition to better crystallinity and slightly textured References
polycrystalline structures (Fig. 2) that were effective at improving
piezoelectric response [36], the superior performance of Sample A to [1] Y.-T. Wang, K.-S. Chang, Piezopotential-induced Schottky behavior of Zn1-xSnO3
nanowire arrays and piezophotocatalytic applications, J. Am. Ceram. Soc. 99
that of Sample B was partially attributable to strong out-of-plane
(2016) 2593–2600.
distortion (spontaneous polarization along a z-direction for perov­ [2] Y.-C. Chiang, K.-S. Chang, Fabrication of Ba1− xZnO2 nanowires using solvothermal
skite). As indicated in the TEM analyses (Figs. 5 and 6), increases in the synthesis: piezotronic and piezophototronic performance characterization, Appl.
in-plane d110 and d200 for Sample A (approximately 0.7% and 2.4%, Surf. Sci. 445 (2018) 71–76.
[3] J.-W. Yeh, S.-K. Chen, S.-J. Lin, J.-Y. Gan, T.-S. Chin, T.-T. Shun, C.-H. Tsau, S.-
respectively) were smaller than those for Sample B (approximately 4.6% Y. Chang, Nanostructured high-entropy alloys with multiple principal elements:
and 3.1%, respectively). Assuming the same volume of Ba(Ti,Hf,Zr,Fe, novel alloy design concepts and outcomes, Adv. Eng. Mater. 6 (2004) 299–303.
Sn)O3 unit cells in Samples A and B, substantial out-of-plane distortion [4] B. Cantor, I.T.H. Chang, P. Knight, A.J.B. Vincent, Microstructural development in
equiatomic multicomponent alloys, Mater. Sci. Eng., A 375− 377 (2004) 213–218.
was expected in Sample A. However, Sample B still exhibited a fairly [5] M.C. Gao, J.-W. Yeh, P.K. Liaw, Y. Zhang, High-Entropy Alloys, Springer,
high d33 value, despite its overall poor crystallinity because of nano­ Switzerland, 2016.
scopic polarity in nearly amorphous states [37], as evidenced by the [6] R.-Z. Zhang, M.J. Reece, Review of high entropy ceramics: design, synthesis,
structure and properties, J. Mater. Chem. A 7 (2019) 22148–22162.
XRD and TEM analyses. The piezoelectricity of Sample B could be [7] C.M. Rost, E. Sachet, T. Borman, A. Moballegh, E.C. Dickey, D. Hou, J.L. Jones,
considerably intensified when [001]-oriented crystalline Ba(Ti,Hf,Zr,Fe, S. Curtarolo, J.-P. Maria, Entropy-stabilized oxides, Nat. Commun. 6 (2015)
Sn)O3 nanorod arrays could be fabricated. A poling process was applied 8485–8488.
[8] D. Bérardan, S. Franger, D. Dragoe, A.K. Meena, N. Dragoe, Colossal dielectric
to study the ferroelectricity of the samples. After poling, enhanced
constant in high entropy oxides, Phys. Status Solidi Rapid Res. Lett. 10 (2016)
piezoelectricity was observed for most of the piezoelectric samples 328–333.
(dashed lines), indicating their ferroelectric effect. No apparent varia­ [9] A. Sarkar, R. Djenadic, N.J. Usharani, K.P. Sanghvi, V.S.K. Chakravadhanula, A.
S. Gandhi, H. Hahn, S.S. Bhattacharya, Nanocrystalline multicomponent entropy
tion was observed for ZnO between before and after poling because of its
stabilised transition metal oxides, J. Eur. Ceram. Soc. 37 (2017) 747–754.
lack of ferroelectric response. The d33 of Sample A (dashed red line, [10] F. Okejiri, Z. Zhang, J. Liu, M. Liu, S. Yang, S. Dai, Room-temperature synthesis of
≈126.7 pm∙V− 1) more than doubled after poling. The value was still far high-entropy perovskite oxide nanoparticle catalysts through ultrasonication-based
superior to those of all other samples, and the ferroelectric improvement method, ChemSusChem 13 (2020) 111–115.
