Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Japanese Journal of Applied

Physics

REGULAR PAPER You may also like


- Characterization of (111)-oriented epitaxial
Growth of epitaxial (K, Na)NbO3 films with various (K0.5Na0.5)NbO3 thick films deposited by
hydrothermal method
orientations by hydrothermal method and their Takahisa Shiraishi, Mutsuo Ishikawa,
Hiroshi Uchida et al.

properties - Ferroelectric and piezoelectric properties


of KNbO3 films deposited on flexible
organic substrate by hydrothermal method
To cite this article: Yoshiharu Ito et al 2019 Jpn. J. Appl. Phys. 58 SLLB14 Takahisa Shiraishi, Noriyuki Kaneko,
Mutsuo Ishikawa et al.

- In-plane orientation and composition


dependences of crystal structure and
electrical properties of {100}-oriented
View the article online for updates and enhancements. Pb(Zr,Ti)O3 films grown on (100) Si
substrates by metal organic chemical
vapor deposition
Shoji Okamoto, P. S. Sankara Rama
Krishnan, Satoshi Okamoto et al.

This content was downloaded from IP address 140.114.252.97 on 05/11/2023 at 05:03


Japanese Journal of Applied Physics 58, SLLB14 (2019) REGULAR PAPER
https://doi.org/10.7567/1347-4065/ab3958

Growth of epitaxial (K, Na)NbO3 films with various orientations by hydrothermal


method and their properties
Yoshiharu Ito1, Akinori Tateyama1, Yoshiko Nakamura1, Takao Shimizu1 , Minoru Kurosawa2, Hiroshi Uchida3,
Takahisa Shiraishi4 , Takanori Kiguchi4 , Toyohiko J. Konno4 , Mutsuo Ishikawa5, and Hiroshi Funakubo1*
1
Department of Materials Science and Engineering, Tokyo Institute of Technology, J2–43, 4259 Nagatsuta-cho, Midori-ku, Yokohama, Kanagawa,
226–8502, Japan
2
Department of Electrical and Electronic Engineering, Tokyo Institute of Technology, G2–32, 4259 Nagatsuta-cho, Midori-ku, Yokohama, Kanagawa,
226–8502, Japan
3
Department of Materials and Life Sciences, Sophia University, 7–1 Kioi-cho, Tokyo, 102–8554, Japan
4
Institute for Materials Research, Tohoku University, Building 2, 6th Floor, 2–1–1, Katahira, Aobaku, Sendai, Miyagi, 980–8577, Japan
5
Department of Clinical Engineering, Toin University of Yokohama, 1614, Kurogane-cho, Aoba-ku, Yokohama, Japan, 225–8503
*
E-mail: funakubo.h.aa@m.titech.ac.jp
Received June 6, 2019; revised June 26, 2019; accepted August 6, 2019; published online September 2, 2019

{100}c-, {110}c- and {111}c-oriented epitaxial (K, Na)NbO3 thin films were grown at 240 °C on (100)cSrRuO3//(100)SrTiO3, (110)cSrRuO3//
(110)SrTiO3, and (111)cSrRuO3//(111)SrTiO3 substrates by the hydrothermal method. Their film thicknesses increased with the deposition time
and then eventually saturated at longer deposition times. Their saturated film thicknesses were mainly determined by their orientation and the order
was {100}c-, {110}c- and {111}c-orientation regardless of any experimental conditions. These films consisted of grains with characteristic
morphologies. All of the films exhibited similar ferroelectric and piezoelectric properties irrespective of the film orientation. The remnant
polarizations (Pr) and coercive fields (Ec) of the {100}c-, {110}c- and {111}c-oriented films at the maximum electric field of 500 kV cm−1 were
31 μC cm−2 and 111 kV cm−1, 27 μC cm−2 and 94 kV cm−1, and 25 μC cm−2 and 110 kV cm−1, respectively, while their effective values of
piezoelectric coefficient (d33) were approximately 31–33 pm V−1. Similar films are associated with its mixed domain structure.
© 2019 The Japan Society of Applied Physics

