You are on page 1of 7

Open Ceramics 15 (2023) 100401

Contents lists available at ScienceDirect

Open Ceramics
journal homepage: www.sciencedirect.com/journal/open-ceramics

Phase transformation of cold-sintered doped barium titanate ceramics


during the post-annealing process
Maliha Siddiqui a, *, Daniel Valášek a, Yang Bai b, David Salamon a
a
Central European Institute of Technology, Brno University of Technology, Purkyňova 123, 612 00, Brno, Czech Republic
b
Microelectronics Research Unit, Faculty of Information Technology and Electrical Engineering, University of Oulu, FI-90014, Oulu, Finland

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

Handling Editor: Dr Catherine Elissalde The cold sintering technique has been developed to densify electroceramics at significantly low temperatures to
tackle high energy consumption as well as compatibility of electroceramic components in modern electronics.
Keywords: However, obstacles still exist since, to obtain desired electrical properties, a post-annealing procedure at elevated
Cold sintering process temperatures must follow to trigger compulsory phase formation and/or grain growth. This work investigates the
Phase transition
BaTiO3 and (Ba,Sr)TiO3 ceramics via both the cold and conventional sintering routes, and then compares their
Annealing
phase transformation behaviors during the post-annealing process. Results suggest that, given similar particle
Electroceramic
Doping and grain sizes, the cold-sintered ceramics tend to retain a substantial part of the pristine tetragonal phase at
Low temperature densification annealing temperatures of <800 ◦ C, while the cubic phase tends to dominate when annealed at >800 ◦ C. This
indicates a process-stimulated phase transformation that may guide or limit the selection of annealing temper­
ature in practice.

1. Introduction surface energy using nanoparticles as the raw materials. Nevertheless, in


practice the densification still occurs in the temperature range of
Barium titanate (BaTiO3) is a well-known multifunctional electro­ 1000–1300 ◦ C. Meanwhile, the high-temperature PLS may lead to
ceramic material exhibiting versatile properties such as high dielectric coarse, overgrown grains, which is considered detrimental to densifi­
permittivity, piezoelectric effect, and ferroelectric switching. This al­ cation by leaving large pores between grains [6]. To solve these issues,
lows it to be used in a broad range of applications, including multilayer several room-temperature densification methods have been proposed,
ceramic capacitors (MLCC), piezoelectric transducers, sensors and ac­ including ultra-low temperature cofired ceramics (ULTCC). However,
tuators, and ferroelectric random-access memories (FeRAM) [1–4]. The the highly unevenly distributed particle sizes of the raw materials in
dielectric, piezoelectric and ferroelectric properties of BaTiO3 ceramics these techniques may cause powder moistening and may also result in
can be influenced by the synthesis method, grain size, and doping. As a problems such as clay-like clusters, non-uniform density, warpage, and
member of the ABO3 perovskite family, BaTiO3 may be doped by, for cracking [7,8]. Another drawback of these methods is the long pro­
instance, Sr2+ on the A site which helps to tune the phase transition cessing time varying from hours to days to consolidate dense ceramics
temperatures and thus to modify the dielectric properties [5]. Like all [9].
ceramics, BaTiO3 usually requires high sintering temperatures ranging The cold sintering process (CSP) has recently been intensively
from 1300 ◦ C to 1400 ◦ C for necessary densification in the conventional, studied and it has the potential to overcome issues of other low-
pressureless sintering (PLS), which has recently raised concerns on high temperature processing methods. The CSP is a non-conventional sin­
energy consumption and thus environmental issues. In addition, the tering technique for densifying ceramic materials at low temperatures of
high sintering temperatures make electroceramics incompatible with <300 ◦ C. It involves liquid-phase sintering and shares common princi­
the lower-temperature procedures adopted in modern electronics (e.g. ples with the hydrothermal synthesis where water is used as a transient
the Si industry) and hence making the cofiring of electroceramic com­ solvent under uniaxial pressure [10]. Initially, the addition of a small
ponents with the substrates less viable. In principle, a potential way to amount of solvent dissolves the surface of the ceramic powders,
reduce the consumption of energy is to minimize the microscopic reducing the interfacial areas and rearranging the particles aided by the

* Corresponding author.
E-mail address: siddiquimaliha021@gmail.com (M. Siddiqui).

