Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Ceramics International 47 (2021) 21618–21624

Contents lists available at ScienceDirect

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

Microstructure evaluation of 3YSZ sintered by Two-Step Flash Sintering


Isabela R. Lavagnini a, *, João V. Campos a, Eliria M.J.A. Pallone a, b
a
Postgraduate Programme in Materials Science and Engineering, University of São Paulo, USP/FZEA, Av. Duque de Caxias Norte, 225, 13635-900, Pirassununga, SP,
Brazil
b
Department of Biosystem Engineering, Faculty of Animal Science and Food Engineering (FZEA), University of São Paulo (USP), 13635-900, Pirassununga, SP, Brazil

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

Keywords: Flash Sintering (FS) is an energy-efficient sintering technique, which allows a considerable reduction in pro­
Zirconia cessing time and temperature. However, a recurring problem of flash sintered samples is the microstructural
Yttria-stabilized zirconia heterogeneity, which occurs mostly on samples with a high surface-to-volume ratio (e.g., cylindrical samples).
Flash sintering
Thus, aiming to obtain homogeneous microstructures and smaller grain sizes, this work evaluated the application
Two step sintering
of Two-Step Flash Sintering (TSFS) in 3YSZ and compared them with samples sintered by FS. The TSFS exper­
iments consisted of applying a first step of current density with greater magnitude (100 mA mm-2 and 200 mA
mm-2) followed by the second step with smaller magnitude (50 mA mm-2 and 75 mA mm-2). In the FS experi­
ments, a single step of current density was used (100 mA mm-2 and 200 mA mm-2). The Results showed that TSFS
significantly reduced the microstructural heterogeneity compared to FS, promoting smaller and more homoge­
neous grain sizes in different regions of the samples.

1. Introduction energy savings when compared to conventional sintering [15,17,18].


Despite these advantages, some studies have reported microstruc­
Refined microstructures (smaller grain sizes) result in properties tural heterogeneity within flash sintered materials, i.e., regions with
changes in material, such as increased fracture toughness, increased different grain sizes in the same sample. Some studies presented het­
hardness, among others [1–6]. The microstructural control of ceramic erogeneities related to the interior and surface of the sample [19–23].
materials, and, consequently, of their final properties, can be performed Others reported heterogeneities between the regions close to the elec­
during the material processing stages [6–9]. Among these stages, sin­ trodes (cathode and anode) [24–26]. Anyway, this is a challenge that
tering is one of the most critical in obtaining a dense material with a must be overcome to achieve better properties of the final material.
refined microstructure, due to the inherent grain growth of the process Thus, several studies are looking for ways to improve the FS to obtain
[10–12]. Thus, several unconventional sintering techniques have been materials with homogeneous microstructure [19,21,27–33].
studied to obtain materials with high densification and refined micro­ Steil et al. [19] presented a variation of FS, the Double Flash Sin­
structure [13–16]. tering. This technique consists of applying two steps of current density:
Flash Sintering (FS) is an unconventional sintering technique pre­ the first one for a short time; and then, after an interval without current
sented for the first time in 2010 [15]. This technique consists of applying density, a second step for a longer time. With this methodology, the
an electric field in a ceramic sample, concomitant to the furnace heating authors sintered 8YSZ to full density. Nie et al. [27] proposed another
up. At a critical temperature, the electric current flowing through the variation of FS, the Two-Step Flash Sintering (TSFS). In this variation,
sample rises abruptly due to the thermal runaway (cycle of increase in two steps of electric current density were applied: one of greater
temperature and electrical conductivity caused by the Joule heating in magnitude (for a few seconds), only to activate the densification
materials with negative temperature coefficient), leading to rapid mechanisms; and another one of lower magnitude to densify the mate­
densification of the sample. The furnace temperature in which the rial without too much grain growth. Nie et al. sintered ZnO by TSFS and
thermal runaway occurs is usually much lower than the temperature obtained smaller grain sizes (up to three times) compared to those sin­
used in conventional sintering [15]. Thus, FS increases the processing tered by FS.
speed, reduces the processing temperature, and consequently promotes Thus, the present work aims to evaluate the effect of two steps of

