Sunteți pe pagina 1din 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/302976805

A review of two-step sintering for ceramics

Article  in  Ceramics International · May 2016


DOI: 10.1016/j.ceramint.2016.05.065

CITATIONS READS

15 1,180

5 authors, including:

Nayadie Jorge Lóh Lisandro Simão


Universidade do Extremo Sul Catarinense (UNESC) Federal University of Santa Catarina
9 PUBLICATIONS   23 CITATIONS    22 PUBLICATIONS   33 CITATIONS   

SEE PROFILE SEE PROFILE

C. A. Faller Agenor De Noni Jr.


Universidade do Extremo Sul Catarinense (UNESC) Federal University of Santa Catarina
9 PUBLICATIONS   29 CITATIONS    57 PUBLICATIONS   261 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Ceramic tiles View project

Development of geopolymeric low energy binders from industrial wastes as coatings passively acting in the control of indoor air quality View project

All content following this page was uploaded by Oscar Rubem Klegues Montedo on 08 June 2016.

The user has requested enhancement of the downloaded file.


Ceramics International 42 (2016) 12556–12572

Contents lists available at ScienceDirect

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

Review article

A review of two-step sintering for ceramics


N.J. Lóh a,b, L. Simão a,b, C.A. Faller a, A. De Noni Jra,b, O.R.K. Montedo a,b,n
a
Grupo de Pesquisa em Cerâmica Técnica (CERTEC), Universidade do Extremo Sul Catarinense (UNESC), Av. Universitária, 1105 - P.O. Box 3167, 88806-000
Criciúma (SC), Brazil
b
Programa de Pós-Graduação em Ciência e Engenharia de Materiais (PPGCEM), Universidade do Extremo Sul Catarinense (UNESC), Av. Universitária,
1105 - P.O. Box 3167, 88806-000 Criciúma (SC), Brazil

art ic l e i nf o a b s t r a c t

Article history: The development of high-density ceramic materials with fine-grained microstructures has been studied
Received 27 January 2016 to considerably improve their properties for high-performance applications. Many alternatives have been
Received in revised form searched to refine their microstructure by changing their composition and/or processing. Among such
30 March 2016
alternatives, the densification of ceramic materials by sintering curve control is an effective, simple and
Accepted 10 May 2016
Available online 11 May 2016
economical microstructure refinement method. Thus, different thermal treatment techniques such as
spark plasma sintering and microstructural forms of control such as the control of sintering conditions
Keywords: have been used to obtain nanostructured materials. One of the techniques widely used in recent years is
Ceramic materials two-step sintering. Two-step sintering (TSS) is a promising method used to obtain high-density bodies
Two-step sintering
and smaller grain sizes. Two TSS methodologies are known: sintering with thermal pretreatment at a low
Suppression of grain growth
sintering temperature, followed by a second stage at elevated temperature, and the more recent ap-
Final microstructure
Properties proach presented by Chen and Wang, which has been the most widely used. In addition to the sintering
conditions (temperature, heating rate and sintering holding times) that must be suitable for each
composition type, the starting materials, particle size and processing method may influence the obtained
microstructure, especially the reduced grain size and increased densification. The current review of two-
step sintering presents the effect of this technique on the grain density and sizes of different ceramic
materials. The influence of the addition of doping agents and its effect on the mechanical properties in
different systems is also presented in the current study.
& 2016 Elsevier Ltd and Techna Group S.r.l. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12557
2. Historical context. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12557
3. Two-step sintering method by Chu . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12557
4. Two-step sintering method by Chen and Wang. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12559
4.1. Yttrium oxide (Y2O3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12559
4.2. Zinc oxide (ZnO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12560
4.3. Aluminum oxide (Al2O3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12560
4.4. Yttria (Y2O3)-stabilized zirconia (ZrO2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12561
4.5. Titanium dioxide (TiO2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12563
4.6. Barium titanate (BaTiO3, BTO). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12564
4.7. Carbides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12564
4.8. Corundum abrasive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12565
4.9. Piezoelectric ceramics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12566
4.10. Bioceramics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12566
5. Miscellaneous . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12567

n
Corresponding author at: Grupo de Pesquisa em Cerâmica Técnica (CERTEC), Universidade do Extremo Sul Catarinense (UNESC), Av. Universitária, 1105 - P.O. Box 3167,
88806-000 Criciúma (SC), Brazil.
E-mail addresses: nayadie@gmail.com (N.J. Lóh), lisandrosimao@gmail.com (L. Simão), cristian.faller@hotmail.com (C.A. Faller), agenordenoni@gmail.com (A. De Noni Jr),
oscar.rkm@gmail.com (O.R.K. Montedo).

http://dx.doi.org/10.1016/j.ceramint.2016.05.065
0272-8842/& 2016 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572 12557

6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12570
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12571
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12571

1. Introduction size in each ceramic material type identified in the literature and
compares it to conventional sintering (CS). It also presents a CS
The properties of polycrystalline ceramics are controlled by the contribution to the improvement of mechanical properties.
microstructure. The density, grain size and presence of hetero-
geneities in the microstructure are carefully controlled to improve
the properties and reliability of ceramics [1]. Highly dense cera- 2. Historical context
mics with nanometric or submicrometric grain size are difficult to
obtain through conventional sintering. Studies about controlled sintering methodology date back to
However, there are some manufacturing routes available for the mid-1960s, which was when the first studies on the RCS
the production of these ceramics, including colloidal powder method emerged [13–24]. This term was first introduced by Pal-
processing with controlled distribution of particle sizes [2,3], the mour and Johnson in 1965 [18]; the methodology consists of mi-
use of sintering additives [4], pressure-assisted sintering [5,6], crostructure refinement by sintering curve manipulation, and it
spark plasma sintering (SPS) [7] and pulsed electric current sin- features advantages such as simplicity and economy. The proce-
tering (PECS) [8]. These sintering methods may not be cost-ef- dure consists of determining the relationship between the densi-
fective for any materials because the devices are more complex, fication rate and the grain growth rate, thus suggesting tempera-
expensive and difficult to implement. ture and sintering time values for material densification [20,21].
Solid-phase sintering requires relatively high temperatures to Therefore, ultrafine nickel powders were sintered at controlled
facilitate diffusion and, through different mechanisms, promote rates, and high-density materials (  99% relative density) with a
material densification. However, diffusion is the matter transport grain size smaller than 100 nm were obtained.
mechanism, which promotes not only densification but also grain Thus, controlled-rate sintering is an effective methodology in the
growth. Therefore, sintering conditions that allow densification to obtainment of dense bodies with a nanocrystalline structure [21].
occur without simultaneously stimulating grain growth are sui- In 1991, Chu et al. [10] conducted a study about the sintering
table for microstructural refinement. This dissociation between process of conventional alumina (Al2O3) and magnesium oxide
densification and grain growth is what allows highly dense cera- (MgO) powders. Conventional isothermal sintering and pre-coar-
mics and nanometric grains to be produced [9]. sening at a low temperature were comparatively evaluated before
An outstanding technique used to manipulate the micro- the use of conventional isothermal sintering. The two-step sin-
structure during the sintering stage is known as rate-controlled tering technique suggested by Chu (TSS-C) was favorable for sys-
sintering (RCS), which was first reported in 1965 [18]. There are tems with wide-particle-size-distribution conventional powders,
several studies about this issue; however, the equipment advance and it produced uniform ceramic bodies in a simple way [10].
and the improvement in the theoretical and empirical models However, in 2000, Chen and Wang [11] suggested a new
related to the microstructure development and densification al- methodology for the two-step sintering (TSS-CW) technique, ac-
lowed the advancement in the RCS technique. cording to which the ceramic body is rapidly subjected to a high
In the 1990s, Chu et al. [10] introduced the two-step sintering temperature and is subsequently cooled and kept at a lower tem-
(TSS) technique, which is described in the current study as TSS-C. perature (sintering holding time). The authors report that densifi-
According to this technique, the first stage (pretreatment) was cation occurs with no grain growth in a certain temperature range
performed at a relatively low temperature and was followed by a called the "kinetic window". Residual porosity is eliminated at this
higher-temperature stage and subsequent cooling. The process temperature level during sintering with no grain growth, which
allows refinement of the microstructure and, thus, improvement occurs at the final stage [11,12]. The grain growth suppression, but
of the material properties. not its densification, is determined by a network of grain bound-
Later, Chen and Wang [11] suggested a modification in the two- aries anchored by joints in the triple points, which have higher
step sintering technique. This modification, which is identified in activation energy for migration than the grain boundaries. However,
the current study as TSS-CW, has become widely used. This tech- the critical density must be achieved in the first stage so that there
nique consists of suppressing the accelerated grain growth, which are sufficient triple joints throughout the body [12].
usually occurs in the final sintering stage. High-temperature A sufficiently high relative density (70% or higher) is re-
heating is performed and is followed by structural freezing. Fast commended in this first stage [11,25]. The sintering temperature
cooling at a constant rate suppresses grain growth but allows decreases to a critical level, leading the triple joints to interrupt
densification to occur [11,12]. The technique may be successfully the grain growth, whereas densification is not affected. The sam-
applied to many ceramic materials, thus enabling high-density ples must then be exposed to prolonged heating at a low tem-
microstructure refinement and improving several properties of the perature in the second isothermal stage [11]. Fig. 1 schematically
materials. depicts the comparison between the TSS-C and TSS-CW sintering
The aim of the current review is to present the state of the art curves.
of two-step sintering applied to ceramic materials. Thus, the study
is divided into three main topics. First, a brief history of the two-
step sintering technique (TSS) evolution will be presented, fol- 3. Two-step sintering method by Chu
lowed by a presentation of papers that used the methodology
suggested by Chu (TSS-C) and, finally, studies that addressed the The two-step sintering suggested by Chu (TSS-C) was applied to
two-step sintering technique suggested by Chen and Wang (TSS- different ceramic materials. Huang et al. [26] evaluated the ob-
CW). This section shows the effect of the addition of doping agents tainment of silicon nitride (Si3N4) through this method and found
and that of the processing characteristics on the grain density and increased material densification, although the grain size growth
12558 N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572

Fig. 1. Illustrative representation of the TSS-C and TSS-CW sintering plots.

occurred at a higher temperature in the second stage. This growth of denser regions in the initial densification stages. Thus, the
was related to the great structural disarray. density variations in the ceramic body were reduced, leading to a
The use of high-purity Al2O3, both with and without MgO ad- more homogeneous final microstructure. Fig. 2 shows the sche-
dition, was evaluated by applying thermal pretreatments. The use matic representation of the TSS technique.
of low temperatures (800 °C) and long periods (50 h) improved Sato and Carry [28] determined the effect of pretreatments on
the material densification and the microstructure [27]. The effect two types of ultrafine aluminas (submicron-sized particles with
of thermal pretreatment performed in a high-quality alumina was median volume diameters of 0.2 mm and 0.4 mm). Before the sin-
also evaluated [1]. The delay in the initial densification stage was tering, pretreatments were executed at 820 or 920 °C for 50 h. They
observed at low temperature levels owing to the elimination of the concluded that the use of pretreatment delays the onset of abnor-
finer particles. Therefore, the elimination of these particles in the mal grain growth, resulting in a more uniform microstructure be-
first level reduced the differential densification and the formation fore densification.

Fig. 2. Schematic diagram illustrating the microstructural refinement produced by the TSS technique [1].
N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572 12559

ceramics with 3-mm-mean-size uniform grains were obtained


following the pretreatment.

4. Two-step sintering method by Chen and Wang

The two-step sintering technique suggested by Chen and Wang


(TSS-CW) has been widely used because it is a methodology used
to obtain ceramic bodies with a controlled microstructure.
The effects of the addition of doping agents and the processing
characteristics on grain density and size are detailed in this section
for each type of ceramic material identified in the literature, and
their contributions to the improvement of mechanical properties
are discussed.