ferroelectric and piezoelectric properties of hydrothermally


1. Introduction deposited (K, Na)NbO3 films, orientation control is one of the
Alkaline niobates, tantalates, and their solid solutions have crucial factors as they strongly depend on the crystal
attracted significant interest owing to their superior ferro- orientation.31–35) In addition, the orientation can also affect the
electric, piezoelectric, pyroelectric, photorefractive, and photo- film thickness of the resulting films, i.e., the deposition rate, as
catalytic properties.1–7) Among them, the solid-solution of frequently observed for other perovskite-based ferroelectric
KNbO3 and NaNbO3, i.e., (K, Na)NbO3 and their related films such as PZT films.36) However few studies have reported
materials have been proposed as alternative lead-free piezo- the dependence of orientation for thin film used by the
electric materials to Pb(Zr, Ti)O3, because (K, Na)NbO3-based hydrothermal technique.
materials have better piezoelectric properties and higher Curie We already reported a successful growth of epitaxial
temperatures than other lead-free piezoelectric materials.8–11) KNbO3 films on SrRuO3-coated SrTiO3 single-crystal sub-
These materials are expected to be applied for various strates with (100), (110) and (111) cut.33) In this study, we
applications including sensors, actuators, and vibration energy prepared epitaxial (K, Na)NbO3 films on SrRuO3-coated
harvesters,12–18) in the form of thin film as well as the sintered SrTiO3 single-crystal substrates with each crystal cuts by the
body. The fabrication of (K, Na)NbO3-based materials is hydrothermal method. In addition, we investigated the effects
generally challenging owing to the high volatilities of K and of the crystal orientation on the deposition rate and surface
Na components, leading to K- and Na-deficient composition morphology of the deposited films as well as ferroelectric and
that degraded their piezoelectric properties. This difficulty was piezoelectric properties of these films.
also reported for films prepared by sputtering, sol-gel, and
chemical vapor deposition techniques.19–25) Therefore, the 2. Experimental methods
preparation method that allows us to depress the volatilizations (K, Na)NbO3 films were deposited on (100)cSrRuO3//
of them is highly required. (100)SrTiO3, (110)cSrRuO3//(110)SrTiO3 and (111)cSrRuO3//
In the present study, we focused on the depositions of (111)SrTiO3 substrates by the hydrothermal method. 50-nm-
(K, Na)NbO3 films by the hydrothermal method. The hydro- thick epitaxial SrRuO3 layers were grown by the radio-
thermal method prevents the volatilization of K and Na frequency magnetron sputtering method as the bottom electrodes
components because the films are prepared at a lower tempera- on single-crystal SrTiO3 substrates.
ture below 300 °C under high pressure. L. S. Wojciech reported Aqueous solutions of KOH and NaOH (Kanto Chemical
the growth of epitaxial KNbO3 films deposited on SrTiO3 and Co., Inc., 5–7 mol l−1, [KOH]/([KOH] + [NaOH]) = 0.9,
LiTaO3 single crystal substrates and Goh et al. the growth on 20 cm3) and Nb2O5 powder (Kanto Chemical Co., Inc.,
SrTiO3 by the hydrothermal method.26,27) In our previous study, 0.25–2.00 g) were used as starting materials. The materials
the film thickness of hydrothermally grown KNbO3 increased and substrates put into a Teflon bottle were sealed in an
using conductive bottom layers of SrRuO3 with a perovskite autoclave. The substrates were suspended by a Teflon fixture
structure.28) Based on these results, we reported the growth of that is almost the same as reported one by Morita.37) Three
epitaxial (K, Na)NbO3 films with various K/(K + Na) ratio on kinds of substrates were put into the same Teflon bottle at one
single-crystal substrates at 240 °C.29–31) In order to enhance the time. The autoclave was placed in an oven kept at 240 °C for
SLLB14-1 © 2019 The Japan Society of Applied Physics
Jpn. J. Appl. Phys. 58, SLLB14 (2019) Y. Ito et al.