* Corresponding author.
E-mail address: isabela.lavagnini@usp.br (I.R. Lavagnini).

https://doi.org/10.1016/j.ceramint.2021.04.174
Received 29 March 2021; Received in revised form 15 April 2021; Accepted 18 April 2021
Available online 21 April 2021
0272-8842/© 2021 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
I.R. Lavagnini et al. Ceramics International 47 (2021) 21618–21624

electric current density (TSFS) on the densification and microstructure sake of simplicity, we did not take the capacitive response of the sample
of 3YSZ using an alternate electric field. into account in the BBR model. Therefore, assuming that I and V were in
phase, we used their RMS values in equation (2).
2. Material and methods Cylindrical samples used here presented larger dimensions when
compared to dog-bone samples. Thus, in our experiments, the heat
The commercial 3 mol% yttria-stabilized zirconia powder from exchanged between the sample flat surfaces and the electrodes cannot be
Tosoh (TZ-3Y-E, particle size 40 nm, 6.05 g cm− 3 theoretical density) neglected at the BBR model. To consider the electrodes as heat sinks, the
was prepared as described in a previous work [21]. The cylindrical area (A), used in equation (2), was the total surface area of the sample
samples (6 mm in diameter and 5 mm in height) were formed using (radial + flat surfaces). With that, we assumed that conduction heat flux
uniaxial pressing followed by isostatic pressing at 200 MPa, then between the flat surfaces and the electrodes was similar to radiation heat
calcined at 600◦ C for 1 h using a heating rate of 5◦ C.min-1. flux between the furnace and samples radial surfaces.
The sintering was carried out in an automated vertical tubular After sintering, the samples were cut in their radial centres in the
furnace [26] using platinum disks as electrodes. Those disks were placed longitudinal direction. The cut surfaces were polished and thermally
on the flat surfaces of the cylindrical sample (top and bottom). A uni­ etched at 1450◦ C for 10 min with a heating rate of 10◦ C.min-1. The
axial mechanical pressure (3 bar) was applied to maintain contact be­ microstructural analysis was performed using a scanning electron mi­
tween the sample and the electrodes since no conductive paste was used. croscope (SEM, FEG-XL 30). Different regions of the cut surface were
An alternate electric field of 120 V cm-1 (RMS), frequency of 1000 Hz, analysed: core, radial surface (located at half height of the cylindrical
was applied since the beginning of the furnace heating (heating rate of samples), and the flat surface (region close to the electrodes). Fig. 1
20◦ C.min-1). A displacement sensor (LVDT type) recorded the specific shows the longitudinal radial section of the cylindrical sample and the
linear shrinkage of the samples in situ during the sintering procedure. regions where the micrographs were obtained from the polished
For the TSFS experiments, the first step of a higher electric current surfaces.
density (100 and 200 mA mm-2) was applied for 5 s. Then, in the second The ImageJ software (National Institutes of Health) was used to
step, a lower current density (75 and 50 mA mm-2) was applied for 55 s. calculate the grain size in each analysed region. The homogeneity
The experiments in FS were performed with the maximum current evaluation, according to the average grain sizes in each region of the
density limits (100 and 200 mA mm-2) applied for 60 s. All the current same sample, was made by analysis of variance according to the Tukey
densities described above were RMS values. The different conditions test, with a significance level of 5%.
studied are represented and labelled in Table 1. J1, J2, Δt1 and Δt2 are,
respectively, the electric current densities of the first and second steps 3. Results and discussion
and the time of those steps. Each condition was performed in triplicate.
The apparent densities of the samples were determined using the Fig. 2 shows the curves of electric current density (J), electric field
Archimedes’ principle according to ASTM C373-8826; the averages of (E), power density (P), and linear shrinkage (dL/Lo) of all studied con­
the relative densities were expressed as a function of the theoretical ditions. It is observed that the curves presented are in accordance with
density of 3YSZ. The Black Body Radiation model (BBR) was used to the standard curves of the FS, in which the increase in the electric cur­
estimate the temperature on the samples surfaces during the first and rent density provides a drop in the value of the applied electric field as
second step, using equation (1) [34]. The samples surface temperatures well as the instantaneous shrinkage of the material [15]. For TSFS
for each step were calculated by the average temperature of the last 3 s samples, the current density curves presented two steps regarding the
for each step. magnitude of the current. The power density curves showed a high
( ) power peak in the samples sintered by FS and only in the first step of the
dTs
mCp = VI − σεA Ts4 − Tf4 (1) samples sintered by TSFS. The power density is related to the samples’
dt
temperature.
In equation (1) m is the sample mass (kg), Cp is the sample heat ca­ Fig. 2 also shows that the behaviour of the electric field was different
pacity (600 J kg-1 K-1 [35]), Ts is the sample temperature (K), t is time between the experiments. It is noted that the electric field of the samples
(s), V is the voltage (V), I is the current (A), σ is Stefan’s Boltzmann in TSFS presented two well-defined steps. In the case of TSFS200/75 and
constant (5.67 10-8 W m-2 K-4), ε is the sample emissivity (assumed 0.8 TSFS200/50, the transition between the first and second step reached
[36]), A is the sample surface area (m2), and Tf is the furnace temper­ lower values of electric field than the transitions of TSFS100/75 and
ature (K). The numerical solution of equation (1) is presented in terms of TSFS100/50. This behaviour indicates that the samples’ conductivity
iterations (i) in equation (2). Worth noting that the acquisition rate of changes during the steps transitions due to change in the samples’
the setup was three points per second. temperature. This effect was more evident in TSFS200/50 because there
was a greater difference between the current density of the steps.
t(i + 1) − t(i) [ ( )]
Ts (i + 1) = Ts (i) + V(i)I(i) − σεA(i) Ts (i)4 − Tf (i)4 (2) However, despite the curves of current density, electric field, and
mCp
power density of the samples in TSFS present two well-defined steps, this
Carvalho et al. [37] reported that 3YSZ samples under alternate is not observed in the linear shrinkage curves. This is because most of the
electric field with 1000 Hz presented similar linear shrinkage as samples linear shrinkage of the sample takes place at the first step, regardless of
under a direct electric field. It indicates that the capacitive effects have a
minor influence on the I and V when using this frequency. Thus, for the