4.1. Yttrium oxide (Y2O3)

As previously mentioned, Chen and Wang [11] suggested a new


Fig. 3. Grain size of Y2O3 in TSS and CS [11].
method for sintering in two stages. Yttrium oxide powders with
sizes ranging from 10 to 100 nm were evaluated. High densifica-
tion was achieved without grain growth in pure Y2O3. The first
Kim and Kishi [29] studied the effect of pretreatments on the stage of temperature (T1) was 1310 °C, and that of the second
resistance and the subcritical crack growth in alumina. The alu- stage (T2) was 1150 °C for 20 h. Smaller grain size was achieved by
minas sintered by the hot-pressing technique had flexural strength using a lower value for T1 (1250 °C). Fig. 3 shows the relationship
values of 400–500 MPa, whereas the pretreated ones (1000– between the relative density and the grain size of pure Y2O3 ob-
1200 °C for 10 h) increased their resistance to 750 MPa. The au- tained through conventional and two-stage sintering processes.
thors concluded that the pretreatment increases the fracture High-density samples (approximately 100%) with grain size
toughness of the grain boundary, reducing the subcritical growth smaller than 200 nm were obtained through TSS, and they were
rate of sintering flaws and resulting in strengthening. The result much smaller than those obtained through CS (  400 nm).
led to increased material resistance. The bodies reached relative The effect of magnesium and niobium additions was also
density exceeding 90% in SrBi2(VxNb1  x)2O9 ferroelectric ceramics evaluated. The addition of 1 wt% Nb2O5 shifted the kinetic window
(x between 0 and 0.3) [30]. to a higher temperature. However, the addition of 1 wt% MgO had
A study was conducted on pure alumina and on alumina doped the opposite effect. According to the authors, the use of doping
with 140, 500 and 2500 ppm of MgO, vacuum sintered at 1700 °C, agents is even more satisfactory when the difference in the ki-
with and without thermal sintering pretreatment at 800 °C for netics between grain-boundary diffusion and grain-boundary mi-
50 h in air [31]. The thermal pretreatment and the greater MgO gration is exploited to achieve densification without growth at
addition led to decreased grain growth, resulting in a more lower temperatures. A fully dense microstructure with a smaller
homogeneous structure. The grain size in the samples was reduced grain size in the range of 60 nm was obtained using Y2O3 doped
from 16 mm to less than 10 mm owing to the addition of 500 ppm with 1 wt% MgO, as shown in Fig. 4. The sintering process occurred
MgO. The grain size was reduced to 8 mm with 2500 ppm MgO. at T1 1080 °C, with 76% initial relative density. The material was
Porosity was also reduced. However, the addition of dopant agents, then kept at 1000 °C for 20 h (T2).
although low, reduced the density of the samples. The combina- Wang et al. [12] also evaluated yttria powders with different
tion of MgO and Al2O3 and the thermal pretreatment technique particle sizes ranging from 10 to 60 nm. Two dopant agents, i.e.,
proved to be helpful in preventing abnormal grain growth.
Han et al. [32] obtained uniform hexagonal grains of yttria-
stabilized zirconia. Three different conformation methods were
applied: tape casting (TCS), tape calendaring (TCL) and gel casting
(GC). Grain sizes of 3–5 mm and between 8 and 15 mm were ob-
tained for TCS and GC samples, respectively, in the conventional
treatment (1450 °C for 2–4 h). Samples with 1–4 mm grain sizes
were produced using the TSS-C process (T1 at 1000 °C, holding
time of 2 h, and T2 at 1400–1450 °C for 2 h) and the TCL con-
formation process.
ZnO was doped with aluminum (Al) and initially subjected to
thermal treatment at a low temperature and to external pressure
by using the same methodology. The specimen was heated to 800–
1000 °C at a 5 °C min  1 heating rate under external pressure (0,
1 or 2 MPa) for 10 min Subsequently, the pressed specimens were
sintered without pressure at 1250, 1300, and 1350 °C for 2 h in air
using a separate furnace. Effective TSS-C implementation reduces
the final sintering temperature to obtain similar density to dif-
ferent dopant contents. It was also reported that the cohesion
between the particles during the initial thermal treatment must be
established and that the high specific area should be preserved for
successful application of TSS-C [33]. Liu et al. [34] studied
KSr2Nb5O15 (KSN) ceramics manufactured through TSS-C. Dense Fig. 4. Microstructure of fully dense Y2O3 doped with 1 wt% MgO [11].
12560 N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572

magnesium (Mg2 þ ) and niobium (Nb5 þ ), were used in addition to at temperatures up to 900 °C. The amount of liquid phase de-
pure yttria, and they replaced yttria at 1 wt%. The samples were creased in the second stage because of cooling at 825 °C, and the
sintered through CS and TSS-CW. The heating rate in CS was ZnO-ZnO direct contact was improved by suppressing the grain
10 °C min  1 to 1600 °C (no holding time) with natural cooling. boundary migration and maintaining the active grain boundary
Regarding TSS-CW, the heating rate was 10 °C min  1 to T1, in diffusion to obtain densification. Thus, full densification was
which it cooled at the rate of 50 °C min  1 to T2 and remained achieved in the second stage (10-h holding time) without grain
there for 6–30 h. By comparing the results found by Wang et al. growth.
[12] to the study by Chen and Wang [11], it appears that both Another study using ZnO varistors [41] found that the molar
yielded pure Y2O3 through TSS-CW with grain sizes smaller than composition of the varistor was 95.01% ZnO and 4.99% other oxi-
200 nm. Relative density of 99% and a grain size of 123 nm were des (4.99% Bi2O3, Sb2O3, Mn3O4, Co3O4, NiO, Cr2O3 and
obtained using 1250 and 1100 °C as T1 and T2, respectively. The Al(NO3)3  9 h2O). Other sintering systems have been tested with
same evidence was verified regarding the influence of doping T1 ranging from 1000 to 950 °C and T2 ranging from 875 to 825 °C.
agents. Finally, Wang et al. [12] concluded that the Y2O3 TSS-CW, The decrease in the temperature of the first stage reduced the
with or without doping agents, develops by grain boundary dif- density and grain size at the beginning of the second stage. Thus,
fusion. The grain growth suppression may be attributed to triple- more time was required in the second stage to achieve an almost
point immobility, which is facilitated by pre-firing high-tempera- fully densified microstructure. A smaller grain size (750 nm) was
ture single-step sintering. obtained with sintering temperatures at 950 and 825 °C in the first
and second stages, respectively, and a 10-h holding time in the
4.2. Zinc oxide (ZnO) second stage.
By comparing this study with those conducted by Durán et al.
Mazaheri et al. [35] evaluated the effect of temperature on the [39,40], it appears that despite the different compositions (in re-
grain density and size of high-purity ZnO compacts (99.7%) with a lation to the used dopant agents), the decrease in T1 and T2, even
mean particle size of 31 nm. Three different sintering regimes with longer T2, allowed material with higher density and smaller
were used, i.e., 850–750, 800–700 and 800–750 °C, respectively, grain size to be obtained.
for T1 and T2. Specimens with density and grain size similar to Among the studies presented herein, it is concluded that the
those of the CS method (approximately 4 mm) were obtained using addition of dopant agents to ZnO in conjunction with TSS-CW
the temperatures of 850 and 750 °C. The samples were dense, but helped suppress grain growth.
there was no grain growth suppression. Grain growth and 86%
densification were observed using lower temperatures such as 800 4.3. Aluminum oxide (Al2O3)
and 700 °C (T1 and T2, respectively). The best condition was ob-
tained at the temperatures of 800 and 750 °C. Submicrometric Alumina is the most studied ceramic. The initial alumina con-
grains ( 680 nm) were obtained with 98% relative density. It was ditions (pre-sintering), such as the smaller particle size, the chosen
concluded that the temperatures of both sintering stages play a conformation method and the reduced pore size in the green body,
significant role in densification and grain growth. affect the success of TSS-CW and the ceramic crystal symmetry.
Wu [36] conducted a study on pure ZnO with mean particle Adding dopant agents to the alumina also hinders grain growth.
size of 400 nm and sintered in four different sintering plans. The Li and Ye [42] applied the TSS-CW method to manufacture
temperature ranged from 1000 to 1300 °C in the first stage and alumina-based nanoceramics. Alumina with a mean particle size of
from 950 to 1250 °C in the second stage, with a 12-h holding time. 10 nm was synthesized using the doped Al2O3 seed poly-
The best condition was obtained at 1100 and 1050 °C, wherein the acrylamide gel method. Two sintering plans were used in their
relative density was 95.1% and the grain size was 3.9 mm. Com- study, both of which used alumina conformed at 1120 MPa. The
pared with the study by Mazaheri et al. [35], the TSS-CW method samples were first heated at 1450 °C for 1 h in the first plan and
is inefficient for ZnO in regard to larger particle sizes and higher then at 1350 °C for 34 h in the second sintering stage. The relative
sintering temperatures. Wu [36] explained this inefficiency based density of the samples was 96%, and the mean grain size was
on the studies by Maca et al. [37,,38]. According to these authors, approximately 90 nm, as estimated from the peak broadening of
the TSS-CW efficiency depends more on the crystalline structure X-ray diffraction. The second sintering plan occurred at 1380 °C for
than on other factors such as particle size, and the cubic structure 1 h. Subsequently, the samples were cooled at 1330 °C with a 50-h
may be more easily densified by TSS-CW than the hexagonal and holding time. Despite the reduced temperature and the longer
tetragonal structures. holding time in the second stage, the relative density was 95%, and
ZnO with 2 wt% Al2O3 was also evaluated [36]. Five temperature the obtained grain size was slightly smaller than 70 nm.
plans (1400–1350, 1450–1400, 1500–1250, 1500–1350 and 1500– Bodišová et al. [43] assessed the application of TSS-CW to
1450 °C) were used, and the density and grain size were evaluated. submicrometric alumina powders. The material had a mean par-
Samples with 99.8% relative density and 3.6 mm grain size were ticle size of 150–200 nm. The samples were sintered at 1330–
obtained at temperatures of 1500 and 1350 °C. The authors report 1450 °C in the first stage and at 1100–1160 °C in the second stage,
that the doping changed the TSS-CW kinetic window and helped with 3- and 24-h holding times, respectively. According to the
densify the material. The Al2O3 reacted with ZnO to form ZnAl2O4 authors, the minimum density to be achieved in the first sintering
(spinel), and these precipitates contributed to the higher density stage should correspond to 92% relative density. They compared
and prevented grain growth. A relative density improvement from this value to the density obtained in the study by Li and Ye [42],
97.4% to 99.8% was observed compared with CS (1500 °C for 12 h). which was 82% in the first stage, and attributed the densification
The grain size was reduced from 9.1 to 3.6 mm [36]. increase to the smaller particle size and the smaller pore size in
Duran et al. [39,40] evaluated ZnO varistors doped with Bi2O3, the green body. The sintering plan that yielded the best results
Sb2O3, CoO, and MnO. Two sintering profiles (900–825 and 940– was T1 at 1400–1450 °C and T2 at 1150 °C. The obtained relative
825 °C) were evaluated by varying the holding time in T2 between density was 98.8% with a 0.9 mm grain size. Regarding an ap-
10 and 6 h, respectively. The grain size for the first profile was proximate relative density value, the grain size was 1.6 mm com-
500 nm. Regarding the second profile and the shorter holding time pared with CS (temperature at 1350 °C with a 1-h holding time).
in T2 (6 h), the grain size increased to 700 nm. The presence of Hesabi et al. [44] applied TSS-CW to 150 nm submicrometric
dopant agents enabled the liquid phase formation in the first stage alumina. Five sintering plans were used in the TSS-CW to sinter
N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572 12561