https://doi.org/10.1016/j.oceram.2023.100401
Received 1 April 2023; Received in revised form 22 June 2023; Accepted 29 June 2023
Available online 29 June 2023
2666-5395/© 2023 The Authors. Published by Elsevier Ltd on behalf of European Ceramic Society. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
M. Siddiqui et al. Open Ceramics 15 (2023) 100401

capillary force. Subsequently, a uniaxial pressure is applied and rapidly Table 1


transports the particles to achieve an early-stage densification. Finally, a Summary of the compositions and fabrication conditions for different samples in
higher level of densification is reached at the sintering stage thanks to this work.
the dissolution-precipitation mechanism. Therefore, the combination of Sample Dopant Fabrication Method Temperature (◦ C)
pressure and heat makes the CSP more energy-efficient than the PLS BT a
– – –
[11]. For instance, a dense BaTiO3 ceramic (~90% relative density)
Sintered CB2 Ba(OH)2.8H2O CSP 225
could be fabricated at 180 ◦ C via the CSP [10]. However, the solely
PB2 PLS
cold-sintered BaTiO3 ceramics did not possess the desired electrical CS2 Sr(OH)2.8H2O CSP
properties and density. This could be overcome by annealing the PS2 PLS
cold-sintered samples at elevated temperatures of 700–900 ◦ C. Such a Annealed CB5 Ba(OH)2.8H2O CSP 500
further heat treatment modified the grain boundary structures, removed PB5 PLS
CS5 Sr(OH)2.8H2O CSP
the residual BaCO3 amorphous phase, and improved the density [10,12]. PS5 PLS
In recent years, this limitation of CSP has been overcome through the CB8 Ba(OH)2.8H2O CSP 800
implementation of flux-assisted single -step CSP for fabrication of PB8 PLS
BaTiO3-polymer nanocomposites. This approach utilizes eutectic and CS8 Sr(OH)2.8H2O CSP
PS8 PLS
hydrate fluxes, such as NaOH–KOH and Ba(OH)2.8H2O. The hydrate flux
CB11 Ba(OH)2.8H2O CSP 1100
not only enables the densification of BaTiO3-polymer nanocomposites at PB11 PLS
lower temperature without damaging the included polymer but also CS11 Sr(OH)2.8H2O CSP
improves the resistivities of the resulting nanocomposite [13,14]. Since PS11 PLS
2015, the CSP has proven to be a useful low-temperature approach for a
BT-acid treated BaTiO3 powder.
fabricating a variety of dense ceramics and composites. Despite its
success on undoped compositions, the lack of deep understanding of the
mechanisms on doped materials hinders further development of
high-performance, low-temperature fabricated ceramics modified by
additives. For instance, it is widely known that BaTiO3 ceramics need to
be doped by Sr2+ to tune and optimize the functionalities for certain use
cases [15]. Randall et al. tried to use the CSP to fabricate (Ba,Sr)TiO3
ceramics by introducing the dopant, Sr2+, in form of a sintering flux (i.e.,
(Sr(OH)2.8H2O)) instead of as a solid reactant (e.g. SrCO3) in the PLS.
Their (Ba1-xSrx)TiO3 ceramics had a relative density of ~91% but
exhibited poor dielectric properties. It seemed that the Sr2+ concen­
trated in the intergranular regions while the grains remained largely Fig. 1. SEM images of BaTiO3 powders: a) before acid treatment and b) after
undoped. The residual flux resulted in defects at the grain boundaries, acid treatment.
which was considered to be the major reason for the poor dielectric
properties [16]. [17–19].
Therefore, this paper aims to contribute toward a better under­
standing of Sr-doped BaTiO3 with sintering flux fabricated via the CSP. 2.2. Cold sintering process of doped BaTiO3
The influence of the post-annealing on the phase transformation
(tetragonal-to-cubic) and on the microstructure is investigated. In 0.75 g of the acid-treated BaTiO3 powder was mixed with 0.1264 g of
addition to the Sr2+ doping, Ba2+ doping is used as a reference. Along­ Sr(OH)2.8H2O (99.9%, Sigma Aldrich, USA) in a mortar with pestle with
side the CSP, the conventional PLS is also implemented for comparison. 10 wt% of water. In parallel, for another set of samples, 0.75 g of the
The formation of secondary phases during post-annealing, the grain acid-treated BaTiO3 powder was mixed with 0.15 g of Ba(OH)2.8H2O
morphology and grain boundary changes after post-annealing, and the (98%, Merck, Germany) in 10 wt% water using the same procedures.
impact on the low-temperature densification by the CSP compared to the The latter set of samples was used as the standard reference in this work.
PLS are discussed. The molar concentration of the crystallization water was the same in
both sets of the samples. CSP was performed in the spark plasma sin­
2. Experimental procedures tering (SPS) device (SPS-625, Dr. Sinter, Fuji Electronic Industrial Co.,
LTD, Japan). The experimental setup consisted of a high-pressure die
2.1. Pretreatment of BaTiO3 raw powder with tungsten carbide punches, enclosed within a graphite die, which
provided heating similar to the conventional cold sintering process. To
Before the actual fabrication, 4g of BaTiO3 powder (purchased from prevent the self-heating of the punches and their interaction with the
Nanografi, Turkey, 99.95% purity, average particle size 280 nm, sample, the high-pressure die was electrically isolated. The mixture was
tetragonal phase) was stir-mixed with 20 ml of 1 M acetic acid at 80 ◦ C added to the inner die with a diameter of 10 mm, and a constant uniaxial
for 1 h. The powder was then collected by filtration, washed with pressure of 270 MPa was applied while employing two sets of heating
deionized water, and dried at 120 ◦ C for 24 h [16]. The sample com­ rates and dwell times. In the first stage, the heating rate was kept slow at
positions and fabrication conditions are described in Table 1. 3 ◦ C/min until the temperature reached 70 ◦ C and then the temperature
Morphologies of the untreated and acid-treated BaTiO3 powders was held for 20 min. In the following stage, the temperature was
observed under SEM (scanning electron microscope) are shown in Fig. 1. increased to 225 ◦ C at a faster heating rate of 8 ◦ C/min. The samples
The treated particle surface was smooth and became less coarse after were then held for 60 min at 225 ◦ C. It should be noted that the SPS
being heated for 1 h at 80 ◦ C, as shown in Fig. 1b. The acid treatment chamber was kept under vacuum during the entire heating cycle.
step facilitated the densification due to the removal of the BaCO3 im­
purity that was present as a contamination layer (between 0.7 wt% and 2.3. Pressureless sintering of doped BaTiO3
1.2 wt%) at the surface of undoped commercial BaTiO3 powder. The
means that acid treatment decontaminated and modified the BaTiO3 For comparison of phase transformation, another set of the Ba
surface by removing carbonate layer and forming a nanometer amor­ (OH)2⋅8H2O and Sr(OH)2⋅8H2O doped samples was prepared. The stoi­
phous layer which could dissolve easier than a crystalline surface chiometry and mixing process were kept identical to the case of the CSP