The thicknesses and chemical compositions of the obtained


films were determined by X-ray fluorescence spectroscopy
(XRF, Panalytical Axios Advance PW4400) calibrated using
standard samples. The film thickness of (K0.88Na0.12)NbO3
films was estimated by the intensity of Nb(Kα) because it
(a) was ascertained to linearly increase with the film thicknesses
by cross-sectional scanning electron microscopy (SEM,
JEOL JEM-6610LA). The crystal structures and preferred
orientations of the films were characterized by X-ray
diffraction (XRD, Philips X’Pert MRD system) with Cu
Kα1 radiation. The pseudo-cubic unit cell denoted by (hkl)c
was employed for SrRuO3 despite their orthorhombic sym-
metry because of the small crystal anisotropy, according to
the previous study.28) The pseudo-cubic cell was also used
for (K, Na)NbO3 in this study. Figure 2(a) shows the
relationships between the orthorhombic unit cell and
(b) pseudo-cubic cell, where the axis of (K, Na)NbO3 orthor-
hombic unit cell is denoted as [hkl]o and [hkl]pc, and their
polar axis is along [001]o, [101]pc, respectively. Figure 2(b)
shows a schematic illustration of orientation relationship of
{100}c-oriented (K, Na)NbO3 epitaxial film deposited on
(100)cSrRuO3//SrTiO3 for the easy understanding of the
relationship between (K, Na)NbO3, and SrRuO3 and
SrTiO3. In this case, the index by a pseudo-cubic cell of (K,
Na)NbO3 denoted as [hkl]c is found to be easier to understand
their crystal structure relationship. The pseudo-cubic cell is
regarded as a special monoclinic lattice reflected in the
(c) orthorhombic symmetry, where the β angle is not equal to
90° but the lattice length of [100] and [001] is equal.40) It is
expected to observe the splitting of diffraction peaks from
(K0.88Na0.12)NbO3 film into 0k0pc and h00pc/00lpc, in view of
Fig. 1. (Color online) (a) Temperatures of the air and the autoclave surface XRD studies.
in the oven as a function of the deposition time. Film thicknesses as a The surface morphologies of the films were observed by
function of the deposition time for the films prepared with (b) various scanning electron microscopy (SEM, JEOL JEM-6610LA).
concentrations of the KOH and NaOH mixed solution (green squares, blue
Pt top electrodes with a diameter of 100 μm were deposited
triangles, and red circles represent concentrations of 5, 6, and 7 mol l−1,
respectively) and (c) input mass of Nb2O5 (red closed circles and open by electron-beam evaporation to fabricate a capacitor struc-
diamonds represent input mass of Nb2O5 of 0.25 and 2.00 g, respectively). ture of Pt/(K0.88Na0.12)NbO3/SrRuO3. The ferroelectric and
piezoelectric properties were measured at room temperature
using a ferroelectric tester (Toyo Corp., FCE Fast) and an
3–24 h for a hydrothermal reaction. The deposition time was atomic force microscope (AFM; SII SPA400) calibrated
counted from the moment when the autoclave was placed in using standard samples, respectively.
the oven. Figure 1(a) shows the temperatures of the air and
the autoclave surface in the oven as a function of the 3. Results and discussion
deposition time. The setting temperature of the oven was 3.1. Deposition characteristics
240 °C. It took approximately 2 h to saturate the air and Figures 1(b) and 1(c) show the film thickness as a function of
autoclave temperatures. After the deposition time, the ob- the deposition time for {100}c-oriented (K0.88Na0.12)NbO3
tained films were ultrasonically cleaned with distilled water, films on (100)cSrRuO3//(100)SrTiO3 substrates prepared with
ethanol, acetone, and methanol, and then dried at 150 °C for various source-material conditions. The film deposition
1 h in the dashed air under atmospheric pressure. started from the deposition time of 2 h and then the film
In this research, we selected (K0.88Na0.12)NbO3 because thickness increased with the deposition time irrespective of
this composition shows the highest deposition rate in our the source conditions. This time period of 2 h almost
previous work.29) Also, epitaxial (K0.88Na0.12)NbO3 film corresponds to the saturation time of the autoclave tempera-
showed a relatively large piezoelectric response together ture as shown in Fig. 1(a). In Fig. 1(b), the film thickness
with relatively low dielectric constant among those of the (K, almost linearly increased with the deposition time. The slopes
Na)NbO3 films that are favorable for the sensor and vibration of these relations, corresponding to the deposition rate,
harvester applications.38) In addition, no phase separation decreased with the decrease in the concentration of the
was observed in the (K0.88Na0.12)NbO3 film, in contrast to the KOH and NaOH mixture, from 7 to 5 mol l−1. The film
two-phase coexistence of the (K0.75Na0.25)NbO3 and thickness was then saturated after a certain time. Also, as
(K0.36Na0.64)NbO3 films.39) Details of the [KOH]/([KOH]+ shown in Fig. 1(c), the both of film thickness and time before
[NaOH]) dependency on the composition and the crystal saturation were almost independent on the input mass of
structure of the obtained film were already reported.38) Nb2O5 powder. Our previous work clarified that the
SLLB14-2 © 2019 The Japan Society of Applied Physics
Jpn. J. Appl. Phys. 58, SLLB14 (2019) Y. Ito et al.

(a) (b)

Fig. 2. (Color online) (a) Relationships between the orthorhombic unit cell and pseudo-cubic unit cell of (K, Na)NbO3 and (b) the schematic illustration of
orientation relationship of {100}c-(K, Na)NbO3 epitaxial film deposited on (100)cSrRuO3//(100)SrTiO3.

hydrothermal process proceeds with competitive reactions of (K0.88Na0.12)NbO3 films deposited on all oriented substrates.
film deposition and powder formation and that the saturation Each film thicknesses increased almost linearly with the
time agreed with the time required for consuming the residual deposition time within the time range of 2–5 h irrespective of
Nb source.41) The data shown in Fig. 1(b) and 1(c) revealed the film orientation, where their deposition rates depended on
that the consumption time of the residual Nb source each orientation of their substrates. The order of the deposition
decreased when the concentration of the mixture solution rate agreed well with that of the saturation thickness [see
increased. On the other hand, the deposition rate of the film Fig. 3(a)]. The deposition rates of (K0.88Na0.12)NbO3 films on
before saturation time clearly shows that the deposition of the (100)cSrRuO3//(100)SrTiO3, (110)cSrRuO3//(110)SrTiO3 and
films was rate-limited by the concentration of the KOH and (111)cSrRuO3//(111)SrTiO3 were 2.9, 2.0 and 1.8 μm h−1,
NaOH within the limit of the present condition. respectively. The deposition rates of (K0.88Na0.12)NbO3 films on
Further, the dependences of the film thickness on the (100)cSrRuO3//(100)SrTiO3 and (110)cSrRuO3//(110)SrTiO3
orientation of the substrate, which will correspond to the film substrates were 1.5 and 1.1 times larger than that on
orientation, were evaluated under various hydrothermal condi- (111)cSrRuO3//(111)SrTiO3 substrate. As discussed later, the
tions. Figure 3(a) shows the film thicknesses as a function of the ratios of the deposition rate accord to the ratios of the saturation
deposition time for (K0.88Na0.12)NbO3 films deposited on thickness. This indicates that the saturated thickness was
(100)cSrRuO3//(100)SrTiO3, (110)cSrRuO3//(110)SrTiO3 and determined by the deposition rate of each oriented film. This
(111)cSrRuO3//(111)SrTiO3 substrates. The concentration of is because these times for starting deposition and saturating the
the KOH/NaOH solution and input mass of Nb2O5 powder film thickness were almost independent of the film orientation.
were fixed at 7 mol l−1 and 0.25 g, respectively. All of the films Figures 3(c) and 3(d) show the film thicknesses as a
deposited on their oriented substrates exhibited similar trends function of the concentration of the KOH/NaOH solution and
for their film thicknesses against the deposition time. Their film input mass of Nb2O5 powder, respectively. Their film
thicknesses increased with the deposition time and then thicknesses increased monotonously with the concentration
saturated above 4 h. No dependency of the saturation time on of the KOH/NaOH solution in the range of 5–7 mol l−1 for all
the substrate orientation is reasonable because the saturation oriented films. On the other hand, their film thicknesses
time was mainly determined by the time for consuming the Nb increased with the input mass of Nb2O5 powder up to 1 g and
source in solution. The film thicknesses after saturation time in then decreased with excessive Nb2O5, irrespective of the
Fig. 3(a) were ordered as follows: {100}c-, {110}c- and orientations of their substrates. The film thicknesses were
{111}c-oriented films. Figure 3(b) shows their film thicknesses ordered as follows: {100}c-, {110}c- and {111}c-oriented
as a function of the deposition time enlarged from 0 to 5 h for films, similarly to the result in Fig. 3(a). The decrease in the
SLLB14-3 © 2019 The Japan Society of Applied Physics
Jpn. J. Appl. Phys. 58, SLLB14 (2019) Y. Ito et al.