the material. T1 ranged from 1200 to 1300 °C, whereas T2 ranged limitations in the compaction process is usually the reason why
from 1100 to 1200 °C. Holding times of 15, 50 and 64 h were the ceramics are not fully densified. Therefore, all compositions
evaluated for T2. Two compaction types were tested–namely, were sintered for 1 h at 1450 °C, which is the condition required
uniaxial pressing and cold isostatic pressing. The authors con- for the complete densification of non-doped powders (1 h at
cluded that, unlike the study by Bodišová et al. [43], the initial 75% 1300–1350 °C). The maximum relative density did not exceed 99%,
density of the samples conformed by cold isostatic pressing is which was considered to be the limit for samples compacted by
sufficient to remove pores in the second stage. In addition, the uniaxial pressing and by cold isostatic pressing. To minimize grain
difference between the relative density values in the first stage growth, the temperature in the second stage was determined
may be attributed to the green compact conformation method. based on the lowest temperature at which the relative density of
The best results were achieved by using 1250 °C in T1 and 1150 °C 99% was achieved in less than 24 h of holding time. The tem-
in T2. By comparing their results to those obtained by Bodišová peratures in the second stage were 1150 °C for pure alumina,
et al. [43], it was possible to see that T1 was significantly lower 1200 °C for MgO-doped alumina, and 1300 °C for
(1250 °C), although T2 (1150 °C) remained the same and that the Y2O3-and-ZrO2-doped alumina. Regarding the non-doped alumi-
final grain size was 0.5 mm. It was also concluded that the size and na, coarsening was found in the microstructure at the final sin-
distribution of the pores after the first sintering stage determined tering stage, and a final grain size of 650 nm was obtained. The
the critical density, which ensured the success of two-step sin- doping using MgO, Y2O3 and ZrO2 resulted in grain growth sup-
tering [44]. The mechanical properties of the alumina samples pression at the final sintering stage. The doped samples were
were evaluated. Samples with the best relative density (98%) and sintered up to the relative density near 100%, and the mean grain
smaller grain size (0.5 mm) had a hardness value of 18.2 GPa, ex- size was 500 nm. The MgO-based doping was considered to be
ceeding the 15.01 GPa value of a sample with larger grain size more advantageous because of the lower sintering temperature
(0.85 mm). The fracture toughness value between 4.24 and required to achieve high final density. The grain size obtained by
4.29 MPa m1/2 was similar in both samples and is in the range of adding dopant agents to the alumina (500 nm) was the same value
values reported in the literature. found by Hesabi et al. [44] in a study conducted on pure alumina.
A zirconia-doped alumina (5 wt%) with submicrometric grains Two types of hexagonal α-alumina with particle sizes of
and high density was evaluated [45]. The alumina particle size was 100 nm (TAI) and 240 nm (REY), one 60 nm tetragonal zirconia
150 nm. Two tests were initially performed. The first test was (Z3Y) and one 140 nm cubic zirconia (Z8Y) were investigated. The
performed to determine the first sintering stage. The samples were thermal treatment was performed at heating rates of 10 °C min  1
heated at 10 °C min  1 to 1525 °C. The second test consisted of to 800 °C and 5 °C min  1 to T1 (it varied according to each ma-
isothermal sintering in the range of 1250–1450 °C at the rate of terial). Subsequently, the samples were kept at T2 for 0, 5, 10, 15
10 °C min  1 for 0–8 h. TSS-CW was performed after the pre- and 20 h using a cooling rate of 60 °C min  1. By comparing the
liminary tests. T1 and T2 ranged from 1400 to 1450 °C and from results, the authors found little success in the alumina sintered
1300 to 1400 °C, respectively. The holding time in the second stage through TSS-CW. The grain size values were lower, but they
was evaluated for each sintering plan at 4, 8, 12 and 24 h. The showed little difference: 0.66 mm TAI for CS, 0.51 mm for TSS-CW
authors concluded that 1450 °C is the best temperature for the and a decrease from 0.93 to 0.82 mm for REY alumina. The zirconia
first sintering stage because lower final density was obtained using samples showed Z8Y improvement and no change in Z3Y. The Z8Y
lower temperatures, regardless of the holding time in the second relative density was 99.54%, and the grain size was 3.58 mm in CS.
stage. Regarding the second sintering stage, 1350–1400 °C was Regarding the TSS-CW, T1 was 1440 °C and T2 was 1340 °C for
found to be the best temperature range because there was good 10 h, reaching 99.54% relative density and 1.71 mm grain size. A
material densification, without the apparent surface diffusion ef- smaller microstructural refinement was found in the hexagonal
fect on grain growth. The authors also evaluated the presence of alumina, and no effect was found in the tetragonal zirconia. The
zirconia. Compared with the study by Bodišová et al. [43], the authors suggested that the efficiency of two-step sintering in-
elevated temperature in the second stage was explained by the creases because of the ceramic crystal symmetry, which is more
presence of zirconia, which increases the activation energy to influential than the particle size and the microstructure of the
densify the alumina. Another observation was related to the grain green body [38].
growth inhibition caused by zirconia, which induced a change in
the kinetic window to a higher temperature and changed the re- 4.4. Yttria (Y2O3)-stabilized zirconia (ZrO2)
lationship between grain size and density, compared with pure
alumina. The final ceramics obtained by the introduction of zir- The TSS-CW application to yttria-stabilized zirconia has sa-
conia had relative density exceeding 99% and grain size of 0.62– tisfactorily reduced grain growth compared with the conventional
0.88 mm. By comparing the study by Hesabi et al. [44] and that by method (CS); however, a certain limitation was observed in tech-
Bodišová et al. [43], who both worked with uniaxial pressing, it niques such as SPS (sparking plasma sintering) and microwave
was possible to see that the effect of zirconia was decisive in grain sintering. The good homogeneity of the green samples proved to
growth inhibition. However, Hesabi et al. [44] obtained a smaller be an important condition for TSS-CW applicability.
grain size (0.5 mm). Lee [47] evaluated 3 mol% Y2O3-stabilized zirconia (Z3Y pow-
Galusek et al. [46] evaluated the influence of doping agents on der) and organic polymers of polyvinyl butanol (PVB) and dibu-
alumina, such as MgO, which is known to increase densification tylphthalate (DBP) in solvent ethanol. The mean particle size was
and grain growth, and Y2O3 and ZrO2, which are known as grain 0.27 mm. The conformation stage was carried out on an injector
growth and densification inhibitors. These oxides were added with 680-bar pressure at 140 °C, and the samples were thermally
(500 ppm) to an alumina with a 150 nm particle size and sintered treated through CS and TSS-CW. The CS technique used a heating
through TSS-CW. The temperature in the first stage was de- rate of 10 °C min  1 to 1500 °C with a 30-min holding time. Re-
termined by heating the samples at a temperature range between garding the sintering through TSS-CW, the same heating rate was
1300 and 1475 °C and immediately cooling them to room tem- used to T1 at 1500 °C for 5 min; the second stage was carried out
perature. Regarding the dopant-free compositions with MgO ad- at 1300 °C (T2) for 10 h. The TSS-CW technique obtained samples
dition, the temperature in the first stage was 1360–1400 °C. In with a 0.59 mm grain size, whereas the CS samples had a 1.05 mm
contrast, the temperature was higher (1450–1475 °C) in the com- grain size. The flexural strength of the samples obtained through
positions with Y2O3 and ZrO2. The presence of defects due to the TSS-CW technique was approximately four times greater than
12562 N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572

that of the samples obtained though CS, i.e., 1078 and 295 MPa, in TZ0 (90 nm) compared with 3YTZ (115 nm). By comparing the
respectively. used sintering processes, it was found that the TSS-CW produced
Mazaheri et al. [48] studied ZrO2 with 3 mol% yttria, with a samples with a smaller grain size than those found in the CS.
mean particle size of 75 nm. The powder was compacted in a cy- However, between both techniques, SPS yielded samples with
lindrical die at 150 MPa to produce green compacts. The samples smaller grain sizes [50].
were sintered in a single stage at a heating rate of 5 °C min  1 and Xiong et al. [51] evaluated 3 mol% yttria-doped zirconia. The
a temperature range from 1100 to 1500 °C with 50 °C intervals powders were spherical granules ranging from 60 to 120 mm by
between the temperatures. The holding time ranged from 1 to 8 h. thermal spraying. Two different methods were evaluated for the
The TSS-CW specimens were heated at 5 °C min  1 to T1 (1300 and first stage of the thermal treatment. One of them was pressureless
1350 °C), kept at T1 for 1 min, and cooled at 50 °C min  1 to T2 sintering, which was carried out in a self-made air furnace. First,
(1050 and 1250 °C), with holding times between 2 and 30 h. The samples were conformed by uniaxial pressing and then compacted
CS technique obtained compacts with a grain size of 275 nm; by cold isostatic pressing. The other method was the high-pressure
however, the use of CW-TSS showed that the grain growth rate spark plasma sintering (referred to as HPSPS) technique. The CS
might be further reduced. The sintering program that showed the technique used a heating rate of 10 °C min  1 to 1250–1350 °C,
best results used T1 at 1300 °C for 1 min The critical density ob- with a 1-h holding time. Regarding the TSS-CW, the heating rate
tained in the first stage was 83% of the theoretical density, and the was 10 °C min  1 to T1 at 1250–1300 °C, with a 1-min holding
second stage was processed at 1150 °C for 30 h. Samples with time. The temperatures of the second stage were 1150 and 1175 °C
maximal densification and a 110 nm mean grain size were ob- with 30- and 20-h holding times, respectively. The second method
tained. By comparing the study by Mazaheri et al. [48] to the study was carried out with the initial pressure of 100 MPa up to 600 °C
by Lee [47], it is observed that the difference in the powder par- with a 3-min holding time. The pressure was then increased to
ticle size (0.27 mm and 75 nm) and the conformation processing 300 MPa, where it remained until the end of the cycle with a
(injection and pressing process) affected the final grain size of the temperature of 1000 °C for 5 min The temperature in the second
samples. Mazaheri et al. [48] obtained bodies with a smaller grain stage was 1100–1175 °C, with a 30-h holding time. By comparing
size; however, the TSS-CW technique proved to be satisfactory the sintering methods, it was found that the TSS-CW (T1 at
when compared with the CS technique in both studies. 1300 °C and T2 at 1175 °C for 20 h) achieved 99.2% relative density,
By using 3 mol% Y2O3-doped tetragonal ZrO2 powder with a mean with a small grain size (184 nm). The TSS treatment through
particle size of 65 nm, Suárez et al. [49] investigated the effects of the HPSPS (T2 at 1175 °C for 30 h) resulted in a smaller grain size
bead-milling treatment on colloidal processing. A comparison be- (173 nm); however, the density was lower (97.46%). Finally, the
tween two processes, i.e., colloidal processing and bead-milling high relative density of 99.92% was achieved in CS (1350 °C);
processing, was performed. Aqueous ceramic suspensions with however, the grain size reached 221 nm [51].
30 vol% of solid loading were prepared in the colloidal processing by Ghosh et al. [52] studied zirconia containing 8 mol% yttria in
dispersing the ZrO2 powder in distilled water using different nanocrystalline form synthesized by the co-precipitation method.
amounts of polyelectrolyte. The aqueous ceramic suspensions were The formation of an easily filterable hydroxide was facilitated by
dispersed in an ultrasonic stirrer for 15 min The slip-cast samples the addition of ammonium sulfate and polyethylene glycol during
were dried slowly to avoid cracking and were cold isostatically precipitation. The precipitate was then calcined to produce na-
pressed at 392 MPa. The bead-milling processing uses bead-milling nocrystalline powder. CS was carried out between 1000 and
equipment comprising ZrO2 beads with a 50 mm diameter at 1400 °C. The heating rate was 5 °C min  1 to 1000 °C, followed by
4000 rpm. The suspension obtained after 60 min of treatment was decreasing at 1 °C min  1 to the desired temperature (3-h holding
evacuated, slip-cast and cold isostatically pressed at 392 MPa for time). The TSS-CW used a heating rate of 5 °C min  1 to 1000 °C,
comparison with the cast samples without bead milling. The tem- which decreased to 1 °C min  1 until it reached 1125 °C (3-h
perature in the first TSS-CW sintering stage was 1300 °C, without holding time). T2 was 1090 °C (20-h holding time). The powder
holding time, and the temperature in the second stage was 1200 °C enabled the obtainment of relative density exceeding 95% at a low
for 15 h. The temperature of 1250 °C for 3 h was used in CS. By using temperature (1150 °C) in CS. By comparing the grain sizes between
the bead-milling equipment and the TSS-CW, it was possible to see the two sintering processes, CS and TSS-CW, little difference (150–
better sinterability in the samples, with relative density near 100% 180 and 185–220 nm grain sizes, respectively) was found. Re-
and smaller grain size (155 nm for CS and 125 nm for TSS-CW, re- garding the mechanical evaluation, the hardness range was 9–
spectively). The samples without milling showed 99% densification, 11 GPa, and the fracture toughness was 5.3 MPa m1/2 for the two
with grain sizes of 160 nm for CS and 150 nm for TSS-CW. evaluated thermal treatments.
Suárez et al. [50] evaluated the effect of starting powders on Lourenço [53] studied two commercial yttria-stabilized zirco-
sintering. The used starting materials were high-purity monoclinic nia powders synthesized by co-precipitation. One powder had a
zirconia (TZ0, particle size of 74.8 nm), commercial Y2O3 powder tetragonal phase partially stabilized with 3 mol% yttria (Z3Y), and
and 3 mol% Y2O3-stabilized tetragonal ZrO2 (3YTZ, 65 nm particle the other had a cubic phase stabilized with 8 mol% yttria (Z8Y).
size). Suspensions were prepared through colloidal processing and The mean particle size ranged from 25 to 70 nm. The TSS-CW
bead-milling processing [49], and the thermal treatment was sintering process enabled grain growth inhibition through the Z3Y
carried out through three different methodologies: CS, TSS-CW powder and produced dense microstructures with a grain size of
and SPS. Regarding CS, a heating rate of 3 °C min  1 to 1250 °C was approximately 100 nm. The sintering program used T1 at 1320 °C
used, with a 5-h holding time. By comparing TZ0 and 3YTZ, the and T2 at 1250–1270 °C for 12 h. The mean grain size of the single-
authors found that the relative densities were the same (greater step sintered bodies (1500 °C for 2 h) was approximately 200 nm.
than 99.5%) at this temperature, whereas the grain sizes were 113 The grain growth was also inhibited in the Z8Y composition, and a
and 175 nm, respectively. The TSS-CW temperatures were defined mean grain size of 450–460 nm was obtained under the conditions
as T1 at 1300 °C without holding time and T2 at 1200 °C for 15 h. of T1 at 1320–1340 °C and T2 at 1270 °C (12-h holding time). The
However, because the initial compositions were different, there grain size of the samples was 1 and 2 mm in the Cs cycle. The Z8Y
was a decrease in grain size in TZ0 compared with 3YTZ, i.e., 95 grain size found by Lourenço [53] was higher (450–460 nm) than
and 125 nm, respectively. The SPS method was carried out at that obtained by Ghosh et al. [52]. The difference in the results
1150 °C, with a heating rate of 300 °C min  1, pressure of 150 MPa obtained in these two studies may be explained by the different
and 30-min holding time. There was a reduction in the grain size used starting powders and by the TSS-CW sintering programs.
N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572 12563