2
M. Siddiqui et al. Open Ceramics 15 (2023) 100401

described above. The mixed powders were then uniaxially pressed at 50


MPa in a 12 mm-diameter die under a semiautomatic press (BSML 21,
Brio Hranice LTD, Czech Republic) and sintered conventionally in a
muffle furnace (Clasic CZ LTD, Czech Republic) at 225 ◦ C for 60 min
with a same heating profile as was described above for the CSP samples.

2.4. Post-annealing of doped BaTiO3

To study the influence of annealing on the phase transformation, the


samples sintered at 225 ◦ C were further heat treated at 500 ◦ C, 800 ◦ C
and 1100 ◦ C, consecutively, for 3 h in the muffle furnace (Clasic CZ LTD,
Czech Republic). After each round of treatment, the samples were
characterized before going to the next post-annealing temperature. The
treatment profiles can be found in Table 1. The relative densities of the
sintered and treated samples were calculated from the dimensional
method, i.e., weight/volume.

2.5. Material characterization

The phase compositions of the samples were investigated with the X-


ray diffraction (XRD) at room temperature. Two wavelengths were
adopted during the XRD measurements: (i) Co-Kα radiation (SmartLab
Rigaku, USA) and (ii) Cu-Kα radiation (SmartLab Rigaku, Japan). To
ensure valid visual comparisons, the raw data collected with the Co-Kα
radiation were converted to the equivalence of those of the Cu-Kα ra­
diation. The lattice parameters and crystallite size were extracted by the
Rietveld profile refinement. The reliability factors are in ranges of Rp =
5.30–14.81%, and Rwp = 9.58–18.6% due to the presence of multi­
phases. Prior to the characterization, the cold-sintered samples were
polished with fine abrasive paper to remove the residual graphite papers
which attached to the sample surfaces after the CSP. The microstructure
and elemental distribution of the samples (carbon coated) were inves­
tigated by high resolution SEM (FEI Verios 460L, Thermo Fisher, Czech
Fig. 2. SEM micrographs of the CSP (the left column) and PLS (the right col­
Republic) equipped with the energy-dispersive X-ray (EDX) system. The umn) BaTiO3 samples doped by Ba(OH)2⋅8H2O. The relative density and
average grain size of the sintered samples was determined using the average grain size of each sample is also labelled.
standard linear intercept method (EN ISO 13383–1) and multiplied by a
coefficient 1.56 [20].
Fig. 5 compares the XRD reflections of the CSP and PLS samples as
well as the post-annealed ones. The reflections of the pristine, tetragonal
3. Results
BaTiO3 (ICCD no. 00-005-0626), which could be as the major phase in
all the CB- and CS-samples, are also provided. After being annealed at
Figs. 2 and 3 show the micrographs of the Ba(OH)2⋅8H2O and Sr
>800 ◦ C, the secondary phase, Ba2TiO4, appeared in all samples. Figs. 6
(OH)2⋅8H2O doped BaTiO3 samples, respectively, sintered by the CSP
and 7 reveal the close-ups of selected individual reflections and lattice
and PLS as well as annealed at different temperatures. The relative
parameters respectively. Table 2 lists the identified phase structures
density (represented by the ratio of the measured bulk density to the
with crystallite size of each sample.
theoretical density of pure BaTiO3, 6.03 gcm− 1) of each sample is also
labelled on the corresponding micrograph. The sample codes can be
4. Discussion
referred to those listed in Table 1.
As can be seen, the CB2 and CS2 (cold sintered at 225 ◦ C) had relative
The obtained results in this work indicated significant changes of the
densities of about 80%. Such densities were retained after the annealing