[11] S. Jiang, T. Hu, J. Gild, N. Zhou, J. Nie, M. Qin, T. Harrington, K. Vecchio, J. Luo,
was substantially greater than that observed for Sample B (dashed blue A new class of high-entropy perovskite oxides, Scripta Mater. 142 (2018) 116–120.
line), indicating a predominant factor of crystallinity [38]. [12] Y. Sharma, B.L. Musico, X. Gao, C. Hua, A.F. May, A. Herklotz, A. Rastogi,
D. Mandrus, J. Yan, H.N. Lee, M.F. Chisholm, V. Keppens, T.Z. Ward, Single-crystal
high entropy perovskite oxide epitaxial films, Phys. Rev. Mater. 2 (2018),
4. Conclusions 060404− 6.
[13] Y. Pu, Q. Zhang, R. Li, M. Chen, X. Du, S. Zhou, Dielectric properties and
High-entropy Ba(Ti,Hf,Zr,Fe,Sn)O3 piezoelectric films with various electrocaloric effect of high-entropy (Na0.2Bi0.2Ba0.2Sr0.2Ca0.2)TiO3 ceramic, Appl.
Phys. Lett. 115 (2019) 223901–223905.
morphologies on FTO substrates were fabricated through the adjustment [14] S. Zhou, Y. Pu, Q. Zhang, R. Shi, X. Guo, W. Wang, J. Ji, T. Wei, T. Ouyang,
of pH values, complex agents, and annealing temperatures using a facile Microstructure and dielectric properties of high entropy Ba
spin-coating and hydrothermal method. EDS mapping revealed homo­ (Zr0.2Ti0.2Sn0.2Hf0.2Me0.2)O3 perovskite oxides, Ceram. Int. 46 (2020) 7430–7437.
[15] A. Mao, H.-Z. Xiang, Z.-G. Zhang, K. Kuramoto, H. Yu, S. Ran, Solution combustion
geneous distributions of all constituent elements for the fabricated HEO synthesis and magnetic property of rock-salt (Co0.2Cu0.2Mg0.2Ni0.2Zn0.2)O high-
films. Two distinct types of Ba(Ti,Hf,Zr,Fe,Sn)O3 films were studied: entropy oxide nanocrystalline powder, J. Magn. Magn Mater. 484 (2019) 245–252.
Sample A, consisting of slightly textured polycrystalline particles, and [16] W. Hong, F. Chen, Q. Shen, Y.-H. Han, W.G. Fahrenholtz, L. Zhang, Microstructural
evolution and mechanical properties of (Mg,Co,Ni,Cu,Zn)O high-entropy ceramics,
Sample B, comprising highly textured [110]-oriented nanorods J. Am. Ceram. Soc. 102 (2019) 2228–2237.
(adjusted by EDTA− 2Na) with marginal crystallinity. Sample A [17] D. Bérardan, S. Franger, A.K. Meena, N. Dragoe, Room temperature lithium
exhibited a superior d33 value (≈60 pm∙V− 1) compared with other superionic conductivity in high entropy oxides, J. Mater. Chem. A 4 (2016)
9536–9541.
samples, including Sample B (≈35 pm∙V− 1), MEO− Fe (≈7.9 pm∙V− 1),
[18] A. Sarkar, L. Velasco, D. Wang, Q. Wang, G. Talasila, L.D. Biasi, C. Kübel,
MEO− Ti (≈7.5 pm∙V− 1), and single-crystal ZnO bulk (≈9.9 pm∙V− 1). T. Brezesinski, S.S. Bhattacharya, H. Hahn, B. Breitung, High entropy oxides for
The excellent piezoelectric performance was attributable to the combi­ reversible energy storage, Nat. Commun. 9 (2018) 3400–3409.
nation of high crystallinity, strong out-of-plane growth, slightly textured [19] J.L. Braun, C.M. Rost, M. Lim, A. Giri, D.H. Olson, G.N. Kotsonis, G. Stan, D.
W. Brenner, J.P. Maria, P.E. Hopkins, Charge-induced disorder controls the
polycrystalline structures, and cocktail effects. Furthermore, its d33 thermal conductivity of entropy-stabilized oxides, Adv. Mater. 30 (2018)
more than doubled (≈126.7 pm∙V− 1) after poling, and the enhancement 1805004–1805008.

7
Y.-W. Chen et al. Ceramics International xxx (xxxx) xxx

[20] A. Mao, H.-Z. Xiang, Z.-G. Zhang, K. Kuramoto, H. Zhang, Y. Jia, A new class of [30] L. Spiridigliozzi, C. Ferone, R. Cioffi, G. Accardo, D. Frattini, G. Dell’Agli, Entropy-
spinel high-entropy oxides with controllable magnetic properties, J. Magn. Magn stabilized oxides owning fluorite structure obtained by hydrothermal treatment,
Mater. 497 (2020) 165884–165885. Materials 13 (2020), 558− 12.