Lourenço et al. [54] considered the experimental data by grain size compared with those of DP. The DP grain size was
Lourenço [53] and evaluated the TSS-CW method using predic- 2.15 mm, whereas the SC grain size was 1.3 mm. DP specimens were
tions of sintering kinetics based on shrinkage data with Taguchi sintered through microwave sintering and TSS-CW to evaluate the
fractional experimental plans. The authors concluded that T1 may sintering process effect. A smaller grain size (0.295 mm) was ob-
be defined as the temperature at which 75% of the relative density tained through TSS-CW. By using the same conformation techni-
is achieved in Z3Y. The temperature range with typical Arrhenius que and microwave sintering, the grain size was 0.9 mm. According
dependence was taken as a guideline to determine T2 and its re- to the authors, pinning the grain boundaries through immobile
spective holding time. The data found from the sintering kinetics triple junctions in the two-step sintering method ensured grain
predictions were similar to the results of the experiments planned growth inhibition [57].
according to the Taguchi method. However, the methodology was Schwarz and Guillon [58] evaluated zirconia with 8 mol% yttria
not considered suitable for Z8Y because the temperature range and a mean particle size of 58 77 nm. Two methods were applied
required to attain unstable microstructures already yielded an in the conformation of the specimens: pressure filtration and dry
increase in the grain size by nearly one order of magnitude. pressing. A suspension was prepared and filtered under pressure
Nanocrystalline 8 mol% yittria-stabilized zirconia was synthe- in the pressure filtration process. It was then pressed in a me-
sized using modified glycine-nitrate [55]. The TSS-CW used T1 at chanical testing machine with a maximum pressure of 50 MPa,
1250 °C and a heating rate of 5 °C min  1 and T2 at 950 and which was kept for 15 min Dry pressing was the other compaction
1050 °C, with a cooling rate of 50 °C min  1 and a 20-h holding performed at 50 MPa in the graphite tool. The thermal treatment
time. Single-step sintering was performed using the same heating was carried out through the field-assisted sintering technique,
rate at 1500 °C, without holding time, and it was compared with which is known as spark plasma sintering (SPS). The heating rate
the TSS-CW. The variation in T2 allowed the authors to observe was 200 °C min  1 to T1 at 1150 °C (kept for 20 s) at an initial
that the highest temperature (1050 °C) showed the best result pressure of 10 MPa, which was increased to 50 MPa when the
with respect to the grain growth and densification process. When temperature was 200 °C from T1. T2 was 1050 °C, with 2- and 5-h
a lower temperature (950 °C) was used, no densification was ob- holding times. The samples were heterogeneous in the dry pres-
served even after prolonged soaking treatment. However, there sure conformation, and a micrometric grain size was reached
was significant reduction in the grain size values: 2.15 mm for the (1.67 mm), despite the TSS-CW and the incomplete densification
samples sintered through CS and 295 nm for those sintered (93.6% relative density). The samples conformed by pressure fil-
through TSS-CW. Hardness and fracture toughness were also ob- tration showed lower grain growth (190 nm) even for a high
served: 12.87 GPa and 1.61 MPa m1/2 for CS; and 13.51 GPa and density (99.8% relative density). According to the authors, the
3.16 MPa m1/2 for TSS-CW, respectively. homogeneous densely packed green body plays an important role
Mazaheri et al. [56] conducted a comparative analysis of three in the grain growth control for the success of the TSS-CW meth-
sintering techniques using nanocrystalline 8 mol% yittria-stabi- odology and the pressure-assisted sintering method.
lized zirconia synthesized using a glycine-nitrate process. Micro-
wave sintering was evaluated in addition to the two treatments 4.5. Titanium dioxide (TiO2)
(CS and TSS-CW) used in the study by Mazaheri et al. [55]. A
temperature of 1500 °C was used in the microwave sintering at Mazaheri et al. [59] used TiO2 nanopowders with particle size
two different heating rates: 5 °C min  1 (lower-rate microwave in the range of 11–27 nm and containing 77% anatase and 23%
sintering, LMS) and 50 °C min  1 (higher-rate microwave sintering, rutile. The powder was uniaxially pressed at 100 MPa and ther-
HMS). After the samples reached the maximum temperature, they mally treated though CS and TSS-CW. The CS treatment used a
were cooled to room temperature. The relative density value in all heating rate of 5 °C min  1 to temperatures in the range of 500 and
sintering methods was 97%. The grain size was 2.14 mm for CS, 1000 °C for 1 h. The TSS-CW treatment used a heating rate of
2.35 mm for LMS, 0.9 mm for HMS, and 0.29 mm for TSS-CW. A 5 °C min  1 to T1, which was held for 1 h, and a cooling rate of
comparison was also performed between the fracture toughness 50 °C min  1 to T2 (700 °C), with a holding time of up to 30 h. Two
values. The samples treated through HMS (3.17 MPa m0.5) and TSS- temperatures were used as T1 to reveal the structural change
CW (3.16 MPa m0.5) had higher values than did those that were between the anatase and rutile phases: 800 °C with a 25-h holding
conventionally treated (1.61 MPa m0.5) and those treated through time in T2 (TSS1) and 750 °C with a 30-h holding time in T2 (TSS2).
LMS (1.9 MPa m0.5). Despite the larger grain size in the HMS Grain sizes between 1 and 2 mm were obtained in CS for sintering
sample, the fracture toughness value was similar to that of TSS- at 1000 °C for 1 h. By comparing the TSS-CW results, it is possible
CW. According to the authors, this was due to the higher micro- to see that the TSS1 process reached a grain size of approximately
structural homogeneity caused by HMS. It was also reported that 250 nm with a relative density higher than 98% using an amount
despite the known ability of the TSS technique to inhibit ac- of rutile near 100% at the beginning of T2. TSS2 reached a grain
celerated grain growth, factors such as the lack of microstructural size of approximately 100 nm and a relative density greater than
homogeneity and the long total sintering time (4 20 h), compared 98% using an initial amount of 88% rutile and 12% anatase at the
with that of HMS (29 min), show that there are restrictions on the beginning of T2. The use of TSS-CW along with the phase trans-
application of the TSS-CW method [56]. formation evaluated at the TSS2 process allowed a more refined
The influence of two conformation methods, uniaxial dry structure to be obtained (100 nm).
pressing (DP) and slip casting (SC) [57], was evaluated for nano- Mazaheri et al. [60] compared CS and TSS-CW to the SPS pro-
crystalline 8 mol% yttria-stabilized zirconia powder processed cess evaluated by Lee et al. [61] to evaluate the TSS-CW process
through the glycine-nitrate combustion method [55]. Several applicability to TiO2 nanostructures. The CS heating rate of
pressures were tested in the first method (DP), and 600 MPa was 5 °C min  1 was used to temperatures in the range of 500–1000 °C
chosen because it achieved 48% relative density. Regarding SC, for 1 h. The TSS-CW consisted of a heating rate of 5 °C min  1 to
suspensions were prepared with 42–67 wt% solids, and 65 wt% 800 °C (T1), which was held for 1 h, and cooled at the rate of
was used with 60% relative density. The same thermal treatments 50 °C min  1 to 700 °C (T2), which was held for 25 h. By comparing
of the previous study [56] were used: CS, TSS-CW and microwave the obtained data to the study by Lee et al. [61], it was possible to
sintering. By comparing the two conformation methods sintered see that the SPS method used to sinter the TiO2 resulted in grain
by CS, the SC method yielded better results. The improved green size and relative density similar to those found in the TSS-CW
properties resulted in a lower sintering temperature and smaller method. The grain size in the SPS treatment was 200–300 nm, and
12564 N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572

the relative density was 98%, whereas the grain size in the TSS-CW 900 °C for 4 h in CS. By comparing these results with those found
was 250 nm and the relative density was 98%. Considering the SPS by Kim and Han [63], who obtained Dy-doped BTO samples with a
process, the authors state that the TSS-CW method is a simple and 200–400 nm grain size, it is possible to see that Wang et al. [64]
feasible process in the obtainment of TiO2 nanostructures. obtained samples with a smaller grain size (35 nm). The con-
Li et al. [62] used TiO2 with a mean particle size between 30 and formation method, the lower T1 and T2 temperatures, and the
40 nm and performed a two-step sintering thermal treatment using shorter holding time in the second stage were the main differ-
the spark plasma sintering (SPS) parameters assisted by TSS. A low ences observed.
pressure of up to 30 Pa was used in the beginning of the (SPS) Karaki et al. [65] evaluated hydrothermally synthesized BaTiO3
process to reach the maximum temperature. The heating rate was with a 100 nm particle size. The specimens were compacted at
100 °C min  1, and the holding time ranged from 1 to 60 min A 200 MPa. During the sintering stage, the samples were heated at
temperature of 850 °C with a 1-min holding time was used in T1. T2 10 °C min  1 to T1 (1230–1340 °C) and held for 1 min Next, they
featured a temperature range between 810 and 840 °C with high were cooled to T2 (1150 and 1200 °C) at 30 °C min  1 and held for
densification and invariable grain size. The best relative density and 2, 4, 15 and 20 h. The sintering program with T1 at 1320 °C, T2 at
grain size values, 97.7% and 40 nm, respectively, were found using 1150 °C and a 15-h holding time showed the best result. The mean
T1 at 850 °C and T2 at 830 °C. By comparing the TSS-CW, SPS and grain size was 1.6 mm with relative density of 98.3%. The study also
TSS-assisted SPS processes through the D/Do ratio value (found evaluated the piezoelectric material properties, and the authors
grain size/initial grain size), values near 10 were found for TSS-CW, stated that the grain size and density strongly influenced the final
between 3 and 6 for SPS, and near 1 for TSS-assisted SPS. The TSS- properties and noted the advantage that BaTiO3 was manufactured
assisted SPS value showed virtually no grain growth. These results using a low-cost process.
demonstrate that the combination of the SPS and TSS techniques Isostatic pressing was performed at 200 MPa using high-purity
may suppress the growth of grains more effectively than when the BTO with a mean particle size ranging from 10 to 30 nm [66]. The
two techniques are used separately. applied TSS-CW used a heating rate of 10 °C min  1 to T1 (920–
It is clear that the TSS-CW method applied to TiO2 favors grain 1250 °C), which was immediately cooled at 30 °C min  1 to T2
growth reduction, with values similar to those of the expensive (1150–850 °C) with holding times in the range of 4–20 h. CS was
SPS process. In addition, lower T2 combined with a longer holding performed at a heating rate of 10 °C min  1 to 1200 °C for 2 h, and
time allowed smaller grains to be obtained, as observed in other it led to 97% relative density and a 1.2 mm grain size. Regarding
materials. TSS-CW, the relative density also reached 97% using T1 at 980 °C
and T2 at 900 °C with a 4-h holding time; however, the grain size
4.6. Barium titanate (BaTiO3, BTO) was reduced to 70 nm. Powder with a particle size smaller than
10 nm was cold pressed at 200 MPa and treated at 920 °C in T1 and
The BaTiO3-related literature shows that lower sintering tem- at 850 °C for 20 h in T2, reaching 99.6% relative density and a
peratures are necessary in both TSS-CW process stages to obtain a smaller grain size between 8 and 10 nm. According to the authors,
smaller grain size. The initial particle size of BaTiO3 and the ad- it was the first study that obtained a dense BTO nanocrystalline
dition of dopant agents also help obtain a smaller final grain size. structure with such a reduced grain size using pressure-free
Kim and Han [63] evaluated a BaTiO3 nanometric composition sintering.
with a particle size of approximately 17 nm and a BaTiO3-doped Huan et al. [67,68] used TSS-CW to prepare BTO with an initial
mixture of 1 mol% dysprosium (Dy), whose samples were pressed particle size of 100 nm and different final grain sizes. The com-
at 300 MPa. The TSS-CW treatment used T1 at 1300 °C and T2 at pacts were isostatically pressed at 200 MPa. The samples were
1100 °C for 0–20 h. The CS temperature was 1200 °C, for 1–5 h. The heated at 10 °C min  1 to T1, and they were then cooled at
density achieved in the CS treatment with a 5-h holding time was 20 °C min  1 to T2. Different sintering conditions were applied. The
5.6 g cm  3, corresponding to more than 93% of the theoretical T1 temperatures were 1250, 1280, 1310, and 1340 °C, and the T2
density and grain size in the range of 5–6 mm. Regarding the TSS- temperatures were 800, 850, 900, 950, 1000, 1050 and 1100 °C. The
CW treatment, the best result was found using a 20-h holding holding time in the second stage was 24 h in all evaluated con-
time, in which the relative density was 5.7 g cm  3 (approximately ditions. A number of BTO ceramics with high density (greater than
95% relative density) with a grain size near 1 mm. The use of 95%) and different grain sizes (from 0.29 to 8.61 mm) were suc-
1 mol% BaTiO3-doped Dy led to approximately 93% relative density cessfully produced through TSS-CW. In addition, it was observed
with a smaller grain size (between 200 and 400 nm). that the piezoelectric coefficient is strongly influenced by grain
Wang et al. [64] evaluated high-purity BaTiO3 powders with a size, enabling the obtainment of high-performance BTO ceramics
particle size between 10 and 30 nm and a nanocrystalline ferrite through TSS-CW.
composed of Ni0.2Cu0.2Zn0.6Fe2O4 with a 10 nm particle size. The
conformation was performed by isostatic pressing at 200 MPa. 4.7. Carbides
Different sintering programs were analyzed for TSS-CW. The
heating rate was 10 °C min  1 to T1 at 950–1250 °C for BTO and at SiC (silicon carbide) nanostructured ceramics were obtained
850–930 °C for ferrite. The cooling rate for T2 was 30 °C min  1, through TSS-CW. According to the literature, TSS-CW allows the
and T2 varied from 1150 to 850 °C for BTO (holding time between sintering temperature to be reduced by 50–200 °C compared with
2 and 20 h) and from 750 to 870 °C for ferrite (holding time be- SC, with the same relative density. Thus, the particle size obtained
tween 4 and 12 h). Samples with more than 96% relative density through CS is approximately 2–3 times greater than that obtained
were obtained for all sintering runs by applying TSS-CW. Up to through TSS-CW, with significant influence on the mechanical
100% relative density was achieved in a ferrite sample. The sample properties.
sintered at 950 °C in T1 and at 900 °C in T2 for 2 h showed the best Lee et al. [69] evaluated the two-step sintering process used to
result for BTO, in which the obtained grain size was 35 nm obtain SiC (silicon carbide) nanostructured ceramics in the pre-
(starting particle size of 10 nm). A grain size of 1.2 mm was sence of the liquid phase. The SiC composition containing 7 wt%
achieved in CS (starting particle size of 30 nm) at 1200 °C for 2 h. Al2O3, 2 wt% Y2O3 and 1 wt% CaO and a mean particle size of
The ferrite sintered at 850 °C in T1 and at 800 °C for 6 h in T2 20 nm was processed by milling in a vibratory mill. The compacts
exhibited the smallest grain size (200 nm) among the tested were sintered at 20 MPa using three different sintering programs.
samples. Samples with a grain size of 2 mm were obtained at The CS specimens were sintered at 1500 and 1550 °C for 8 h and at
N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572 12565