Table 1
Experimental conditions of TSFS and FS using different step times and maximum
electric current densities.
Condition Label J1(mA.mm-2) Δt1 (s) J2 (mA.mm-2) Δt2 (s)

FS100 100 60 – –
TSFS100/50 100 5 50 55
TSFS100/75 100 5 75 55
FS200 200 60 – –
TSFS200/50 200 5 50 55
Fig. 1. Radial longitudinal section of a cylindrical sample and its polished
TSFS200/75 200 5 75 55
surface, with emphasis on regions where the microstructure was analysed.

21619
I.R. Lavagnini et al. Ceramics International 47 (2021) 21618–21624

Fig. 2. Electric current density (J), electric field (E), power density (P), and linear shrinkage (dL/Lo) curves as a function of time for samples during FS and TSFS.

the magnitude of the electric current density applied, in the same way as
Table 2
the shrinkage of the samples in FS. However, it is observed that the
Power density peak, average power density, and temperature estimated by BBR
samples in TSFS that received 75 mA mm-2 (regardless of the current
for samples during FS and TSFS followed by the standard deviation for all
density limit applied in the first step) showed higher values of linear conditions studied.
shrinkage, indicating that there was some shrinkage at the second step
Condition Power Peak Average Power (mW. Temperature BBR (◦ C)
as well.
Label (mW.mm-3) mm-3)
Table 2 shows the power density peak, the average power density,
1◦ step 1◦ step 2◦ step 1◦ step 2◦ step
and the calculated temperature using the BBR model for FS and for each
step of TSFS samples. It is observed that the sample temperature is FS100 1469.8 ± 528.4 ± – 1662.7 ± –
proportional to the current density applied, i.e., higher current densities 25.3 52.2 14.1
TSFS100/50 1325.5 649.1 ± 261.1 ± 1731.9 1416.4 ±
lead to higher temperatures [36]. Worth noting that the temperatures
± ±
53.0 16.5 5.4 12.2 10.3
estimated by BBR for FS and the first steps of TSFS samples using the TSFS100/75 1394.4 ± 643.7 ± 406.4 ± 1737.3 ± 1558.9 ±
same current density were slightly different. It happened because the 67.8 17.0 5.1 9.6 9.7
first steps of TSFS were too short (5 s). In this short time, the samples FS200 2782.5 ± 1189.3 ± – 2035.0 ± –
224.4 49.7 15.2
were still densifying. Thus, their electrical resistivity was still changing
TSFS200/50 2714.2 ± 1431.7 ± 242.8 ± 2087.4 ± 1401.1 ±
and consequently their temperature. In FS, the temperature estimated by 179.4 45.5 19.9 13.6 11.7
BBR was lower because the step was long enough (60 s) to densify and TSFS200/75 2830.5 ± 1455.7 ± 395.5 ± 2111.8 ± 1560.3 ±
stabilize the samples’ electrical resistivity. 201.7 48.3 21.7 15.7 8.1
The temperatures showed in Table 2 suggest that the samples
TSFS200/75, TSFS200/50, and FS200 reached temperatures higher than the
platinum melting point. However, no evidence of molten platinum was
observed on the electrodes, suggesting that the BBR model

21620
I.R. Lavagnini et al. Ceramics International 47 (2021) 21618–21624

overestimated the samples’ temperatures, up to 350◦ C as already re­ samples presented the same size [20].
ported by Charalambous et al. [38]. Fig. 3 shows the micrographs in different regions (core, radial sur­
The furnace temperature at the moment of the flash was 995◦ C face, and flat surface) of the samples under different conditions in FS and
(±5◦ C) for all the studied conditions. The variation in the onset tem­ TSFS. Table 3 shows the relative densities and the average grain size
perature of flash between all the conditions was slight because the same measured for these regions. It is observed that the samples in FS showed