(a)
(b)

(c)

Fig. 4. (Color online) Film thickness of the {100}c- oriented


(K0.88Na0.12)NbO3 films (red circles) and {110}c-oriented (K0.88Na0.12)NbO3
films (blue triangles) plotted against that of the {111}c-oriented
(K0.88Na0.12)NbO3 films shown in Fig. 3. The dashed black line denotes the
case when the film thickness of {100}c- and {110}c-oriented films are the
same as that of {111}c-oriented films.

(111)cSrRuO3//(111)SrTiO3 substrates, using experimental


data in Fig. 3. These three kinds of films were simultaneously
deposited at the same batch. The dashed (black) line in
Fig. 4 indicates that the film thicknesses deposited on
(100)cSrRuO3//(100)SrTiO3 and (110)cSrRuO3//(110)SrTiO3
(d) substrates were coincident with that on the (111)cSrRuO3//
(111)SrTiO3 substrate, suggesting no dependency on the
crystal orientation. The obtained film thicknesses increased in
the order {100}c-, {110}c- and {111}c-oriented films, irre-
spective of the deposition conditions shown in Fig. 3. In
addition, almost all of the experimental results were almost
on the straight lines across zero, indicating the ratios of the
film thicknesses also independent of the deposition condi-
tions. This means that these film thicknesses of the {100}c-
and {110}c-oriented films were approximately 1.6 and 1.2
times larger than those of the {111}c -oriented film irrespec-
tive of the deposition conditions. Note that the data in Fig. 4
includes not only these thicknesses of films deposited
for a longer time than saturation time but also that for a
shorter time. In fact, these ratios of these film thicknesses
Fig. 3. (Color online) (a) Film thicknesses as a function of the deposition
time for 0–18 h and (b) the enlarged one for 0–5 h with 7 mol l−1 KOH/
almost agreed with the ratios of the deposition rate shown in
NaOH solution and 0.25 g Nb2O5 powder, together with film thicknesses as a Fig. 3(b); the deposition rates of {100}c- and {110}c-oriented
function of (c) the concentration of KOH/NaOH solution (deposition time: (K0.88Na0.12)NbO3 film were 1.5 and 1.1 times larger than
6 h, Nb2O5 powder: 0.25 g) and (d) input mass of Nb2O5 powder (deposition that of {111}c-oriented (K0.88Na0.12)NbO3 film. This sug-
time: 6 h, KOH/NaOH solution: 7 mol l−1). The films were simultaneously gests that the resultant thickness variation by orientations is
deposited at the same batch on (100)cSrRuO3//(100)SrTiO3 (red circles),
(110)cSrRuO3//(110)SrTiO3 (blue triangles), and (111)cSrRuO3//(111)SrTiO3
caused by the deposition rate that depends on the film
(green squares) substrates. orientations.
The similar tendency, i.e. the anisotropy of the deposition
rate, was reported for Pb(Zr, Ti)O3 films deposited on (100)c,
film thicknesses at 2 g Nb2O5 in Fig. 3(d) can be attributed to (110)c, and (111)cSrRuO3//SrTiO3 substrates prepared by
the homogenous nucleation of the (K0.88Na0.12)NbO3 powder metal organic chemical vapor deposition.42) In that case,
accelerated by excessive Nb2O5 addition, which consumes the deposition rates of {100}c- and {110}c-oriented films
the Nb2O5 source in the reaction system and then suppresses were approximately 1.7 and 1.2 times larger than that of the
the film deposition on single crystal substrates. {111}c-oriented film, similar to the present study for
Figure 4 shows the thicknesses of these films deposited on (K0.88Na0.12)NbO3 films by hydrothermal deposition. This
(100)cSrRuO3//(100)SrTiO3 and (110)cSrRuO3//(110)SrTiO3 could be considered to be attributed to the crystal orientation
substrates plotted against that of the films deposited on dependence of the growth rate of the perovskite oxide film,
SLLB14-4 © 2019 The Japan Society of Applied Physics
Jpn. J. Appl. Phys. 58, SLLB14 (2019) Y. Ito et al.