phases after being post-annealed for both the CSP and PLS samples.
of up to 800 ◦ C. Distinctive densification behaviors were discovered
According to Fig. 5 and Table 2, both the CB2 and CS2 formed the
between the CB- and CS-samples after being annealed at 1100 ◦ C, where
secondary phases of BaCO3 and BaO. The sintering flux, Ba(OH)2.8H2O,
the CB11 drastically increased the density to ~97% while, in compari­
as a dopant source for densification was considered suitable for the
son, the CS11 exhibited a negligible improvement of densification to
formation of the BaTiO3 phase at low temperatures as it helped to ease
only ~84%. The change of the densities was also evidenced by the
the precipitation of the solute ions (Ba2+). Thermodynamic modeling
highly porous microstructure in CB2-CB8 and CS2-CS11 as well as the
suggest that in the Ti–Ba(OH)2 system, the formation of the BaTiO3
similar particle sizes in these samples to those of the loose BT powders
phase is dependent on the concentration of the solute ions, temperature,
shown in Fig. 1 but the observed grain growth and significant reduction
and pH value of the solution. The presence of additional Ba(OH)2
of porosity exclusively in CB11 (Figs. 2 and 3).
maintained the necessary high pH value of the solution for the reaction
The PLS samples showed similar trends with the only difference that
of Ba2+ and Ti4+ toward the desired perovskite crystal structure of
the density increase from PB8 to PB11 was identical to that from PS8 to
BaTiO3 at low temperatures. Meanwhile, as a negative factor, it stimu­
PS11, both from approximately 48% to roughly 55%. Another signifi­
lated the formation of undesirable phases, for instance, BaCO3 in a CO2
cant fact was that all the PLS samples, regardless of being sintered at
containing atmosphere, which further easily decomposed to BaO [21].
225 ◦ C or being annealed at 500 ◦ C and 800 ◦ C, barely achieved any
BaCO3 was also found in the CS2, implying that its formation was indeed
noticeable densification. Even the high annealing temperature, 1100 ◦ C,
due to the high pH value caused by the additional Sr(OH)2 that was
was not as helpful as the sole CSP at low temperatures (see CB2 and CS2
analogic to the Ba(OH)2 in the CB2. Therefore, it is recommended to
in Fig. 4a).

3
M. Siddiqui et al. Open Ceramics 15 (2023) 100401

A site. Since the radius of Sr2+ (1.32 Å) is smaller than that of Ba2+ (1.49
Å) and thus is likely to induce contraction of the unit cells, the (110)
reflection of the Ba1-xSrxTiO3 phase should tend to broaden and shift
toward a higher angle compared to that of the BaTiO3 phase [15]. As
shown in Fig. 7a, it was observed that the unit cell volume of CS2
initially decreased compared to CB2. This indicates that the CSP at
225 ◦ C may induce the reaction and result in the substitution of Sr2+ ions
into the BaTiO3 perovskite unit cells. Furthermore, both CB2 and CS2
samples continue to exhibit a decreasing cell volume until reaching
800 ◦ C, which could be attributed to phase transformation. However, a
sharp decrease in the cell volume is observed in CS11, implying the
complete incorporation of Sr2+ ions in the BaTiO3 unit cell. These
finding align to those reported in literature [25,26].
On the other hand, the secondary phase, SrCO3, was observed in the
PS2 (Table 2). The dopant, Sr(OH)2.8H2O, decomposed at around
200 ◦ C and then absorbed CO2 from the air at room temperature, leading
to the unwanted SrCO3, as explained by Equations (4) and (5) [27,28].
210 ◦ C
Sr(OH)2 .8H2 O̅̅̅̅→Sr(OH)2 + 8H2 O (4)