electric field limit (120 V cm-1) [39] was used in all cases, and all the larger grains in its core than in the other regions, regardless of the

Fig. 3. SEM micrographs of the different regions analysed for each condition studied.

21621
I.R. Lavagnini et al. Ceramics International 47 (2021) 21618–21624

current density limit used. Also, the sintered samples with a current should lead to higher densification. Although, as observed in Fig. 3, the
density limit of 200 mA mm-2 showed larger grains for all regions, samples submitted to 200 mA mm-2 presented higher coarsening and
accompanied by the presence of pores. Dong and Chen [40] also cavitation, leading to lower densification, as reported previously [21].
observed the presence of pores at flash sintered 8YSZ in severe condi­ Besides that, in Fig. 1, the final linear shrinkage of the samples was
tions of current density. These pores were attributed to the cavitation higher for conditions using 200 mA mm-2. However, the relative den­
effect, due to reduction-generated voids by condensation of supersatu­ sities showed the opposite behaviour. It can be happening because the
rated oxygen vacancies [21,40]. linear shrinkage was measured only on the vertical axis (height of the
The samples in TSFS with a limit of 100 mA mm-2 presented grains cylinders). Thus, considering the superplasticity effect during flash
with similar sizes for all analysed regions. In contrast, samples with a presented by the material [42–44], it is expected that the samples may
limit of 200 mA mm-2, in the first step, showed much bigger grains in have deformed during these severe current density conditions. There­
their core than the other regions (radial and flat surfaces). Regarding the fore, the linear shrinkage measured by the LVDT took into account the
current densities applied at the second step (50 or 75 mA mm-2), no shrinkage (caused by sintering) and deformation of the sample (caused
significant differences were observed. by the 3-bar mechanical pressure).
The exaggerated growth in grain size inside samples in FS and TSFS, Fig. 4 shows the evaluation of microstructural homogeneity for all
when using a current density limit of 200 mA mm-2, happens due to the conditions studied according to the Tukey’s mean difference. The
thermal gradient formed in cylindrical samples during FS. This thermal comparison between two regions of the same sample shows an average
gradient is caused by the difference between the heat gain rate (due to difference value, and how close to zero this value indicates if the regions
Joule heating) and the heat loss rate of the material [21,33,40,41]. In have statistically similar grain sizes.
the case of TSFS samples with a limit of 100 mA mm-2, it was possible to In Fig. 4, TSFS100/75 and TSFS100/50 presented a homogeneous
control the thermal gradient. microstructure for all regions analysed. However, FS100 presented ho­
Regarding the relative density of the samples, it can be seen in mogeneity only between the radial and flat surfaces. All samples that
Table 3 that the samples with a current density limit of 100 mA mm-2 received a current density limit of 200 mA mm-2, both in FS and TSFS,
showed higher densities than the samples with 200 mA mm-2. This presented heterogeneous microstructures in all regions. Although, in the
behaviour was counterintuitive if one thinks that higher temperature case of TSFS samples, the heterogeneity was slighter than those sintered