[21] Z. Grzesik, G. Smoła, M. Miszczak, M. Stygar, J. Dąbrowa, M. Zajusz, K. Świerczek, [31] P.B. Meisenheimer, T.J. Kratofil, J.T. Heron, Giant enhancement of exchange
M. Danielewski, Defect structure and transport properties of (Co,Cr,Fe,Mn,Ni)3O4 coupling in entropy-stabilized oxide heterostructures, Sci. Rep. 7 (2017)
spinel-structured high entropy oxide, J. Eur. Ceram. Soc. 40 (2020) 835–839. 13344–13346.
[22] J. Gild, M. Samiee, J.L. Braun, T. Harrington, H. Vega, P.E. Hopkins, K. Vecchio, [32] D. Seol, B. Kim, Y. Kim, Non-piezoelectric effects in piezoresponse force
J. Luo, High-entropy fluorite oxides, J. Eur. Ceram. Soc. 38 (2018) 3578–3584. microscopy, Curr. Appl. Phys. 17 (2017) 661–674.
[23] K. Chen, X. Pei, L. Tang, H. Cheng, Z. Li, C. Li, X. Zhang, L. An, A five-component [33] B.D. Cullity, S.R. Stock, Elements of X-Ray Diffraction, third ed., Prentice Hall, New
entropy-stabilized fluorite oxide, J. Eur. Ceram. Soc. 38 (2018) 4161–4164. York, USA, 2001.
[24] A.J. Wright, Q. Wang, S.-T. Ko, K.M. Chung, R. Chen, J. Luo, Size disorder as a [34] D. Nath, F. Singh, R. Das, X-ray diffraction analysis by Williamson-Hall, Halder-
descriptor for predicting reduced thermal conductivity in medium- and high- Wagner and size-strain plot methods of CdSe nanoparticles- a comparative study,
entropy pyrochlore oxides, Scripta Mater. 181 (2020) 76–81. Mater. Chem. Phys. 239 (2020) 122021–122029.
[25] F. Li, L. Zhou, J.-X. Liu, Y. Liang, G.-J. Zhang, High-entropy pyrochlores with low [35] W.-C. Lu, H.-D. Nguyen, C.-Y. Wu, K.-S. Chang, M. Yoshimura, Modulation of
thermal conductivity for thermal barrier coating materials, J. Adv. Ceram. 8 (2019) physical and photocatalytic properties of (Cr, N) codoped TiO2 nanorods using soft
576–582. solution processing, J. Appl. Phys. 115 (2014) 144305–144308.
[26] A.R. West, Solid State Chemistry and its Applications, John Wiley & Sons, Ltd, [36] M. Acosta, N. Novak, V. Rojas, S. Patel, R. Vaish, J. Koruza, G.A. Rossetti, J. RÖdel,
Chichester, UK, 1991. BaTiO3-based piezoelectrics: fundamentals, current status, and perspectives, Appl.
[27] A. Kirnbauer, C. Spadt, C.M. Koller, S. Kolozsvári, P.H. Mayrhofer, High-entropy Phys. Rev. 4 (2017), 041305− 53.
oxide thin films based on Al–Cr–Nb–Ta–Ti, Vacuum 168 (2019) 108850–108855. [37] D. Yu, M. Zhao, C. Wang, L. Wang, W. Su, Z. Gai, C. Wang, J. Li, J. Zhang,
[28] M.R. Chellali, A. Sarkar, S.H. Nandam, S.S. Bhattacharya, B. Breitung, H. Hahn, Enhanced piezoelectricity in plastically deformed nearly amorphous Bi12TiO20-
L. Velasco, On the homogeneity of high entropy oxides: an investigation at the BaTiO3 nanocomposites, Appl. Phys. Lett. 109 (2016), 032904− 5.
atomic scale, Scripta Mater. 166 (2019) 58–63. [38] J.D. Mackenzie, Y. Xu, Ferroelectric materials by the sol-gel method, J. Sol. Gel Sci.
[29] M. Biesuz, L. Spiridigliozzi, G. Dell’Agli, M. Bortolotti, V.M. Sglavo, Synthesis and Technol. 8 (1997) 673–679.
sintering of (Mg, Co, Ni, Cu, Zn)O entropy-stabilized oxides obtained by wet
chemical methods, J. Mater. Sci. 53 (2018) 8074–8085.

You might also like