(a) (b) (c)


(a) (d)

(e)
(b)

(c) (f)

Fig. 6. (Color online) XRD θ–2θ profiles near (a) (K, Na)NbO3 {300}c,
Fig. 5. (Color online) (a)–(c) XRD θ–2θ profiles and (d)–(f) X-ray pole (b) (K, Na)NbO3{220}c and (c) (K, Na)NbO3 {222}c, for (K0.88Na0.12)NbO3
figures of the (K0.88Na0.12)NbO3 films grown on [(a), (d)] (100)cSrRuO3// films grown on (a) (100)cSrRuO3//(100)SrTiO3, (b) (110)cSrRuO3//
(100)SrTiO3, [(b), (e)] (110)cSrRuO3//(110)SrTiO3, and [(c), (f)] (110)SrTiO3 and (c) (111)cSrRuO3//(111)SrTiO3 substrates, respectively.
(111)cSrRuO3//(111)SrTiO3 substrates. Data shown in (d)–(f) was measured
at fixed 2θ angles corresponding to the (K0.88Na0.12)NbO3{110}c, {100}c,
and {111}c for films deposited on [(a), (d)] (100)cSrRuO3//(100)SrTiO3,
[(b), (e)] (110)cSrRuO3//(110)SrTiO3 and [(c), (f)] (111)cSrRuO3// (110)cSrRuO3//(110)SrTiO3 and (111)cSrRuO3//(111)SrTiO3
(111)SrTiO3 substrates, respectively. substrates, respectively. These peaks were found to separate
into two or three, suggesting the presence of a multi-domain
structure. {300}c peaks consist of about 82% of (003)pc and
perhaps due to the crystal structure and surface energy (300)pc and 18% of (030)pc ones in integrated intensities as
anisotropies. shown in Fig. 6(a) for the (K0.88Na0.12)NbO3 films deposited on
3.2. Crystal structures and surface morphologies of (100)cSrRuO3//(100)SrTiO3 substrates, while {220}c peaks
the deposited films consist of about 23% (202)pc, 34% of (20-2)pc and 43% of
In this section, crystal structure and surface morphology were (022)pc, (0–22)pc, (220)pc and (2–20)pc for the films deposited
observed using these films with 3–4 μm in thickness grown on (110)cSrRuO3//(110)SrTiO3 substrates. Finally {222}c peaks
on (100)cSrRuO3//(100)SrTiO3, (110)cSrRuO3//(110)SrTiO3, consist of 56% (222)pc and (2–22)pc and 44% of (−222)pc and
and (111)cSrRuO3//(111)SrTiO3 substrates. It must be noted (22-2)pc for the films on (111)cSrRuO3//(111)SrTiO3 substrates.
we cannot detect an obvious change in electrical properties Taking account of the fact that (K0.88Na0.12)NbO3 films have an
with the film thickness within the present experimental orthorhombic symmetry,30) splittings of the XRD peaks can be
conditions. attributed to the domain structure of these films. In fact, R. M.
Figure 5 shows representative XRD θ–2θ patterns and Fernando et al., reported that that the (K, Na)NbO3 ceramics
X-ray pole figures at fixed 2θ angles corresponding to the with orthorhombic symmetry have 180°, 90°, 60° and 120°
(K0.88Na0.12)NbO3 {110}c, {100}c, and {111}c for films domains with [101]c polar axis.43) The existence of these multi
deposited on (100)cSrRuO3//(100)SrTiO3, (110)cSrRuO3// orientations affects the ferroelectric and piezoelectric properties
(110)SrTiO3, and (111)cSrRuO3//(111)SrTiO3 substrates, re- in Sect. 3.3. It must be noted that we confirmed that the shoulder
spectively. These films were deposited by the hydrothermal that exists at the right side of the 003pc and 300pc peak did not
method with KOH/NaOH concentration of 7 mol l−1 and affect these properties.
input Nb2O5 mass of 0.25 g. Only {h00}c, {110}c, and Figures 7(a)–7(c) show the surface morphologies of the
{111}c (K0.88Na0.12)NbO3 peaks were observed in Figs. 5(a) {100}c-, {110}c- and {111}c-oriented (K0.88Na0.12)NbO3
–5(c), respectively, suggesting that {100}c-, {110}c-, and films. Their films consisted of grains with characteristic
{111}c-oriented films were obtained on (100)cSrRuO3// shapes depending on the crystal orientation of the films and
(100)SrTiO3, (110)cSrRuO3//(110)SrTiO3 and (111)cSrRuO3// substrates. That is, the {100}-oriented film consisted of
(111)SrTiO3 substrates. Fourfold-, twofold-, and threefold- square-shaped grains rotated by 45° from [001] SrTiO3 [in
symmetry peaks at azimuth ψ angles of 45, 45, and 75° were Fig. 7(a)]. Roof-shaped grains elongated along [001] SrTiO3
exhibited in Figs. 5(d)–5(f), respectively, which indicates in- were observed for the {110}c-oriented film [Fig. 7(b)].
plane crystal orientation of these films deposited on Finally, triangle-shaped grains with edges along [1–10]
(100)cSrRuO3//(100)SrTiO3, (110)cSrRuO3//(110)SrTiO3 and SrTiO3 and flat top surface were observed for the {111}-
(111)cSrRuO3//(111)SrTiO3 substrates. These results certified oriented film [Fig. 7(c)].
that the epitaxial films with three different crystal orientations The characteristic shapes of the grains are attributed to the
of {100}c-, {110}c- and {111}c were grown successfully by growth mechanisms of the epitaxial (K, Na)NbO3 films on
the hydrothermal method. the surfaces of single-crystal substrates. Considering that the
Figures 6(a)–6(c) show XRD θ–2θ profiles near (K, Na) {100}c-oriented film grows faster than the {110}c- and
NbO3{300}c, (K, Na)NbO3{220}c and (K, Na)NbO3 {222}c for {111}c-oriented films, as shown in Fig. 4, possible scenarios
(K0.88Na0.12)NbO3 films grown on (100)cSrRuO3//(100)SrTiO3, can be illustrated as described in Figs. 7(d)–7(f). First, initial
SLLB14-5 © 2019 The Japan Society of Applied Physics
Jpn. J. Appl. Phys. 58, SLLB14 (2019) Y. Ito et al.