(5)
29 ◦ C
Sr(OH)2 + CO2 ̅̅̅→SrCO3 + H2 O

Furthermore, the split reflections of (002) and (200) at about 44–46◦


which belong to the tetragonal phase in the treated BaTiO3, as shown in
Fig. 6, were largely preserved in the CS2 and CB2. The intensities of the
(002) reflections in the CS2 and CB2 increased compared to those of the
PB2 and PS2, respectively. This implies a substantial change in the phase
quantity. Based on such observations, as well as on that the CB2, CS2,
PB2 and PS2 showed very similar microstructure and grain distributions
(Fig. 4b), it is reasonable to believe that the CSP triggered the phase
transformation. This is considered a new finding that may relate to the
high relative densities of the CSP samples. It should be noted that the
potential concentration of Sr2+ ions in the grain boundaries barely
Fig. 3. SEM micrographs of the CSP (the left column) and PLS (the right col­ played a role, as can be seen in Fig. 8 which reveals a relatively even
umn) BaTiO3 samples doped by Sr(OH)2⋅8H2O. The relative density and distribution of Sr2+ in the bulk sample.
average grain size of each sample is also labelled. The CS2 and CB2 yielded relative densities of about 80% that is
significantly higher than those of the PS2 and PB2 (Fig. 4a). The
avoid the exposure of CO2 during the synthesis of BaTiO3. Equations (1)– improvement in the density after the CSP at such low temperatures was
(3) explain the reactions toward the creation of BaCO3 and its decom­ attributed to the constant pressure applied throughout the heating cycle.
position [22–24]. In addition, the initial powder treatment with acid could also enhance
the density by surface modification as mentioned in pretreatment sec­
(1)
200 ◦ C
Ba(OH)2 .8H2 O̅̅̅̅→Ba(OH)2 + 8H2 O tion [14]. In comparison, the PLS did not have such an advantage and
thus the low densities of the green bodies could only be expected to
25 ◦ C
Ba(OH)2 + CO2 ̅̅̅→BaCO3 + H2 O (2) increase at around 1200 ◦ C which is the usual case for ordinary BaTiO3
ceramics [29].
800◦ C
BaCO3 ̅̅̅̅→ BaO + CO2 (3) After the post-annealing, the phases continued to change in the CB-
and CS-samples with the increase of the annealing temperature
In the conventional solid-state reaction, Sr2+ substitutes the Ba2+ on the (Table 2). The BaCO3 phase was stable up to 800 ◦ C in both the CB- and

Fig. 4. The relative density a) and average grain size b) of the doped CSP samples with Ba(OH)2.8H2O (CB), Sr(OH)2.8H2O (CS) and doped PLS samples with Ba
(OH)2.8H2O (PB), Sr(OH)2.8H2O (PS).

4
M. Siddiqui et al. Open Ceramics 15 (2023) 100401

Fig. 5. XRD patterns of the CSP (left) and PLS (right) BaTiO3 samples doped by a) Ba(OH)2⋅8H2O and b) Sr(OH)2⋅8H2O.

Fig. 6. Close-ups of the (002)/(200) reflections for a) the Ba(OH)2⋅8H2O doped b) the Sr(OH)2⋅8H2O doped CSP and PLS samples.

CS-samples. When subject to 1100 ◦ C, the Ba2TiO4 emerged as another literature [16]. A similar trend was also found in the transition from the
secondary phase. The appearance of secondary phase above 800 ◦ C PS8 to the PS11 with broadening of the (110) reflections, confirming
could be attributed due to the leaching of Ba2+ ions as the H2O is present that the higher annealing temperatures stimulated the Sr2+ ions to enter
during the sintering process [30]. However, according to the BaO–TiO2 the BaTiO3 crystal lattices.
phase diagram, the Ba2TiO4 phase could be stably observed in the The evolution of the major BaTiO3 phase during the post-annealing
Ba-rich region at >1000 ◦ C [31]. Due to the presence of the BaO and can be systematically seen in Fig. 6. Researchers have widely demon­
BaCO3 phases in the samples of this work, it was expected that the strated that BaTiO3 is a mixture of two symmetries, i.e., the pseudocubic
Ba2TiO4 phase was formed at the higher annealing temperature phase dominated in nanoparticles and the tetragonal phase dominated
following the routes of Equations (6) and (7) [32,33]. in sub-micron particles. The tetragonal-cubic transition temperature
decreases with the decrease of the particle size [34]. In this work, the
BaO + BaTiO3 → Ba2 TiO4 (6)
CSP at 225 ◦ C triggered phase transformation, which was evident from
800− 900 ◦ C the XRD patterns at ~45◦ in both the CB2 and CS2 but still with some
BaTiO3 + BaCO3 ̅̅̅̅̅̅̅̅→Ba2 TiO4 + CO2 (7) extent of the tetragonal phase remained in the structures with the
The (110) reflections of the CS-samples gradually shifted to higher visualized peak splitting at ~45◦ . This apparent peak splitting gradually
angles and then broadened with the increase of the annealing temper­ diminished upon being annealed at 500–1100 ◦ C, which showed the
ature (Fig. 5b). This behavior also evident from the reduction in the prevalence of cubic phase after the tetragonal-cubic phase transition.
volume of unit cell and crystallite size, as depicted in Fig. 7b. This is in Because the relative densities did not change until 1100 ◦ C, the observed
good agreement with the trend of Sr2+ doped BaTiO3 reported in phase transformation was likely to be originated from the CSP at 225 ◦ C.