1750 °C for 5 and 30 min The second sintering program—the so- hot-pressing sintering. Regarding conventional hot-pressing sin-
called "zero-time sintering"—had no holding time at 1700, 1750, tering, the bodies were hot-pressed at 1450–1900 °C, with 0.13 Pa
1800 and 1850 °C. Finally, the TSS-CW was performed with T1 vacuum, a temperature range of 50 °C and a heating rate of
ranging from 1650 to 1750 °C and with T2 at 1500 °C or 1550 °C, 10 °C min  1. The specimens were held at the highest temperature
with an 8-h holding time. Among the evaluated sintering pro- for 3 min under a pressure of 39.6 MPa to obtain a uniform tem-
grams, TSS-CW with T1 at 1750 °C and T2 at 1550 °C for 8 h perature in the entire sample. Regarding TSS-CW, the samples
showed the best results. Specimens with 99% relative density and were heated under the same CS conditions, with T1 at 1750 °C. The
a 43 nm grain size were obtained. According to the authors, the applied cooling rate was 60 °C min  1, and T2 was evaluated at
size and density of the pores obtained after the first sintering stage 1450, 1550 and 1650 °C for 2, 4 and 8 h under a pressure of
are critical to obtain high density in the second sintering stage. 39.6 MPa. A temperature of 1550 °C in the second stage resulted in
Relative densities exceeding 85% in the first stage were required in 99% densification and a 2.58 mm grain size. At 1650 °C, the relative
this study. Lee et al. [69] concluded that TSS-CW might be applied density was also near 99%, but the grains grew to 2.99 mm. The
to sintering in the presence of the liquid phase. temperature of 1550 °C provided a more homogeneous and more
Lee et al. [70] investigated the TSS-CW process parameters consolidated structure than that of CS. The mechanical properties
using the same material: SiC with the addition of Al2O3, Y2O3 and of the nanocomposite were improved; the hardness improved
CaO. Twelve experimental runs were conducted by varying T1 from 16.7 to 18.4 GPa, whereas the flexural strength increased
(between 1550 and 1750 °C), T2 (1550 and 1600 °C) and the from 976.6 to 1283.7 MPa. The fracture toughness, which was
holding time in the second stage (0, 1, 4 and 8 h). After analyzing 10.2 MPa m0.5 in CS, increased to 12.95 MPa m0.5 in TSS-CW.
the results, the authors suggested that SiC nanostructures might Qu and Zhu [74] evaluated the two step hot-pressing sintering
be manufactured using T1 of 1700–1750 °C and T2 of 1550–1600 °C in WC-Al2O3 composites (WC–40 vol% Al2O3). To perform the hot-
for 8 h. The smaller grain size obtained in their study was near pressing sintering, pressure of 39.6 MPa was used in 0.13 Pa va-
40 nm, approximately two times the initial particle size. The au- cuum, with heating rate of 10 °C min  1. The temperature range
thors suggested the existence of a processing window for nanos- used in CS was 1450 b750 °C for 90 min, with temperature inter-
tructured microstructures, i.e., a "quasi-frozen region". In this vals of 50 °C. Regarding the two step hot-pressing sintering, the
window, there is a separation of grain growth densification during samples were heated at 1650 and 1600 °C (T1), kept for 3 min, and
the liquid-phase sintering in other liquid-phase sintered systems. then cooled down to T2 (1500, 1450 and 1400 °C), with holding
Magnani et al. [71] used α-SiC with sintering aids, boron (B) and times of 2, 4 and 6 h. The sintering regime with T1 at 1600 °C and
carbon (C), as the starting powder to manufacture SiC ceramics by T2 at 1450 °C inhibited the accelerated growth of grain size, which
pressureless sintering using the TSS-CW method. The specimens was 2.38 mm (2.79 mm for CS) with 99% relative density. The im-
were obtained by uniaxial pressing at 60 MPa, followed by cold provements in the samples due to microstructure refinement re-
isostatic pressing at 200 MPa. The sintering stage was carried out flected in the mechanical properties. The hardness of the CS-
using an argon stream at 1 atm. CS used the temperature of 2200 °C treated samples increased from 18.65 to 19.71 GPa. The fracture
for 1 h, whereas SST-CW used T1 at 2100 °C and T2 at 2050 °C, with toughness and flexural strength increased from 10.43 MPa m0.5
a 7-h holding time. There was sharp grain growth (more than and 756 MPa m0.5 to 12 MPa m0.5 and 1285 MPa m0.5, respectively.
100 mm) in CS, whereas the grain size was approximately 30 mm in
TSS-CW. These values were obtained for virtually the same relative 4.8. Corundum abrasive
density, 97% and 97.7%, respectively. The authors concluded that the
TSS-CW carried out at a temperature lower than that of CS (2050 °C Few studies related to corundum abrasive sintering through
instead of 2200 °C) leads to the same densification with grain TSS-CW have been identified. However, the obtained results de-
growth suppression. With respect to the evaluated mechanical monstrate that this method is effective in grain growth inhibition
properties, it was observed that the flexural strength was most in- and allows better mechanical properties to be obtained compared
fluenced by grain size reduction. The CS value was 341 MPa, with CS. By considering the same maximum sintering temperature
whereas the SST-CW value increased to 556 MPa. The hardness and and the same holding time, the grain size is 5–10 times greater in
fracture toughness values were the same for both processes. CS, thus significantly affecting hardness, fracture toughness and
Magnani et al. [72] used the same starting material, commercial wear resistance.
SiC powder containing boron carbide and carbon as sintering aids, Li et al. [75] evaluated corundum abrasive prepared by the sol-
to manufacture specimens through uniaxial pressing at 100 MPa, gel process with sintering additives (MgO-CaO-SiO2). Two sinter-
followed by cold isostatic pressing at 200 MPa, by changing the ing programs were applied: CS and TSS-CW. CS was carried out at
sintering temperatures. The CS treatment was carried out at 2130 °C 1250 and 1300 °C (T1) at the rate of 4 °C min  1, without holding
for 1 h, whereas TSS-CW was conducted with T1 at 2030 °C and T2 time, followed by cooling to room temperature (25 °C). Next, the
at 1980 °C for 7 h. By reducing the temperature of the second stage samples were heated again to 1150 °C (T2) and held for 5 h. Al-
from 2050 to 1980 °C, specimens were obtained with a refined though the authors named this sintering plan CS, Li et al. [76]
microstructure comprising equiaxed grains and some elongated renamed it as a third sintering method called CTS. The same
grains with a 10-mm grain size (the grain size is approximately temperatures were used in TSS-CW, i.e., T1 at 1250 and 1300 °C
100 mm in CS). Despite the microstructural differences, samples and T2 at 1150 °C for 5 h. Materials with relative density greater
subjected to CS and TSS-CW exhibited very close relative densities than 99.2% were obtained in both TSS-CW cycles. The grain size of
of 98.6% and 98.4%, respectively. Better results were found using a the samples was smaller than 100 nm, whereas the grain size in CS
lower sintering temperature for SiC consolidation. The flexural ranged from 480 to 610 nm. The liquid phase formation sintering-
strength values were higher (530710 MPa) than those of the CS- aid oxides (MgO-CaO-SiO2) contributed to grain growth suppres-
sintered samples (480712 MPa). In addition, the high flexural sion and to nearly full densification owing to the formation of a
strength value was similar to the values found in more complex and continuous amorphous silicate film in the material. The wear re-
expensive sintering techniques such as hot pressing and spark sistance was also evaluated in the tested processes and, according
plasma sintering. to the authors, the TSS-CW-sintered samples were considered to
An investigation with tungsten carbide and magnesium oxide exhibit excellent results. The wear rate of samples sintered via CS
nanocomposite (WC-4.3 wt% MgO) [73] was performed through was found to be 5–4  10  7 mm3 N  1 m  1, whereas TSS-CW
two-step hot-pressing sintering and compared with conventional routes showed values o2  10  7 mm3 N  1 m  1.
12566 N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572

Li et al. [76] applied three different sintering methods to nano- grains showed a size of 10 mm, most of which were smaller than
crystalline corundum abrasive: TSS-CW and CTS, based on the study 3 mm. The TSS-CW treatment resulted in significantly different and
by Li et al. [75], and CS (temperatures of 1250 and 1300 °C and a 5-h denser microstructures presenting a bimodal grain size distribu-
holding time). The smallest grain size was found in TSS-CW, be- tion, and no distinct pores could be found. The TSS-CW-processed
tween 65 and 85 nm, whereas the CS grain size ranged from 680 to samples had 98% relative density, whereas the CS-processed
800 nm and that of CTS ranged from 400 to 560 nm. Regarding the samples had 95% relative density.
investigated mechanical properties, the higher hardness and frac- Bafandeh et al. [81] investigated KNNLT powders synthesized
ture toughness values were found for the TSS-CW technique by the conventional solid-state reaction method using Na2CO3,
(22.5470.21 GPa and 3.3770.06 MPa m0.5), CTS (20.8270.22 GPa K2CO3, Li2CO3, Nb2O5, Ta2O5, SrO and TiO2 with more than 99%
and 2.9370.09 MPa m0.5) and CS (19.5570.43 GPa and purity as raw materials. Three different sintering methods were
2.8670.08 MPa m0.5), respectively. The TSS-CW-sintered samples applied: CS, TSS-CW and microwave sintering (MWS). The CS
also exhibited excellent compressive strength (over 39.2 N) and samples were sintered between 1060 and 1150 °C for 2 h. TSS-CW
wear resistance less than 1.8670.45  10  7 mm3 N  1 m  1 com- used T1 at 1060 and 1140 °C (for 10 min) and T2 at 1000 and
pared with the other methods (CS and CTS). 1040 °C for 6 and 10 h. Regarding MWS, the specimens were sin-
tered between 950 and 1050 °C (T1) for 10 min and between 850
4.9. Piezoelectric ceramics and 950 °C (T2) with 30- and 120-min holding times. By com-
paring the grain size results, it was found that TSS-CW effectively
The TSS-CW technique has also been successfully applied to inhibited grain growth compared with the conventional process.
piezoelectric ceramics. The grain size in CS was 1.8–2.2 mm, whereas TSS-CW obtained a
Fang et al. [77] investigated a typical Li- and Ta/Sb-modified grain size of 1.0–1.2 mm. However, the samples had a smaller grain
KNN composition, 0.9625(Na0.5K0.5NbO3)-0.0375Li(Ta0.4Sb0.6)O3. size (0.6–0.8 mm) in the MWS treatment. According to the authors,
The samples were treated by CS, with temperatures ranging from the high heating rate during MWS affected the sintering behavior
1075 to 1120 °C and a 4-h holding time. The TSS-CW was per- and slowed the grain growth.
formed with T1 ranging from 1100 to 1160 °C without holding Bafandeh et al. [82] investigated CS and TSS-CW sintering using
time and T2 ranging from 1000 to 1075 °C with holding times of the same material (KNN ceramics). The material was compacted at
12, 24 and 36 h. With respect to CS, the temperature increase in- 50 MPa and sintered by TSS-CW using T1 between 1090 and
itially led to microstructure densification, which was followed by 1140 °C (10-min holding time) and T2 between 1000 and 1040 °C
the formation of a porous microstructure, with abnormal grain (holding times of 6, 8 and 10 h). The authors observed that the
growth. However, TSS-CW resulted in dense microstructures with grain growth rate (ratio between the grain size of the sintered
a bimodal grain size distribution. According to the authors, the material and the particle size of the calcined powder) was ap-
increased holding time in the second sintering stage allows the proximately 8 for CS and 3 for TSS-CW. According to the authors,
pores to be eliminated, thus improving densification. The authors there is evidence of the ability of the TSS-CW method to inhibit
concluded that TSS-CW sintering is an effective way to broaden grain growth. The ceramics also showed better dielectric and
the sintering temperature range of KNN-based ceramics. piezoelectric properties compared with CS.
Li et al. [78] evaluated piezoelectric ceramics composed of Regarding the use of the TSS-CW technique in piezoelectric
0.95(Na0.52K0.44Li0.04)(Nb0.88Sb0.08Ta0.04)O3-0.05KNbO3. The TSS- ceramics, it is concluded that the method mainly helps form
CW sintering process was carried out with a heating rate of denser microstructures. In most of the evaluated cases, the longer
10 °C min  1 to T1 (1130 and 1150 °C) and a 1-min holding time. holding time in the second stage resulted in reduced pores and
Next, the specimens were cooled to T2 (which ranged from 900 to increased relative density. In addition, according to Hao et al. [80],
1040 °C) and held for 7 h. The condition with T1 at 1130 °C and T2 TSS-CW further improved the piezoelectric properties of KxNLN
at 950 °C provided higher relative density (96.5%) and a better ceramics by  10%, thus proving to be a promising technique used
piezoelectric property. The authors also concluded that the driving in the production of such materials.
force of the solution–precipitation process of the crystal particles
for liquid-phase sintering was improved by the increase in T2. This 4.10. Bioceramics
mechanism accelerated the process of liquid phase formation and
left nano-sized spots [78]. Fathi and Kharaziha [83] investigated high-purity forsterite
KNN-based ceramics ((K0.4425Na0.52Li0.0375)(Nb0.8925Sb0.07Ta0.0375)O3) nanopowders. The samples were formed by uniaxial pressing at
were prepared by the conventional mixed-oxide method [79]. The 550 MPa. TSS-CW was performed with T1 ranging from 1100 to
powders were compacted by uniaxial pressing at 300 MPa. The TSS- 1300 °C and a 6-min holding time and with T2 at 750 and 850 °C
CW was set with T1 at 900 °C (10-min holding time) and a heating rate for 2–15 h. The heating rate was 10 °C min  1, and the cooling rate
of 10 °C min  1. The specimens were then rapidly cooled to T2 (be- between T1 and T2 was 50 °C min  1. The best results were ob-
tween 980 and 1040 °C) and held for 5, 10, 15 and 20 h. The micro- tained with T1 at 1300 °C, T2 at 750 °C, and a 15-h holding time.
structures were dense in all TSS-CW-sintered samples, and the ap- Bodies with relative density of approximately 98.5% and a grain
propriate holding time increase in the second sintering stage allowed size of 60–75 nm were obtained. The use of the TSS-CW technique
elimination of the pores and improvement of the density. According to to prepare dense forsterite ceramics showed the possibility of
the authors, T1 at 1130 °C, T2 at 1020 °C and a 15-h holding time al- obtaining good mechanical properties. Compared with the hy-
lowed the samples to exhibit excellent temperature stability over a droxyapatite ceramics (HAp), which shows fracture toughness
wide temperature range, which is useful in the application of KNN- between 0.75 and 1.2 MPa m0.5 and hardness of 700 Hv, the results
based ceramics. suggested that the forsterite might be significantly improved in
With respect to piezoelectric ceramics (KxNa1  x)0.94Li0.06NbO3 terms of fracture toughness and hardness, which were
(abbreviated as KxNLN, x ¼0.34–0.61) [80], the TSS-CW sintering 3.61 MPa m0.5 and 940 Hv, respectively.
process was used with T1 at 900 °C (heating rate of 10 °C min  1 Forsterite suspensions were prepared in ethyl alcohol powder,
and 10-min holding time) and T2 at 1100 and 1050 °C with a 12-h and sponge samples were placed in the slurry and stirred. These
holding time. The CS temperature was held at 1080 °C for 2 h. The samples were removed from the slurry and sintered at different
evaluation of the micrographs showed that the microstructure TSS-CW thermal regimes. T1 ranged from 1300 to 1600 °C, and a
obtained by CS was not homogeneous and that a minority of the holding time of 5 or 60 min was used. The second stage used a
N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572 12567