(a) (d)

(b) (e)

(c) (f)

Fig. 7. (Color online) (a)–(c) Surface SEM images and (d)–(f) possible growth models for the (K0.88Na0.12)NbO3 films grown on [(a), (d)]
(100)cSrRuO3//(100)SrTiO3, [(b), (e)] (110)cSrRuO3//(110)SrTiO3, and [(c), (f)] (111)cSrRuO3//(111)SrTiO3 substrates.

nuclei oriented along the surface normal direction and {111}c from XRD patterns in Fig. 6 and projections with respect to
facets grow on all of the substrate surfaces. Subsequently, the surface normal directions of these domains. The expected
crystal facet grew faster along {100}c. Further studies are saturation polarization for the {100}-oriented film is 1.3 times
ongoing to clarify the growth mechanism. larger than those for other orientations. This shows almost good
3.3. Ferroelectric and piezoelectric properties agreement with present experimental results, which show
Ferroelectric and piezoelectric properties were also evaluated remanent polarization {100}-oriented film is 1.1 times larger
using the films with 3–4 μm thick grown on (100)cSrRuO3// than that for {110}-oriented film and 1.2 times larger than that
(100)SrTiO3, (110)cSrRuO3//(110)SrTiO3, and (111)cSrRuO3// for the {111}-oriented film. Thus, the small orientation variation
(111)SrTiO3 substrates shown in Figs. 5–7. Figures 8(a)–8(c) can be explained by the formation of multi-domain structures.
show room-temperature polarization–electric-field (P–E) hyster- Similar small orientation dependency is also reported
esis loops measured at 100 kHz for the (K0.88Na0.12)NbO3 for {100}c, {110}c, and {111}c-oriented tetragonal
films deposited on (100)cSrRuO3//(100)SrTiO3, (110)cSrRuO3// Pb(Zr0.35Ti0.65)O3 films.44) In this case, the predicted Pr
(110)SrTiO3, and (111)cSrRuO3//(111)SrTiO3 substrates. All of values based on semi-quantitative estimation using the
the films exhibited clear hysteresis loops originating from their projection of the spontaneous polarization to the surface
ferroelectricity. Their remnant polarizations (Pr) and coercive normal direction for each domain and its volume fraction also
fields (Ec) at the maximum electric field of 500 kV cm−1 were almost agreed with the experimental results.
31 μC cm−2 and 111 kV cm−1, 27 μC cm−2 and 94 kV cm−1, Figures 8(d)–8(f) show the S–E curves measured at
and 25 μC cm−2 and 110 kV cm−1, respectively. These results 5.2 Hz. These curves originating from the ferroelectricity
suggest that the ferroelectric properties did not depend on the were clearly observed for all of the films. The effective
orientation considerably. To understand the ferroelectric prop- values of piezoelectric coefficient, d33, of the films
erty, we must take into account the effect of the multi-domain on (100)cSrRuO3//(100)SrTiO3, (110)cSrRuO3//(110)SrTiO3,
structure as already discussed in Fig. 6. This multi-domain and (111)cSrRuO3//(111)SrTiO3 substrates estimated from
structure is considered to result in the small anisotropy of the strain curves were 33, 31, and 31 pm V−1, respectively.
the remnant polarization of these films. The polar axis of These results suggested that the piezoelectricity also did not
(K, Na)NbO3 with orthorhombic symmetry is 〈101〉c and shows depend so strongly on the film orientation. The multi-domain
180°, 90°, 60° and 120° domains, in general. We expect that structure ascertained in Figs. 7(d)–7(f) should contribute to
saturation polarization of polarization for {100}c-, {110}c- and the piezoelectric response, similar to the ferroelectric polar-
{111}c-oriented (K0.88Na0.12)NbO3 films should be 0.57Ps, ization. However, the piezoelectric coefficient is more
0.45Ps, and 0.45Ps, respectively, where Ps is a spontaneous complicated because it is a third-order tensor. Therefore,
polarization, taking into account volume fractions estimated here we discuss them indirectly; that is, generally, the d33 can
SLLB14-6 © 2019 The Japan Society of Applied Physics
Jpn. J. Appl. Phys. 58, SLLB14 (2019) Y. Ito et al.