5
M. Siddiqui et al. Open Ceramics 15 (2023) 100401

Table 2
Phase structures, lattice parameters and crystallite size of the CSP and PLS samples doped by Ba(OH)2⋅8H2O and Sr(OH)2⋅8H2O, respectively, as well as the annealed
ones.
Sample code Major phase Minor phases Lattice Parameter(Å) Crystallite Size (nm)

a-axis c-axis Volume

BT BaTiO3 – 3.99466 4.03537 64.39349 171.64


CB2 BaTiO3 BaO, BaCO3 3.99392 4.03009 64.28543 82.8
CS2 BaTiO3 BaO, BaCO3 3.99255 4.02847 64.21573 64.52
PB2 BaTiO3 BaCO3, unknown 3.99263 4.03086 64.25643 129.26
PS2 BaTiO3 BaO, SrCO3 3.99188 4.03013 64.22077 126
CB5 BaTiO3 BaO, BaCO3 3.99905 4.02529 64.37402 80.31
CS5 BaTiO3 BaCO3 3.99857 4.02311 64.32364 66.03
PB5 BaTiO3 BaCO3, unknown 3.99413 4.02284 64.1765 91.85
PS5 BaTiO3 BaCO3, SrCO3, unknown 3.99387 4.02443 64.19362 93.84
CB8 BaTiO3 Ba2TiO4 3.99893 4.024600 64.35905 94.65
CS8 BaTiO3 Ba2TiO4, BaCO3 3.9993 4.02264 64.33975 69.01
PB8 BaTiO3 BaCO3, unknown 3.99522 4.02346 64.22153 79.05
PS8 BaTiO3 BaCO3, unknown 3.99432 4.02353 64.19386 54.58
CB11 BaTiO3 Ba2TiO4 3.99833 4.02378 64.32687 131.75
CS11 BaTiO3 Ba2TiO4 3.98712 4.01356 63.80415 49.71
PB11 BaTiO3 Ba2TiO4, BaCO3, unknown 3.99537 4.02184 64.20069 81
PS11 BaTiO3 Ba2TiO4, unknown 3.99167 4.02194 64.08344 37.17

Fig. 7. Lattice parameters comparison of the CSP samples for a) unit cell volume comparison between CS and CB and b) volume of unit cell and crystallite size
comparison of the CS-samples annealed at different temperatures.

homogenous distribution of the Sr2+ in the bulk samples has proved an


effective doping. The samples that are cold sintered at 225 ◦ C under 270
MPa pressure have achieved an average relative density of 80% whilst
exhibiting a reduced level of coarseness from the microstructural
morphology compared to the pristine BaTiO3 raw powder. The grains
and particles have remained ungrown and non-uniform alongside the
unchanged densities until being post-annealed at >800 ◦ C, yet such
densities have already been able to significantly influence the phase
structure compared to the situation in the conventional, pressureless
sintering. Meanwhile, the applied pressure during the cold sintering has
affected the incorporation of the Sr2+ ions and then eventually the
Fig. 8. EDX analysis of the CS2 sample: a) the SEM micrograph and b) the phases. The phases have experienced a gradual transformation from
elemental map scanned at the same location. tetragonal to cubic after being post-annealed from well below 800 ◦ C all
the way up to 1100 ◦ C. It consequently concludes that the cold sintering
Such a finding in this study regarding the phase transformation upon process has triggered the phase transformation of the Ba1-xSrxTiO3 solid
thermal treatment differed from the previous works on the cold sintering solution. Such a conclusion provides a guidance or sets a limitation for
process of BaTiO3 [10,12,35]. Here, it is believed that the dopants the selection of post-annealing temperatures in future developments of
distributed by the sintering flux may have played a crucial role in this the cold sintering process toward adoption in functional ceramics
unusual but interesting behavior of phase transformation. including ferroelectrics and piezoelectrics.