Table 3
Relative density and average grain size in the core, radial surface, and flat surface followed by the standard deviation for all conditions.
Condition Relative Densities (%TD) Average grain size (μm)

Core Radial Surface Flat Surface

FS100 92.1 ± 1.4 1.1 ± 0.2 0.5 ± 0.2 0.4 ± 0.2


TSFS100/50 92.9 ± 0.6 0.6 ± 0.2 0.5 ± 0.1 0.4 ± 0.2
TSFS100/75 93.9 ± 0.5 0.6 ± 0.3 0.5 ± 0.2 0.4 ± 0.2
FS200 90.9 ± 1.3 7.1 ± 1.8 0.8 ± 0.4 0.5 ± 0.2
TSFS200/50 90.1 ± 0.2 3.8 ± 1.6 0.6 ± 0.3 0.5 ± 0.2
TSFS200/75 89.3 ± 0.6 6.1 ± 1.8 0.6 ± 0.2 0.5 ± 0.2

Fig. 4. Tukey’s mean difference between the average grain size of each region was analysed for the same condition.

21622
I.R. Lavagnini et al. Ceramics International 47 (2021) 21618–21624

by FS, indicating that the second step minimizes this heterogeneity. [14] Y. Chen, B. Fan, B. Yang, W. Ma, G. Liu, H. Li, Microwave sintering and fracture
behavior of zirconia ceramics, Ceram. Int. 45 (2019) 17675–17680, https://doi.
org/10.1016/j.ceramint.2019.05.334.
4. Conclusions [15] M. Cologna, B. Rashkova, R. Raj, Flash sintering of nanograin zirconia in <5 s at
850◦ C, J. Am. Ceram. Soc. 93 (2010) 3556–3559, https://doi.org/10.1111/j.1551-
2916.2010.04089.x.
The application of two steps of different electric current density
[16] M. Biesuz, R. Sedlák, T. Saunders, A. Kovalčíková, J. Dusza, M. Reece, D. Zhu,
promoted greater microstructural control and homogeneity in the grain C. Hu, S. Grasso, Flash spark plasma sintering of 3YSZ, J. Eur. Ceram. Soc. 39
size of 3YSZ cylindrical samples, showing the potential of the TSFS (2019) 1932–1937, https://doi.org/10.1016/j.jeurceramsoc.2019.01.017.
[17] C.E.J. Dancer, Flash sintering of ceramic materials, Mater. Res. Express 3 (2016)
technique to improve the results of FS even more.
102001, https://doi.org/10.1088/2053-1591/3/10/102001.
However, the current density limit is an important parameter, as it [18] F. Trombin, R. Raj, Developing processing maps for implementing flash sintering
has a great influence on the final microstructure. The samples sintered in into manufacture of whiteware ceramics, Am. Ceram. Soc. Bull. 93 (2014) 32–35.
TSFS with 200 mA mm-2 continued to exhibit microstructural hetero­ [19] M.C. Steil, D. Marinha, Y. Aman, J.R.C. Gomes, M. Kleitz, From conventional ac
flash-sintering of YSZ to hyper-flash and double flash, J. Eur. Ceram. Soc. 33
geneity between the regions studied, in addition to greater grain growth. (2013) 2093–2101, https://doi.org/10.1016/j.jeurceramsoc.2013.03.019.
[20] J.V. Campos, I.R. Lavagnini, J.G. Pereira da Silva, J.A. Ferreira, R.V. Sousa,
R. Mücke, O. Guillon, E.M. Pallone, Flash sintering scaling-up challenges: influence
Declaration of competing interest of the sample size on the microstructure and onset temperature of the flash event,
Scripta Mater. 186 (2020) 1–5, https://doi.org/10.1016/j.scriptamat.2020.04.022.
The authors declare that they have no known competing financial [21] I.R. Lavagnini, J.V. Campos, J.A. Ferreira, E.M.J.A. Pallone, Microstructural
evolution of 3YSZ flash-sintered with current ramp control, J. Am. Ceram. Soc.
interests or personal relationships that could have appeared to influence (2020) 1–7, https://doi.org/10.1111/jace.17037.
the work reported in this paper. [22] Y. Du, A.J. Stevenson, D. Vernat, M. Diaz, D. Marinha, Estimating Joule heating
and ionic conductivity during flash sintering of 8YSZ, J. Eur. Ceram. Soc. 36 (2016)
749–759, https://doi.org/10.1016/j.jeurceramsoc.2015.10.037.