cooling rate of 50 °C min  1, T2 at 800 °C and holding times of 300 holding times of 10 min (T1) and 24 h (T2), and compaction
and 900 min A microstructure with a grain size in the range of 16– pressure of 200 MPa. The best density values of 92.4 and 91.1 and
38 nm and a 25 nm mean grain size was obtained with T1 at grain sizes of approximately 136 and 162 nm were obtained for
1500 °C with a 60-min holding time and T2 at 800 °C for 300 min compositions with x ¼0.0 and 2.0, respectively. They also showed
This method provided several scaffolds with different compressive the best microhardness values: 6.76 and 7.25 GPa. The authors
strengths. Compressive strength values were 0.03–24.16 MPa, and concluded that material density decreased because of the in-
porosity was 58–88%. The authors concluded that the synthesized creased fluoride in the composition; however, it increased because
scaffolds showed a nanostructure, which is crucially important to of the longer holding time in the second sintering stage.
the bioactivity and degradability of forsterite ceramics used in The TSS-CW technique was applied to biphasic calcium phos-
biomedical applications [84]. phate (BCP) bioceramics [90]. The samples were compacted at
Another study [85] used the same procedure to prepare the 400 MPa. CS was performed at a temperature of 1200 °C for 1 h. T1
samples, and the material was processed through TSS-CW. Two at 1100 °C (1- to 30-min holding times) and T2 at 1000 and
lower temperatures were tested in the first stage: 1300 and 1050 °C for 20 h were used in TSS-CW. The heating rate was
1500 °C with a 5- or 60-min holding time and T2 at 800 °C for 300 5 °C min  1 to T1 and 50 °C min  1 between T1 and T2. A dense
and 900 min Other properties such as bioactivity and biodegrad- structure with non-uniform grains and a grain size of 1.4 mm was
ability were analyzed in this investigation to apply this material to observed in CS. The sample grain growth was suppressed by the
tissue engineering. By processing T1 at 1500 °C for 60 min and T2 TSS-CW sintering using T1 at 1100 °C (30-min holding time) and
at 800 °C with a 900-min holding time, the authors obtained T2 at 1050 °C and a 20-h holding time. A grain size of 375 nm was
samples with a grain size in the range of 18–78 nm and a mean obtained with high density. The authors concluded that another
size of 33 nm. The results demonstrated that nanostructure scaf- advantage of the TSS-CW technique was the lowest sintering
folds with a mean crystallite size of approximately 33 nm showed temperature, which enabled the material in question to obtain
suitable bioactivity. benefits pertaining to biological and mechanical behavior. Re-
Samples conformed by pressing at 200 MPa were prepared in garding the mechanical properties, the fracture toughness was
high-purity HAp materials (Ca10(PO4)6(OH)2) [86] with a mean size 1.11 MPa m0.5, and the hardness was 4.9 GPa; these values are
of 93 nm. The samples were sintered through CS and TSS-CW. CS consistent with those obtained in the application of more so-
used a temperature of 1100 °C for 60 s TSS-CW used T1 at 900 °C phisticated sintering techniques, with slightly higher values for
for 60 s and T2 at 800 °C with a 20-h holding time. The same re- fracture toughness.
lative density (98.3%) was obtained in both processes. However, The TSS-CW technique was also applied to a calcium magne-
the grain size was 190 nm in TSS-CW and 1.7 mm in CS. The frac- sium silicate bioceramic called merwinite (Ca3MgSi2O8) [91]. The
ture toughness increased by 95% in the TSS method, i.e., from 0.98 mean particle size of the material was approximately 90 nm. CS
to 1.92 MPa m0.5, and hardness increased from 7.1 to 7.8 GPa. was performed at a temperature of 1400 °C, heating rate of
Lukić et al. [87] evaluated HAp bioceramics compacted at 5 °C min  1 and a 20-h holding time. TSS-CW used T1 at 1250 and
175 MPa and thermally treated through TSS-CW. Different sinter- 1300 °C, a heating rate of 10 °C min  1 and a 5-min holding time.
ing procedures were performed, wherein T1 ranged from 850 to The cooling rate was 60 °C min  1 to 1250, 1200 or 1150 °C in T2,
1150 °C with a 5-min holding time and T2 ranged from 800 to with holding times of 5, 10 and 20 h. Samples with a mean grain
1050 °C with a 20-h holding time. The specimens were sintered in size of 633 nm and 98% relative density were obtained in TSS-CW
CS at 900 and 1200 °C for 1 and 20 h. The influence of the heating using T1 at 1300 °C and T2 at 1250 °C, with a 20-h holding time in
rate (2, 10 and 20 °C min  1) on the non-isothermal sintering be- T2. A significant increase in fracture toughness from 1.77 to
havior was also evaluated. It was found that the higher heating 2.68 MPa m0.5 occurred in the CS specimens compared with the
rate delays the densification. The difference in density became TSS-CW specimens.
evident at 800 °C, ranging from 2.1 g cm  3 for 2 °C min  1 to Zhou et al. [92] prepared porous nano-CaP ceramics through
1.9 g cm  3 for 20 °C min  1. The best results were obtained in the TSS-CW. Biphasic calcium phosphate (BCP) powder was manu-
TSS treatment with a heating rate of 2 °C min  1, T1 at 900 °C (5- factured by the H2O2 foaming method. The sintering stage con-
min holding time), T2 at 850 °C and a 20-h holding time. Samples sisted of heating at 10 °C min  1 to T1 (950 °C), a cooling rate of
with 99.2% relative density and a uniform microstructure were 30 °C min  1 and T2 at 850 °C. The holding time in T2 was 0.5 h.
obtained without uncontrolled grain growth and with few pores in The grain size of the nano-CaP ceramics ranged from 50 to 100 nm,
the grain boundaries. The mean grain size of the samples was whereas CS processed at 1100 °C with an 8-h holding time ob-
approximately 75 nm. High relative density (99.5%) was obtained tained grain sizes between 1 and 2 mm. Porous CaP ceramics with
in CS at 1200 °C with a 20-h holding time; however, the final grain nanometric grains were manufactured at the lowest sintering
size was 2.25 mm. It was possible to manufacture fully dense na- temperatures.
nostructured bioceramics by pressureless sintering at low sinter- Based on the studies reported herein, it is concluded that the
ing temperatures. Marković et al. [88] used this same TSS-CW mechanical properties of TSS-CW-processed bioceramics im-
sintering condition to consolidate nanostructured calcium phos- proved even when compared with more sophisticated sintering
phate functionally graded materials (FGMs). The mean grain size of techniques.
the FGMs was smaller than 100 nm, and according to the authors,
the functionally graded chemical and physical properties could be
obtained through powder processing and reaction in two-step 5. Miscellaneous
sintering.
The consolidation of fluoridated hydroxyapatite (FHA) pow- The two-step sintering technique has also been applied to other
ders, which had a chemical composition of Ca10(PO4)6(OH)2  xFx ceramic oxides in combination with other sintering techniques
(wherein the x values were 0.0, 0.5, 1.0, 1.5 and 2.0) was evaluated (such as microwave and spark plasma sintering) or by changing a
[89]. Uniaxial compaction was used at different pressures, from process condition. These different studies are presented in this
100 to 500 MPa. CS was held in the range of 900–1150 °C in air. T1 section.
at 975, 1000 and 1025 °C for 1, 5 and 10 min, and T2 at 850, 900 TSS was carried out in BaTiO3 under controlled atmosphere by
and 950 °C for 20, 22 and 24 h were investigated in TSS-CW. FHA using a combination of processes, including rapid-rate sintering
nanostructures were prepared using T1 at 1000 °C, T2 at 900 °C, and RCS [93]. BaTiO3 nanopowders were compacted by cold
12568 N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572