5. Conclusions Declaration of competing interest

In this work, the effects of the cold sintering process at 225 ◦ C as well The authors declare that they have no known competing financial
as the post-annealing at different temperatures in the range of interests or personal relationships that could have appeared to influence
500–1100 ◦ C on the phenomena of phase transformation and the the work reported in this paper.
microstructural evolution have been studied for the BaTiO3 ceramics
doped by sintering flux of Ba(OH)2.8 H2O and Sr(OH)2.8H2O. The

6
M. Siddiqui et al. Open Ceramics 15 (2023) 100401

Acknowledgments [14] T. Sada, K. Tsuji, A. Ndayishimiye, Z. Fan, Y. Fujioka, C.A. Randall, High
permittivity BaTiO3 and BaTiO3-polymer nanocomposites enabled by cold
sintering with a new transient chemistry: Ba (OH) 2• 8H2O, J. Eur. Ceram. Soc. 41
The authors are grateful to the JECS Trust for funding [the visit of (2021) 409–417.
Maliha Siddiqui to Microelectronics Research Unit, Faculty of Informa­ [15] M. Ben Chamekh, Z. Ben Achour, A. Thamri, R. Chtourou, E. Dhahri, O. Touayar,
tion Technology and Electrical Engineering, University of Oulu, Structural and electrical characterization of strontium doped barium titanate for
radiometric measurement, Chem. Phys. Lett. 761 (2020), 138008.
Finland] (contract number 2021285). In addition, Czech Nano Lab [16] T. Sada, Z. Fan, A. Ndayishimiye, K. Tsuji, S.H. Bang, Y. Fujioka, C.A. Randall, In
project LM2018110 funded by MEYS CR is also gratefully acknowledged situ doping of BaTiO3 and visualization of pressure solution in flux-assisted cold
for the financial support of the measurements/sample fabrication at sintering, J. Am. Ceram. Soc. 104 (2021) 96–104.
[17] C. Hérard, A. Faivre, J. Lemaître, Surface decontamination treatments of undoped
CEITEC Nano Research Infrastructure. Maliha Siddiqui also grateful to BaTiO3—part I: powder and green body properties, J. Eur. Ceram. Soc. 15 (1995)
VUT for funding this work by specific research grant project CEITEC 135–143, https://doi.org/10.1016/0955-2219(95)93059-C.
VUT-J-22-7838. Yang Bai acknowledges the Centre for Material Anal­ [18] M. Özen, M. Mertens, F. Snijkers, P. Cool, Hydrothermal synthesis and formation
mechanism of tetragonal barium titanate in a highly concentrated alkaline
ysis, University of Oulu for the use and technical support of the char­ solution, Ceram. Int. 42 (2016) 10967–10975.
acterization equipment and facilities, as well as the Infotech Institute of [19] D. Völtzke, H.-P. Abicht, J. Woltersdorf, E. Pippel, Surface modification of pre-
University of Oulu for the financial and strategic support. sintered BaTiO3 particles, Mater. Chem. Phys. 73 (2002) 274–280.
[20] M.I. Mendelson, Average grain size in polycrystalline ceramics, J. Am. Ceram. Soc.
52 (1969) 443–446.
References [21] M.M. Lencka, R.E. Riman, Thermodynamic modeling of hydrothermal synthesis of
ceramic powders, Chem. Mater. 5 (1993) 61–70.
[1] M. Singh, B.C. Yadav, A. Ranjan, M. Kaur, S.K. Gupta, Synthesis and [22] G.M. Habashy, G.A. Kolta, Thermal decomposition of the hydrates of barium
characterization of perovskite barium titanate thin film and its application as LPG hydroxide, J. Inorg. Nucl. Chem. 34 (1972) 57–67.
sensor, Sensor. Actuator. B Chem. 241 (2017) 1170–1178, https://doi.org/ [23] G.L. Haag, Application of the Carbon Dioxide-Barium Hydroxide Hydrate Gas-Solid
10.1016/j.snb.2016.10.018. Reaction for the Treatment of Dilute Carbon Dioxide-Bearing Gas Streams, Oak
[2] H. Kishi, Y. Mizuno, H. Chazono, Base-metal electrode-multilayer ceramic Ridge National Lab., TN (USA), 1983.
capacitors: past, present and future perspectives, Jpn. J. Appl. Phys. 42 (2003) 1, [24] I. Arvanitidis, D. Siche, S. Seetharaman, A study of the thermal decomposition of