Acknowledgments [23] S. Carvalho, E. Muccillo, R. Muccillo, Electrical behavior and microstructural
features of electric field-assisted and conventionally sintered 3 mol% yttria-
This work was supported by the Coordenação de Aperfeiçoamento de stabilized zirconia, Ceramics 1 (2018) 1–10, https://doi.org/10.3390/
ceramics1010002.
Pessoal de Nível Superior – Brasil (CAPES) – Finance Code 001 and the [24] K. Ren, J. Xia, Y. Wang, Grain growth kinetics of 3 mol. % yttria-stabilized zirconia
São Paulo Research Foundation (FAPESP) [2015/07319-8, and 2018/ during flash sintering, J. Eur. Ceram. Soc. 39 (2019) 1366–1373, https://doi.org/
04331-5]. 10.1016/j.jeurceramsoc.2018.11.032.
[25] M. Biesuz, L. Pinter, T. Saunders, M. Reece, J. Binner, V.M. Sglavo, S. Grasso,
Investigation of electrochemical, optical and thermal effects during flash sintering
References of 8YSZ, Materials 11 (2018), https://doi.org/10.3390/ma11071214.
[26] J.V. Campos, I.R. Lavagnini, R.V. Sousa, J.A. Ferreira, E.M.J.A Pallone,
[1] H.L. Yao, C. Yang, Q. Yang, X.Z. Hu, M.X. Zhang, X.B. Bai, H.T. Wang, Q.Y. Chen, Development of an instrumented and automated flash sintering setup for enhanced
Structure, mechanical and bioactive properties of nanostructured hydroxyapatite/ process monitoring and parameter control, J. Eur. Ceram. Soc. 39 (2019) 531–538,
titania composites prepared by microwave sintering, Mater. Chem. Phys. 241 https://doi.org/10.1016/j.jeurceramsoc.2018.09.002.
(2020) 122340, https://doi.org/10.1016/j.matchemphys.2019.122340. [27] J. Nie, Y. Zhang, J.M. Chan, S. Jiang, R. Huang, J. Luo, Two-step flash sintering of
[2] S. Balaji, B.K. Mandal, S. Ranjan, N. Dasgupta, R. Chidambaram, Nano-zirconia – ZnO: fast densification with suppressed grain growth, Scripta Mater. 141 (2017)
evaluation of its antioxidant and anticancer activity, J. Photochem. Photobiol. B 6–9, https://doi.org/10.1016/j.scriptamat.2017.07.015.
Biol. 170 (2017) 125–133, https://doi.org/10.1016/j.jphotobiol.2017.04.004. [28] M. Biesuz, J. Dong, S. Fu, Y. Liu, H. Zhang, D. Zhu, C. Hu, S. Grasso, Thermally-
[3] A. Kusior, J. Banas, A. Trenczek-Zajac, P. Zubrzycka, A. Micek-Ilnicka, M. Radecka, insulated flash sintering, Scripta Mater. 162 (2019) 99–102, https://doi.org/
Structural properties of TiO2 nanomaterials, J. Mol. Struct. 1157 (2018) 327–336, 10.1016/j.scriptamat.2018.10.042.
https://doi.org/10.1016/j.molstruc.2017.12.064. [29] A.G. Storion, J.A. Ferreira, S.C. Maestrelli, E.M.J.A. Pallone, Influence of the
[4] Z. Yin, J. Yuan, Z. Wang, H. Hu, Y. Cheng, X. Hu, Preparation and properties of an forming method on flash sintering of ZnO ceramics, Ceram. Int. (2020), https://
Al2O3/Ti(C,N) micro-nano-composite ceramic tool material by microwave doi.org/10.1016/j.ceramint.2020.08.210.
sintering, Ceram. Int. 42 (2016) 4099–4106, https://doi.org/10.1016/j. [30] R.R. Ingraci Neto, K.J. McClellan, D.D. Byler, E. Kardoulaki, Controlled current-
ceramint.2015.11.082. rate AC flash sintering of uranium dioxide, J. Nucl. Mater. 547 (2021) 152780,
[5] P. fei Liu, Z. Li, P. Xiao, H. Luo, T. hui Jiang, Microstructure and mechanical https://doi.org/10.1016/j.jnucmat.2021.152780.
properties of in-situ grown mullite toughened 3Y-TZP zirconia ceramics fabricated [31] G.M. Jones, M. Biesuz, W. Ji, S.F. John, C. Grimley, C. Manière, C.E.J. Dancer,
by gelcasting, Ceram. Int. 44 (2018) 1394–1403, https://doi.org/10.1016/j. Promoting microstructural homogeneity during flash sintering of ceramics through