isostatic pressing at 300 MPa. The sintering process was carried 5 °C min  1 to room temperature. The thermal treatment in this
out in a precise RCS, high-temperature dilatometer equipped with study was characterized by the control of heating rates rather than
computer-aided design and a gas control system. The partial by the conventional TSS process, in which the cooling process
oxygen pressure during sintering was adjusted to a H2/N2/H2O freezes the structure. Cobalt ferrite discs with 96% relative density
ratio in a blowing gas mixture. Initially, the non-isothermal sin- were obtained by applying the sintering profile [96].
tering was evaluated up to 1300 °C, with a heating rate ranging Nanopowders of yttrium aluminum garnet (YAG) synthesized
from 200 to 3000 °C h  1 under partial oxygen pressures ranging by a modified co-precipitation method were obtained through the
from 10  1 to 10  19 atm. The purpose was to define the best TSS method. Powders with a 50 nm particle size were cold pressed
heating system. Temperatures of 1000–1200 °C with 0.5- to 20-h at 10 MPa and isostatically cold pressed at 200 MPa to 40–45%
holding times were then used in the second sintering stage, and relative density. Sintering occurred between 1500 and 1800 °C
the partial oxygen pressure remained unchanged. Samples with under vacuum (10  3 Pa). The temperatures of the first stage were
grain sizes of 108 711 nm and 99.7% relative density were ob- 1700, 1750 and 1800 °C without holding time, and those of the
tained using T1 at 1300 °C and oxygen pressure at 10  19 atm and second stage were 1500, 1550 and 1600 °C with a 10-h holding
T2 at 1100 °C with a 7-h holding time. By combining these sin- time. CS was performed at the same temperatures used in the first
tering techniques, the authors stated that the process simplified and second TSS stages to compare the obtained results. After the
control of the grains and the morphology of the pores. sample reached the desired temperature in T1, it was cooled and
Chang et al. [94] investigated TSS-CW in (Ba, Ca)(Ti, Zr)O3 held at T2 for 10 h. The results of relative density and grain size for
(BCTZ). The thermal treatment used a heating rate of 10 °C min  1 CS (corresponding to the range from 1700 to 1800 °C) were 90.0–
in T1 (from 1250 to 1300 °C) and T2 ranging from 1150 to 1200 °C. 98.1% and 2.11–5.33 mm, respectively. Regarding CS at 1500–
The cooling rate in T2 was 30 °C min  1, with an 8- to 15-h holding 1600 °C, the relative density (76.8–97.2%) and the grain size (0.65–
time. The morphological analysis of the samples showed that the 1.33 mm) were reduced. Regarding TSS, the densities were higher
sintering conditions, i.e., T1 at 1250 °C for 1 min, T2 at 1200 °C for than 99%, with a grain size ranging from 2.62 to 8.36 mm. T1 at
8 h; and T1 at 1300 °C for 1 min, T2 at 1200 °C for 8 h, resulted in a 1800 °C and T2 at 1550 °C reached 99.9% relative density with a
microstructure with grains smaller than 0.5 mm and high relative 6.04 mm grain size [97].
density (higher than 95%). The grain size obtained in the CS Y3Al5O12-YAG nanopowders were sintered by the co-pre-
samples was 5 mm. cipitation method using ammonium hydrogen carbonate as the
Balaya et al. [95] investigated dense nanocrystalline strontium precipitant [98]. The samples were pressed at 10 MPa, followed by
titanate (SrTiO3). They avoided significant grain growth by apply- isostatic pressing at 200 MPa. The TSS treatment was performed at
ing TSS using hot pressing. Isostatic pressing was applied at 1800 °C (T1) without holding time, with immediate cooling to T2
300 MPa for 5 min in addition to uniaxial hot pressing. TSS was at 1600 °C, and held for 8 h. The process occurred under vacuum
applied to the pressed samples (300 MPa), with T1 at 1220 °C and (10  3 Pa). After sintering, the samples were annealed at 1450 °C
a heating rate of 15 °C min  1 under nitrogen atmosphere and T2 for 6 h. CS was used at temperatures T1 and T2 for comparison.
at 1000 °C with a 20-h holding time. The obtained grain size was After T1 (1800 °C) was reached, the sample was cooled. T2
150 nm. T1 at 1060 °C was used in high-pressure TSS and in hot (1600 °C) was held for 8 h. The samples treated in a single stage
pressing with a heating rate of 150 °C min  1 under nitrogen at- showed density values of 98.2% and 97.1% for T1 and T2, respec-
mosphere subjected to a uniaxial pressure of 125 MPa for 1 min tively. TSS obtained 99.9% relative density and a 4 mm grain size,
The temperature of the second stage was 1020 °C with a 3-min which was lower than that obtained in the study by Chen et al.
holding time. Samples with a smaller grain size (80 nm) and 93% [97].
relative density were obtained. Ultrafine alumina powders with a mean particle size of 0.2 mm
A comparative study [37] used three different oxides to com- were also evaluated by TSS. The TSS suggested by Chu was com-
pare the TSS and SPS thermal treatments: Al2O3, 3 mol% pared with the method developed by Chen and Wang (TSS-CW).
Y2O3-stabilized ZrO2, and SrTiO3, with particle sizes of 240 nm, Temperatures from 1000 to 1050 °C with holding times of 3, 6 and
60 nm and 50 nm, respectively. The evaluation was carried out by 9 h were used regarding the methodology suggested by Chu. The
considering samples compacted at 300 MPa with a 5-min holding temperature was set at 1300 °C with a 3-h holding time in the
time. Heating rates of 10 °C min  1 to 800 °C and 5 °C min  1 to T1, second stage. Regarding the process developed by Chen and Wang,
decreasing by 60 °C min  1 to T2, were used in TSS, and T1 and T2 T1 was 1400 °C with a 5-min holding time. Two different tem-
were in the range of 1000–1400 °C with holding times of 0, 5, 10, peratures (1260 and 1300 °C) with holding times of 3, 6 and 9 h
15 and 20 h. The SPS samples were prepared under vacuum with a were evaluated in T2. Both TSS methods were effective in con-
uniaxial pressure of 200 MPa. The use of TSS did not cause sig- trolling the alumina microstructure compared with CS. The TSS-
nificant changes in the alumina and zirconia microstructures when CW method, which used T1 at 1400 °C and T2 at 1300 °C with a
the grain density and size data were compared to those of CS. 9-h holding time, showed the best relative density and grain size
Significant changes were observed in SrTiO3, wherein T1 was results, 96.58% and 460.7 nm, respectively. The holding time was
1220 °C and T2 was 1000 °C for 15 h in TSS, and the grain size was significantly important to grain growth in both techniques, and it
0.35 mm. CS used a temperature of 1250 °C for 30 min and ob- most influenced the TSS-CW technique [99].
tained a grain size of 0.56 mm, i.e., a grain size reduction of ap- Chinelatto et al. [100] evaluated the TSS processes performed
proximately 40%. The relative density was 95% in both processes. according to the methodology by Chu and the technique by Chen
However, the best result was achieved by the SPS method, with a and Wang. The used material was 5 vol% of nanometric zirconia
temperature of 875 °C for 5 min under a pressure of 200 MPa. The inclusions in alumina matrix. TSS-CW used T1 at 1450 °C (for
SrTiO3 grain size was 80 nm, and the relative density was 98%. 5 min) and T2 at 1400 °C with an 8-h holding time. Samples with a
The use of high-purity cobalt oxide (chromium) and iron oxide grain size of approximately 538 nm and 98.5% relative density
(Fe2O3) resulted in a cobalt ferrite composition (CoFe2O4), which were obtained. The Chu-based TSS used T1 at 1100 °C with a 2-h
was processed based on TSS. The specimens were heated at holding time and T2 at 1400 °C for 2 h, and it obtained a smaller
1 °C min  1 to 550 °C, at 5 °C min  1 to 950 °C, and at 1 °C min  1 to grain size (472 nm) but with lower relative density (97.1%). CS was
T1 at 1310 °C, where they remained for 1–3 h. The specimens were performed at 1500 °C for 2 h and showed a larger grain size with
then cooled at 20 °C min  1 to 1500 °C (T2) with a holding time less density, 960 nm and 96.7%, respectively. The authors stated
ranging from 30 min to 3 h. Finally, the material was cooled at that the TSS-CW technique allowed more effective densification
N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572 12569

and grain size suppression to be obtained, in addition to an in- Y2O3 samples sintered in a single stage at a rate of 10 °C min  1 to
creased microhardness value (18.10 GPa). 1490 °C with a 4-h holding time, a grain size reduction from 3.8 to
The effect of the two-step sintering process (TSS-CW) on the 0.76 mm was observed, and the Vickers hardness increased from
optical transmittance and on the mechanical strength of poly- 686 to 867 Hv and flexural strength increased from 103.5 to
crystalline alumina ceramics was evaluated [101]. Injection-based 158.5 MPa compared with TSS without holding time in T1. Owing
molding was used in a vacuum atmosphere at a high temperature to the use of a holding time (5 min) in the first sintering stage, Ti-
along with the implementation of the TSS-CW method. The first doped Y2O3 showed the best mechanical properties among the
stage was held at 1780 °C, and the second stage was held at 1300 tested samples: Vickers hardness of 874 Hv and flexural strength
and 1600 °C with different holding times, to prevent surface dif- of 307.0 MPa with densification of 99.4%. The flexural strength
fusion and the formation of secondary phases. The increased sin- increased by approximately two times more than the values re-
tering temperature and holding time in the second stage promoted ported in the literature.
a significant increase in grain size. The optical transmittance va- Wang et al. [106] extensively investigated barium-doped cerate
lues significantly increased because of grain growth. The condition ceramics with a nominal composition of BaZr0.1Ce0.7Y0.1Yb0.1O3  δ
with low temperature in the second sintering stage and a shorter (BZCYYb), owing to their high proton conductivity at intermediate
holding time showed the highest flexural strength (191 MPa) and temperatures, which can be applied to proton-conducting solid
the lowest optical transmittance (12.5%). The relative density and oxide fuel cells (SOFCs). The powder with a mean particle size of
grain size values in this condition (1780–1300 °C for 1 h) were 30 nm was uniaxially pressed at 600 MPa. The material was sin-
99.39% and 7.2 mm, respectively. However, grain sizes larger than tered by TSS-CW at 5 °C min  1 up to T1 at 1450 °C for 1 min and
those previously obtained were observed. cooled to T2 at 1300 °C with a cooling rate of 15 °C min  1 and a
Pouchly et al. [102] changed the concept from the master sin- 20-h holding time. Regarding CS, the specimens were heated at
tering curve (MSC) to the master two-step sintering curve using 5 °C min  1 to 1450 °C (5-h holding time). By controlling the sin-
four different ceramics appropriate to represent the TSS-CW pro- tering curve through TSS-CW, a material with a grain size of
cess. Thus, a tetragonal zirconia with 3 mol% Y2O3, two cubic zir- 184 nm was obtained, whereas CS obtained a grain size of 445 nm.
conia with 8 mol% Y2O3, and one alumina were used. Different In addition, the TSS-CW-sintered samples showed less impurity
heating rates (2, 5, 10 and 20 °C min  1) and CS at a temperature of and higher electrical conductivity compared with the CS samples.
1500 °C were used in all materials. With respect to TSS, T1 (1440– Zhang et al. [107] evaluated BS-PT ceramics
1328 °C), T2 (1340–1228 °C) and the holding time (10 or 15 h) var- (0.37BiScO3-0.63PbTiO3, initial particle size of 10 nm) by combin-
ied for each material under study. The application of MSC to the ing the SPS and TSS-CW techniques. The SPS heating rate ranged
TSS-CW-sintered samples demonstrated the need to divide the between 150 and 175 °C min  1 to Ts (sintering temperature) with
curves (plots) into two parts: one with low density and with the a 3-min holding time. The specimen was cooled to room tem-
same activation energy calculated for the single-step sintering ex- perature under vacuum and pressed at 50 MPa for 3 min after
periments and the other with high density and low activation en- reaching Ts. Subsequently, the samples were treated through TSS
ergy. An increase in TSS-CW efficiency was also observed because of at a heating rate of 5 °C min  1 to T1, cooled at 50 °C min  1 to T2,
the reduction in the activation energy at the final sintering stage. and held for 6 h. The treatment at 650 °C (Ts) and a heating rate of
The calculated activation energy values were 770 kJ mol  1 for 150 °C min  1, followed by T1 at 800 °C and T2 at 750 °C, produced
Al2O3, 1270 kJ mol  1 for ZrO2 with 3 mol% Y2O3, 620 kJ mol  1 and samples with a 70 nm grain size and 94.7% relative density. Ts at
750 kJ mol  1 for ZrO2 with 8 mol% Y2O3 ceramics. 700 °C and a heating rate of 175 °C min  1, followed by T1 at
Lee and Kim [103] studied aluminum nitride (ALN) with cal- 800 °C and T2 at 750 °C, obtained a grain size of 33 nm and re-
cium zirconate (CaZrO3) and Y2O3 added. The specimens were lative density of 96.6%. With respect to the sample treated with Ts
treated through TSS, with temperatures between 1450 and at 600 °C, a heating rate of 150 °C min  1, T1 at 900 °C and T2 at
1700 °C. The process occurred under nitrogen atmosphere in a 750 °C, the grain size and relative density were lower at 23 nm and
graphite furnace. The heating rate was 10 °C min  1 to T1 (3-h 95.7%, respectively. The authors concluded that the grain size in-
holding time), and the cooling rate was 25 °C min  1 to T2 for 2 h. creased with increasing Ts, and it decreased according to the
Limited grain growth was verified owing to the use of the two-step heating rate in the SPS process.
process. The highest thermal conductivity was achieved by the Commercial powders of Al2O3, Y2O3 and Nd2O3 (99% purity)
two-step sintering, and the ceramics showed moderate flexural were used to obtain Nd-doped YAG ceramics produced by a solid-
strength (560 MPa). state reaction [108]. Compaction was performed by uniaxial
Hejazi et al. [104] applied TSS-CW to prepare Mg-substituted pressing at 10 MPa and cold isostatic pressing at 200 MPa. The
fluorapatite and MFA/alumina composites using different amounts powders were thermally treated in TSS with T1 at 1800 °C and a
of alumina. T1 at 1050 or 1200 °C (1-min holding time) and T2 at heating rate of 10 °C min  1. T2 ranged between 1550 and 1750 °C
950 or 1100 °C (12- and 20-h holding times) were used. A com- (1- to 8-h holding time) in a vacuum graphite tube furnace under
posite with a fine particle size was obtained using the TSS-CW 10  3 Pa during the holding process. Samples sintered at 1680 °C
process. However, an increase in particle size from 175 nm to (8-h holding time) showed a dense structure, pore-free micro-
590 nm was observed after alumina was introduced into the structure and excellent optical properties.
composite, and it improved the mechanical properties. The TSS process was applied to yttrium and gadolinium-doped
Khosroshahi et al. [105] used a modified TSS process for cation ceria-based electrolytes (20 at% dopant cation) (YCO and GCO)
(Si4 þ , Er3 þ , Yb3 þ , Al3 þ or Ti4 þ )-doped yttria and molar dopant: with and without small Ga2O3- additions (0.5 mol%). The treat-
yttria ratio of 0.01. The samples were compacted by uniaxial ment used T1 of 1250–1300 °C and T2 at 1150 °C with a holding
pressing at 5.3 MPa, followed by cold isostatic pressing at 350 MPa time of 0 or 10 h. The GCO samples reached 93.4% relative density,
for 3 min The SST was performed at a heating rate of 10 °C min  1 using T1 at 1300 °C and T2 at 1150 °C with a 10-h holding time.
to T1 at 1490 °C, when it was cooled to room temperature. Next, YCO using T1 at 1300 °C and T2 at 1150 °C with a 10-h holding
the specimens were reheated at a rate of 20 °C min  1 to T2 at time reached slightly higher relative density (94.3%). Compacts
1350 °C with a 20-h holding time. Regarding Ti-doped Y2O3, Ti with a submicrometric grain size ranging from 150 to 250 nm
showed the best results in terms of densification and mechanical were produced [109].
properties; thus, an optimized TSS with holding times of 0, 1 and The 0.89Bi0.5Na0.5TiO3-0.06BaTiO3-0.05 K0.5Na0.5NbO3 ternary
5 min was used in the first stage. With respect to the Ti-doped system ceramics derived from Bi2O3, Na2CO3, K2CO3, BaCO3, TiO2
12570 N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572