(a) (b) (c)

(d) (e) (f)

Fig. 8. (Color online) (a)–(c) P–E loops measured at 100 kHz and (d)–(f) strain–electric-field (S–E) loops measured at 5.2 Hz for the (K0.88Na0.12)NbO3
films grown on [(a), (d)] (100)cSrRuO3//(100)SrTiO3, [(b), (e)] (110)cSrRuO3//(110)SrTiO3, and [(c), (f)] (111)cSrRuO3//(111)SrTiO3 substrates.

be described as d33 = 2QPrεrε0, where Q, εr and ε0 are an {111}c-oriented films, almost independent on the deposition
electrostrictive coefficient and relative dielectric constant conditions within the present study. These films consisted of
and dielectric constant of the vacuum, respectively. The grains with characteristic shapes depending on the orienta-
εr values measured for (K0.88Na0.12)NbO3 films grown tions of the substrates, which could be explained by the
on (100)cSrRuO3//(100)SrTiO3, (110)cSrRuO3//(110)SrTiO3 promoted crystal growth along the {100}c direction.
and (111)cSrRuO3//(111)SrTiO3 substrates at 1 kHz were However, significant orientation dependencies of the ferroe-
337, 418 and 453, respectively. The measured εr values are lectricity and piezoelectric response were not observed in the
not so largely varied by the crystal orientation, as well as the P–E and S–E characteristics, which suggest the strong
values of remnant polarization, Pr as observed in Figs. 8(a)– contribution of the non-180° domains.
8(c). Therefore, we can consider again that the d33 values in
Acknowledgments
this study also do not show so marked dependence on the
crystal orientation by assuming the almost constant Q value This research was partially supported by the Japan Science
for each orientation. Note that the order of the value of and Technology Agency and the Adaptable and Seamless
polarization is {100}c > {110}c > {111}c, whereas that of Technology Transfer Program through Target-driven R&D
dielectric constant is {111}c > {110}c > {100}c. These or- (A-STEP) Grant No. JPMJTS1616.
ders in opposite relation would also contribute to the piezo-
ORCID iDs
electric responses without significant dependency against the
crystal orientation. Takao Shimizu https://orcid.org/0000-0001-9508-7601
Takahisa Shiraishi https://orcid.org/0000-0002-2154-0589
4. Conclusions Takanori Kiguchi https://orcid.org/0000-0003-4541-8761
{100}c-, {110}c- and {111}c-oriented epitaxial (K0.88Na0.12) Toyohiko J. Konno https://orcid.org/0000-0002-3396-
NbO3 films were deposited on (100)cSrRuO3//(100)SrTiO3, 1049
(110)cSrRuO3//(110)SrTiO3 and (111)cSrRuO3//(111)SrTiO3 Hiroshi Funakubo https://orcid.org/0000-0002-1106-200X
substrates by the hydrothermal method. These film thick-
nesses increased with increasing the deposition time and then
eventually saturated at longer deposition time for all of these 1) C.-R. Cho and A. Grishin, Appl. Phys. Lett. 75, 268 (1999).
films, independent of the concentration of the KOH/NaOH 2) P. Li, S. Ouyang, G. Xi, T. Kako, and J. Ye, J. Phys. Chem. C 116, 7621
solution, and input mass of Nb2O5 powder. However the (2012).
3) Y. Yang, J. H. Jung, B. K. Yun, F. Zhang, K. C. Pradel, W. Guo, and
deposition rate exhibited significant dependency on the Z. L. Wang, Adv. Mater. 24, 5357 (2012).
orientation of the film and substrate; the deposition rate 4) D. Lin, Z. Li, S. Zhang, Z. Xu, and X. Yao, Solid State Commun. 149, 1646
before saturation increased in the order: {100}c-, {110}c- (2009).

SLLB14-7 © 2019 The Japan Society of Applied Physics


Jpn. J. Appl. Phys. 58, SLLB14 (2019) Y. Ito et al.