ceramint.2017.09.151. thermal management, MRS Bull. 46 (2021) 59–66, https://doi.org/10.1557/
[6] R.H.L. Garcia, V. Ussui, N.B. de Lima, E.N.S. Muccillo, D.R.R. Lazar, Physical s43577-020-00010-2.
properties of alumina/yttria-stabilized zirconia composites with improved [32] T.P. Mishra, R.R.I. Neto, R. Raj, O. Guillon, M. Bram, Current-rate flash sintering of
microstructure, J. Alloys Compd. 486 (2009) 747–753, https://doi.org/10.1016/j. gadolinium doped ceria: microstructure and defect generation, Acta Mater. 189
jallcom.2009.06.204. (2020) 145–153, https://doi.org/10.1016/j.actamat.2020.02.036.
[7] P.J. Thomas, P. O’Brien, Nanomaterials Chemistry, WILEY-VCH Verlag GmbH & [33] C.A. Grimley, S. Funni, C. Green, E.C. Dickey, A thermal perspective of flash
Co, Weinheim, 2007, https://doi.org/10.1002/9783527611362. sintering: the effect of AC current ramp rate on microstructure evolution, J. Eur.
[8] A. Shafeiey, M.H. Enayati, A. Al-Haji, The effect of slip casting parameters on the Ceram. Soc. (2020), https://doi.org/10.1016/j.jeurceramsoc.2020.11.040.
green density of MgAl2O4 spinel, Ceram. Int. 43 (2017) 6069–6074, https://doi. [34] M.O. Prado, M. Biesuz, M. Frasnelli, F.E. Benedetto, V.M. Sglavo, Viscous flow flash
org/10.1016/j.ceramint.2017.01.151. sintering of porous silica glass, J. Non-Cryst. Solids 476 (2017) 60–66, https://doi.
[9] F.A.T. Guimarães, K.L. Silva, V. Trombini, J.J. Pierri, J.A. Rodrigues, R. Tomasi, E. org/10.1016/j.jnoncrysol.2017.09.024.
M.J.A. Pallone, Correlation between microstructure and mechanical properties of [35] W. Ji, B. Parker, S. Falco, J.Y. Zhang, Z.Y. Fu, R.I. Todd, Ultra-fast firing: effect of
Al2O3/ZrO2 nanocomposites, Ceram. Int. 35 (2009) 741–745, https://doi.org/ heating rate on sintering of 3YSZ, with and without an electric field, J. Eur. Ceram.
10.1016/j.ceramint.2008.02.002. Soc. 37 (2017) 2547–2551, https://doi.org/10.1016/j.jeurceramsoc.2017.01.033.
[10] D.S.B. Heidary, M. Lanagan, C.A. Randall, Contrasting energy efficiency in various [36] R. Raj, Joule heating during flash-sintering, J. Eur. Ceram. Soc. 32 (2012)
ceramic sintering processes, J. Eur. Ceram. Soc. 38 (2018) 1018–1029, https://doi. 2293–2301, https://doi.org/10.1016/j.jeurceramsoc.2012.02.030.
org/10.1016/j.jeurceramsoc.2017.10.015. [37] S.G.M. Carvalho, E.N.S. Muccillo, R. Muccillo, AC electric field assisted
[11] M.N. Rahaman, Ceramic Processing and Sintering, 1996, https://doi.org/10.1179/ pressureless sintering zirconia: 3 mol% yttria solid electrolyte, Phys. Status Solidi
095066096790151286. Appl. Mater. Sci. 215 (2018) 1–5, https://doi.org/10.1002/pssa.201700647.
[12] N.J. Lóh, L. Simão, C.A. Faller, A. De Noni, O.R.K. Montedo, A review of two-step [38] H. Charalambous, S.K. Jha, R.T. Lay, A. Cabales, J. Okasinski, T. Tsakalakos,
sintering for ceramics, Ceram. Int. 42 (2016) 12556–12572, https://doi.org/ Investigation of temperature approximation methods during flash sintering of ZnO,
10.1016/j.ceramint.2016.05.065. Ceram. Int. 44 (2018) 6162–6169, https://doi.org/10.1016/j.
[13] A. Fregeac, F. Ansart, S. Selezneff, C. Estournès, Relationship between mechanical ceramint.2017.12.250.
properties and microstructure of yttria stabilized zirconia ceramics densified by [39] J.S.C. Francis, R. Raj, Influence of the field and the current limit on flash sintering
spark plasma sintering, Ceram. Int. 45 (2019) 23740–23749, https://doi.org/ at isothermal furnace temperatures, J. Am. Ceram. Soc. 96 (2013) 2754–2758,
10.1016/j.ceramint.2019.08.090. https://doi.org/10.1111/jace.12472.