https://doi.org/10.1143/JJAP.42.1. BaCO 3, Metall. Mater. Trans. B 27 (1996) 409–416.
[3] J.F. Scott, High-dielectric constant thin films for dynamic random access memories [25] H. Jiang, J. Qi, D. Wu, W. Lu, J. Qian, H. Qu, Y. Zhang, P. Liu, X. Liu, L. Chen,
(DRAM), Annu. Rev. Mater. Sci. 28 (1998) 79–100. Atomic-resolution characterization on the structure of strontium doped barium
[4] G.H. Haertling, Ferroelectric ceramics: history and technology, J. Am. Ceram. Soc. titanate nanoparticles, Nano Res. 14 (2021) 4802–4807.
82 (1999) 797–818. [26] G. Dai, S. Wang, G. Huang, G. Chen, B. Lu, D. Li, T. Tao, Y. Yao, B. Liang, S. Lu,
[5] Y. Feng, J. Wu, Q. Chi, W. Li, Y. Yu, W. Fei, Defects and aliovalent doping Direct and indirect measurement of large electrocaloric effect in barium strontium
engineering in electroceramics, Chem. Rev. 120 (2020) 1710–1787. titanate ceramics, Int. J. Appl. Ceram. Technol. 17 (2020) 1354–1361.
[6] M.N. Rahaman, Sintering of Ceramics, CRC press, 2007. [27] R. Dinescu, M. Preda, Thermal decomposition of strontium hydroxide, J. Therm.
[7] H. Kähäri, M. Teirikangas, J. Juuti, H. Jantunen, Improvements and modifications Anal. 5 (1973) 465–473.
to room-temperature fabrication method for dielectric Li 2 MoO 4 ceramics, J. Am. [28] M.K. Mondal, M. Lenka, Solubility of CO2 in aqueous strontium hydroxide, Fluid
Ceram. Soc. 98 (2015) 687–689. Phase Equil. 336 (2012) 59–62.
[8] M. Nelo, J. Peräntie, T. Siponkoski, J. Juuti, H. Jantunen, Upside-down composites: [29] C.A. Randall, P. Yousefian, Fundamentals and practical dielectric implications of
electroceramics without sintering, Appl. Mater. Today 15 (2019) 83–86. stoichiometry and chemical design in a high-performance ferroelectric oxide:
[9] R.E. Riman, V. Atakan, Method of hydrothermal liquid phase sintering of ceramic BaTiO3, J. Eur. Ceram. Soc. 42 (2022) 1445–1473.
materials and products derived therefrom, US Patent 8 (313) (2012) 802. [30] D.-H. Yoon, B.I. Lee, P. Badheka, X. Wang, Barium ion leaching from barium
[10] H. Guo, A. Baker, J. Guo, C.A. Randall, Cold sintering process: a novel technique titanate powder in water, J. Mater. Sci. Mater. Electron. 14 (2003) 165–169.
for low-temperature ceramic processing of ferroelectrics, J. Am. Ceram. Soc. 99 [31] S. Lee, C.A. Randall, Z. Liu, Modified phase diagram for the barium oxide–titanium
(2016) 3489–3507. dioxide system for the ferroelectric barium titanate, J. Am. Ceram. Soc. 90 (2007)
[11] M. Biesuz, S. Grasso, V.M. Sglavo, What’s new in ceramics sintering? A short report 2589–2594.
on the latest trends and future prospects, Curr. Opin. Solid State Mater. Sci. 24 [32] K.I. Othman, A.A. Hassan, O.A.A. Abdelal, E.S. Elshazly, M.E.-S. Ali, S.M. El-Raghy,
(2020), 100868. S. El-Houte, Formation mechanism of barium titanate by solid-state reactions, Int.
[12] H. Guo, J. Guo, A. Baker, C.A. Randall, Hydrothermal-assisted cold sintering J. Sci. Eng. Res. 5 (2014) 1460–1465.
process: a new guidance for low-temperature ceramic sintering, ACS Appl. Mater. [33] B. Chen, F.-H. Liao, H. Jiao, X.-P. Jing, Thermal stability and phase transformation
Interfaces 8 (2016) 20909–20915. of barium orthotitanate (Ba2TiO4), Phase Transitions 86 (2013) 380–390.
[13] K. Tsuji, A. Ndayishimiye, S. Lowum, R. Floyd, K. Wang, M. Wetherington, J.- [34] M.H. Frey, D.A. Payne, Grain-size effect on structure and phase transformations for
P. Maria, C.A. Randall, Single step densification of high permittivity BaTiO3 barium titanate, Phys. Rev. B 54 (1996) 3158.
ceramics at 300 oC, J. Eur. Ceram. Soc. 40 (2020) 1280–1284, https://doi.org/ [35] H. Guo, A. Baker, J. Guo, C.A. Randall, Protocol for ultralow-temperature ceramic
10.1016/j.jeurceramsoc.2019.12.022. sintering: an integration of nanotechnology and the cold sintering process, ACS
Nano 10 (2016) 10606–10614.

You might also like