5) C. Medrano, P. Gunter, P. Amrhein, and E. Voit, Ferroelectrics 92, 289 26) G. K. L. Goh, C. G. Levi, J. H. Choi, and F. F. Lange, J. Cryst. Growth 286,
(1989). 457 (2006).
6) Q. Fu, X. Wang, C. Li, Y. Sui, Y. Han, Z. Lv, B. Song, and P. Xu, RSC Adv. 27) W. L. Suchanek, Chem. Mater. 16, 1083 (2004).
6, 108883 (2016). 28) Q. Gan, R. A. Rao, C. B. Eom, L. Wu, and F. Tsui, J. Appl. Phys. 85, 5297
7) Y. Huan, X. Wang, W. Hao, and L. Li, RSC Adv. 5, 72410 (2015). (2002).
8) X. Wang, J. Wu, D. Xiao, J. Zhu, X. Cheng, T. Zheng, B. Zhang, X. Lou, 29) T. Shiraishi, H. Einishi, S. Yasui, M. Ishikawa, T. Hasegawa, M. Kurosawa,
and X. Wang, J. Am. Chem. Soc. 136, 2905 (2014). H. Uchida, Y. Sakashita, and H. Funakubo, Jpn. J. Appl. Phys. 50, 09ND11
9) S. Zhang, R. Xia, and T. R. Shrout, J. Electroceramics 19, 251 (2007). (2011).
10) Y. Saito, H. Takao, T. Tani, T. Nonoyama, K. Takatori, T. Homma, 30) T. Shiraishi, N. Kaneko, H. Einishi, T. Shimizu, M. Kurosawa, and
T. Nagaya, and M. Nakamura, Nature 432, 84 (2004). H. Uchida, Jpn. J. Appl. Phys. 52, 09KA11 (2013).
11) J. Fu, R. Zuo, H. Qi, C. Zhang, J. Li, and L. Li, Appl. Phys. Lett. 105, 31) T. Shiraishi, M. Ishikawa, H. Uchida, T. Kiguchi, M. K. Kurosawa,
242903 (2014). H. Funakubo, and T. J. Konno, Jpn. J. Appl. Phys. 56, 10PF04 (2017).
12) N. A. A. Nawir, A. A. Basari, M. S. M. Saat, N. X. Yan, and S. Hashimoto, 32) J. R. Duclère, C. Cibert, A. Boulle, V. Dorcet, P. Marchet, C. Champeaux,
ARPN J. Eng. Appl. Sci. 13, 2993 (2018). A. Catherinot, S. Députier, and M. Guilloux-Viry, Thin Solid Films 517, 592
13) T. Mishima, H. Kotera, K. Shibata, R. Yokokawa, Y. Wakasa, and I. Kanno, (2008).
Sens. Actuators A 171, 223 (2011). 33) H. Einishi, M. Ishikawa, M. Nakajima, S. Yasui, T. Yamada, M. Kurosawa,
14) Y. Zhen, M. Wang, S. Wang, and Q. Xue, Ceram. Int. 40, 10263 (2014). and H. Funakubo, Key Eng. Mater. 485, 199 (2011).
15) K. Shibata, R. Wang, T. Tou, and J. Koruza, MRS Bull. 43, 612 (2018). 34) T. Harigai, Y. Tanaka, H. Adachi, and E. Fujii, Appl. Phys. Express 3,
16) R. Ahmed, F. Mir, and S. Banerjee, Smart Mater. Struct. 26, 085031 (2017). 111501 (2010).
17) K. Asif, A. Zafar, K. Heung Soo, and O. Il-Kwon, Smart Mater. Struct. 25, 35) S. Yamazoe, H. Sakurai, M. Fukada, H. Adachi, and T. Wada, Appl. Phys.
53002 (2016). Lett. 95, 062906 (2009).
18) Z. Li, G. Zhou, Z. Zhu, and W. Li, Energies 9, 1 (2016). 36) S. Yokoyama et al., J. Appl. Phys. 98, 094106 (2005).
19) X. Wang, U. Helmersson, S. Olafsson, S. Rudner, L. D. Wernlund, and 37) T. Morita, Materials (Basel). 3, 5236 (2010).
S. Gevorgian, Appl. Phys. Lett. 73, 927 (1998). 38) T. Shiraishi, H. Einishi, S. Ysui, M. Ishikawa, T. Hasegawa, M. Kurosawa,
20) S.-W. Oh, J. Akedo, J.-H. Park, and Y. Kawakami, Jpn. J. Appl. Phys. 45, H. Uchida, Y. Sakashita, and H. Funakubo, J. Ceram. Soc. Jpn. 121, 627
7465 (2006). (2013).
21) A. Onoe, A. Yoshida, and K. Chikuma, Appl. Phys. Lett. 69, 167 (1996). 39) T. Shiraishi, H. Einishi, T. Shimizu, H. Funakubo, M. Kurosawa, H. Uchida,
22) T. Saito, T. Wada, H. Adachi, and I. Kanno, Jpn. J. Appl. Phys. 43, 6627 N. Kumada, T. Kiguchi, and T. J. Konno, J. Mater. Res. 31, 693 (2016).
(2004). 40) N. Lu, R. Yu, Z. Cheng, Y. Dai, X. Zhang, and J. Zhu, Appl. Phys. Lett. 96,
23) M. Fukada, S. Yamazoe, and T. Wada, 2011 Int. Symp. on Applications of 108 (2010).
Ferroelectrics (ISAF/PFM) and 2011 Int. Symp. on Piezoresponse Force 41) A. Tateyama et al., J. Cryst. Growth 511, 1 (2019).
Microscopy and Nanoscale Phenomena in Polar Materials, 2011, 10.1109/ 42) S. Yokoyama, T. Ozeki, T. Oikawa, and H. Funakubo, Integr. Ferroelectr.
ISAF.2011.6014121. 59, 1429 (2003).
24) K. Shibata, F. Oka, A. Ohishi, T. Mishima, and I. Kanno, Appl. Phys. 43) F. Rubio-Marcos, A. Del Campo, R. López-Juárez, J. J. Romero, and
Express 1, 011501 (2008). J. F. Fernández, J. Mater. Chem. 22, 9714 (2012).
25) F. Söderlind, P. O. Käll, and U. Helmersson, J. Cryst. Growth 281, 468 44) T. Oikawa, M. Aratani, H. Funakubo, K. Saito, and M. Mizuhira, J. Appl.
(2005). Phys. 95, 3111 (2004).

SLLB14-8 © 2019 The Japan Society of Applied Physics

You might also like