21623
I.R. Lavagnini et al. Ceramics International 47 (2021) 21618–21624

[40] Y. Dong, I.W. Chen, Electrical and hydrogen reduction enhances kinetics in doped [43] R. Raj, M. Cologna, J.S.C. Francis, Influence of externally imposed and internally
zirconia and ceria: II. Mapping electrode polarization and vacancy condensation in generated electrical fields on grain growth, diffusional creep, sintering and related
YSZ, J. Am. Ceram. Soc. 101 (2018) 1058–1073, https://doi.org/10.1111/ phenomena in ceramics, J. Am. Ceram. Soc. 94 (2011) 1941–1965, https://doi.
jace.15274. org/10.1111/j.1551-2916.2011.04652.x.
[41] J. Park, I.W. Chen, In situ thermometry measuring temperature flashes exceeding [44] J. Cho, Q. Li, H. Wang, Z. Fan, J. Li, S. Xue, K.S.N. Vikrant, H. Wang, T.B. Holland,
1,700◦ C in 8 mol% Y2O3-stablized zirconia under constant-voltage heating, J. Am. A.K. Mukherjee, R.E. García, X. Zhang, High temperature deformability of ductile
Ceram. Soc. 96 (2013) 697–700, https://doi.org/10.1111/jace.12176. flash-sintered ceramics via in-situ compression, Nat. Commun. 9 (2018) 1–9,
[42] J.S.C. Francis, R. Raj, Flash-sinterforging of nanograin zirconia: field assisted https://doi.org/10.1038/s41467-018-04333-2.
sintering and superplasticity, J. Am. Ceram. Soc. 95 (2012) 138–146, https://doi.
org/10.1111/j.1551-2916.2011.04855.x.

21624

You might also like