and Nb2O5 with more than 99% purity were treated through TSS- isostatically pressed at 1500 kg cm  2. A CS heating rate of
CW and CS. The samples were compacted by pressing at 40 MPa. 3 °C min  1 was applied to reach the sintering temperature of
The temperature in CS was 1140 °C with a 3-h holding time. TSS- 1300–1400 °C (4-h holding time). The same heating rate was used
CW used T1 at 1140 °C (no holding time) and T2 at 1000 °C with a in the TSS-CW process up to T1 at 1300–1400 °C (5-min holding
5-h holding time. The cooling rate was 20 °C min  1. The TSS-CW- time). The cooling rate was 25 °C min  1 to 1250 °C (T2) with
sintered samples showed a homogeneous microstructure without holding times of 4, 8 and 12 h. The TSS-CW-sintered samples (T1
abnormal grain growth and with mean size of 1 mm. The grain at 1350/1400 °C and T2 at 1250 °C for 4 h) showed increased mi-
sizes were abnormal and ranged from 1.5 to 3.0 mm in CS [110]. crostructural homogeneity and higher relative densities (higher
Zhou et al. [111] evaluated nano-sized Nd3 þ :Lu2O3 powders than 96%) compared with the CS samples.
with an initial particle size of 40 nm. The thermal treatment was Hosseini et al. [117] evaluated nanostructured forsterite coat-
performed through TSS-CW, and it used T1 at 1720 °C and T2 at ing. The TSS-CW sintering process was used in argon atmosphere
1620 °C with a 10-h holding time. H2 atmosphere was used during with T1 at 1100 °C and a 30-min holding time and with T2 at
the process. The grain size was approximately 400 nm, which was 900 °C for 120 min The heating and cooling rates were 5 °C min  1.
approximately 150 times smaller than that obtained by CS [112]. The novel nanostructured forsterite coating was successfully de-
Khakpour et al. [113] investigated samarium-doped CeO2 (SDC) posited onto the 316 L stainless steel substrate by the electro-
compounds. Two different compositions were prepared: ceria phoretic deposition (EPD) method, followed by TSS.
doped with 20 (2SDC) and 30 (3SDC) mol% Sm. The powder was
compacted at 500 MPa and thermally treated through TSS-CW and
CS. TSS-CW used T1 at 1450 °C and a heating rate of 10 °C min  1. 6. Conclusions
The cooling rate for T2 was 50 °C min  1 to 1300 °C, at which it was
held for 20 h. CS used temperatures of 1400 °C (2SDC) and 1450 °C The sintering of ceramic materials by sintering curve control is
(3SDC) with a heating rate of 2 °C min  1. The holding times were an effective, simple and economical microstructure refinement
5 h for 3SDC and 8 h for 2SDC. A grain size and relative density of method. Thus, two-step sintering (TSS) is a promising method
1.7 mm and 95%, respectively, were obtained for the 2SDC sample used to obtain high-density bodies and smaller grain sizes. Two
processed through CS, whereas the TSS-CW grain size and relative TSS methodologies are known: the sintering with thermal pre-
density values were 1.15 mm and 99.6%, and they were significantly treatment at a low sintering temperature, followed by a second
better. The 3SDC composition obtained a grain size and relative stage at an elevated temperature, and the more recent approach
density of 4.5 mm and 97%, respectively, through CS. However, TSS- presented by Chen and Wang, which has been most widely used.
CW achieved a grain size reduction to 2.13 mm (with 98% relative Among the studies identified in the literature, the vast majority
density). Again, the effectiveness of the TSS-CW technique in have employed the technique developed by Chen and Wang by
suppressing grain growth was demonstrated. adjusting the parameters to each study system.
The TSS technique may also be useful to control the pore size. The sintering specifications, including the temperatures, heat-
The pore size and porosity of porous ceramics depend on the par- ing/cooling rates, sintering holding times and atmosphere types,
ticle size and the green body porosity. Therefore, the sintering plan were determined for each study to obtain the best features for
is an important factor that should be considered in the preparation each material. Most studies reported the obtainment of higher
of porous ceramics. Isobe et al. [114] used 30 vol% α-Al2O3 with relative density through the two-step sintering technique than
70 vol% distilled water and 0.62 wt% poly (ammonium acrylate) to that obtained through conventional sintering, thus inhibiting the
prepare suspensions and shaped them through slip-casting. The CS accelerated grain growth. Nanometric structures were obtained in
thermal treatment used temperatures of 1000–1400 °C for 2 h with some studies. Some general conclusions should be highlighted in
a heating rate of 1.5 °C min  1. The TSS-CW used a heating rate of the current study:
100 °C min  1 to T1 (1000–1300 °C) and T2 at 850–1150 °C with
holding times between 2 and 12 h. The cooling rates were 1. specific particularities were observed in each material owing to
15 °C min  1 from T1 to T2 and 1.5 °C min  1 from T2 to room the addition of dopant agents and the effects of the two-step
temperature. One of the results showed that when T1 is sufficiently sintering technique. Regarding the yttrium, zinc, barium tita-
high, the pore size and porosity do not change. The pores showed a nate and alumina oxides, it was possible to see that dopant
mean size of 61–76 nm for T1 at 1150 °C and T2 at 1000 °C. agents inhibit grain growth;
Li et al. [115] investigated the behavior of a MgO microstructure 2. the initial powder features, such as particle size, molding pro-
with 99.9% purity and a 50 nm particle size. The powder was cessing, microstructural homogeneity, and amount of green
compacted at 6 MPa for 3–4 min The sintering process was carried body pores are factors that significantly influence the success of
out in tube furnace using the TSS-CW and CS processes. CS was two-step sintering, in addition to alumina and yttria-stabilized
performed between 1000 and 1600 °C, with heating and cooling zirconia;
rates of 5 °C min  1. Two holding times were used, 5 min and 2 h, 3. specifically regarding the two-step sintering developed by Chen
and the samples were designated as C1 and C2, respectively. TSS- and Wang (TSS-CW), the temperature reduction of the first
CW used a heating rate of 5 °C min  1 to T1 (5-min holding time). stage (T1) and the second stage (T2), despite the longer holding
The cooling rate was 8 °C min  1 to T2, which was held for 20 h. time in T2, allows smaller grain sizes to be obtained for the
Based on the results obtained for the C1 and C2 samples, the TSS- same relative density compared with conventional sintering;
CW temperatures were determined to optimize the results. T1 was 4. the grain size reduction for the same densification allows ma-
chosen as 1550 °C and T2 as 1400 °C owing to high grain growth at terials with improved properties to be obtained, especially in
temperatures above 1410 °C. The use of TSS-CW produced samples terms of mechanical properties. Two-step sintering indirectly
with high relative density (98%), a 21 mm grain size, and a mean improves these properties;
hardness of 6.1 GPa. Regarding CS, a lower relative density (96.2%) 5. the TSS method is an effective microstructure control method
was achieved for a grain size of 23 mm. Thus, MgO ceramics with compared with more expensive processes, such as spark plasma
greater densification were obtained through TSS-CW without sintering (SPS). However, techniques such as SPS and micro-
using additives in the sintering. wave sintering are more effective for some materials, such as
Eoh et al. [116] used the TSS-CW process to investigate the yttria-stabilized zirconia. The cost/benefit ratio of the material
microstructural features of Zn1.8SiO3.8 ceramics. The powders were under study must be evaluated.
N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572 12571

Acknowledgements (1995) 9–16.


[29] B.-N. Kim, T. Kishi, Strengthening mechanism of alumina ceramics prepared
by precoarsening treatments, Mater. Sci. Eng. A 215 (1) (1996) 18–25.
The authors are very grateful to Coordenação de Aperfeiçoa- [30] Y. Wu, C. Nguyen, S. Seraji, M.J. Forbess, S.J. Limmer, T. Chou, G. Cao, Pro-
mento de Pessoal de Nível Superior (CAPES/Brazil, Edital PRÓ- cessing and properties of strontium bismuth vanadate niobate ferroelectric
DEFESA n. 031/2013, Auxílio n. 031/2013) for funding the current ceramics, J. Am. Ceram. Soc. 84 (12) (2001) 2882–2888.
[31] D.-S. Kim, J.-H. Lee, R.J. Sung, S.W. Kim, H.S. Kim, J.S. Park, Improvement of
study. traslucency in Al2O3 ceramics by two step sintering technique, J. Eur. Ceram.
Soc. 27 (13) (2007) 3629–3632.
[32] M. Han, X. Tang, H. Yin, S. Peng, Fabrication, microstructure and properties of
a YSZ electrolyte for SOFCs, J. Power Sources 165 (2) (2007) 757–763.
References [33] S.-P. Lee, B. Hwang, Y.-K. Paek, T.-J. Chung, S.-H. Yang, S. Lim, W.-S. Seo, K.-
S. Oh, Determination of the initial heat treatment temperature in the two
step sintering of xAl–ZnO (x¼ 1, 2, and 3 wt%), J. Eur. Ceram. Soc. 33 (1)
[1] F.J.T. Lin, L.C. De Jonghe, Microstructure refinement of sintered alumina by a
(2013) 131–137.
two step sintering technique, J. Am. Ceram. Soc. 80 (9) (1997) 2269–2277.
[34] L. Liu, F. Gao, Y. Zhang, H. Sun, Dense KSr2Nb5O15 ceramics with uniform
[2] E.A. Barringer, H.K. Bowen, Formation, packing, and sintering of mono-
grain size prepared by molten salt synthesis, J. Alloy. Compd. 616 (2014)
dispersed TiO2 powders, J. Am. Ceram. Soc. 65 (12) (1982) 199–201.
293–299.
[3] T.-S. Yeh, M.D. Sacks, Low-temperature sintering of aluminum oxide, J. Am.
[35] M. Mazaheri, A.M. Zahedi, S.K. Sadrnezhaad, Two step sintering of nano-
Ceram. Soc. 71 (10) (1988) 841–844.
crystalline ZnO compacts: effect of temperature on densification and grain
[4] X. Kuang, G. Carotenuto, L. Nicolais, A review of ceramic sintering and sug-
growth, J. Am. Ceram. Soc. 91 (1) (2008) 56–63.
gestions on reducing sintering temperatures, Adv. Perform. Mater. 4 (3)
[36] M.-W. Wu, Two step sintering of aluminum-doped zinc oxide sputtering
(1997) 257–274.
target by using submicrometer zinc oxide powder, Ceram. Int. 38 (8) (2012)
[5] S.C. Liao, Y.J. Chen, B.H. Kear, W.E. Mayo, High pressure/low temperature
6229–6234.
sintering of nanocrystalline alumina, Nanostruct. Mater. 10 (6) (1998)
[37] K. Maca, V. Pouchly, Z. Shen, Two step sintering and spark plasma sintering
1063–1079.
of Al2O3, ZrO2 and SrTiO3 ceramics, Integr. Ferroelectr. 99 (1) (2008) 114–124.
[6] Z. He, J. Ma, Grain growth rate constant of hot-pressed alumina ceramics,
[38] K. Maca, V. Pouchly, P. Zalud, Two step Sintering of oxide ceramics with
Mater. Lett. 44 (1) (2000) 14–18. various crystal structures, J. Eur. Ceram. Soc. 30 (2) (2010) 583–589.
[7] L. Gao, J.S. Hong, H. Miyamoto, D.D.L. Torre, Bending strength and micro- [39] P. Durán, F. Capel, J. Tartaj, C. Moure, A strategic two-stage low-temperature
structure of Al2O3 ceramics densified by spark plasma sintering, J. Eur. Cer- thermal processing leading to fully dense and fine-grained doped-ZnO var-
am. Soc. 20 (12) (2000) 2149–2152. istors, Adv. Mater. 14 (2) (2002) 137–141.
[8] Y. Zhou, K. Hirao, Y. Yamauchi, S. Kanzaki, Densification and grain growth in [40] P. Durán, J. Tartaj, C. Moure, Fully dense, fine-grained, doped zinc oxide
pulse electric current sintering of alumina, J. Eur. Ceram. Soc. 24 (12) (2004) varistors with improved nonlinear properties by thermal processing opti-
3465–3470. mization, J. Am. Ceram. Soc. 86 (8) (2003) 1326–1329.
[9] M.J. Mayo, Processing of nanocrystalline ceramics from ultrafine particles, [41] M.M. Shahraki, S.A. Shojaee, M.A.F. Sani, A. Nemati, I. Safaee, Two step sin-
Int. Mater. Rev. 41 (3) (1996) 85–116. tering of ZnO varistors, Solid State Ion. 190 (1) (2011) 99–105.
[10] M.-Y. Chu, L.C. De Jonghe, M.K.F. Lin, F.J.T. Lin, Precoarsening to improve [42] J. Li, Y. Ye, Densification and grain growth of Al2O3 nanoceramics during
microstructure and sintering and sintering of powder compacts, J. Am. Cer- pressureless sintering, J. Am. Ceram. Soc. 89 (1) (2006) 139–143.
am. Soc. 74 (11) (1991) 2902–2911. [43] K. Bodišová, P. Šajgalík, D. Galusek, P. Švančárek, Two-stage sintering of
[11] I.W. Chen, X.H. Wang, Sintering dense nanocrystalline ceramics without fi- alumina with submicrometer grain size, J. Am. Ceram. Soc. 90 (1) (2007)
nal-stage grain growth, Nature 404 (6774) (2000) 168–171. 330–332.
[12] X.-H. Wang, P.-L. Chen, I.-W. Chen, Two step sintering of ceramics with [44] Z.R. Hesabi, M. Haghighatzadeh, M. Mazaheri, D. Galusek, S.K. Sadrnezhaad,
constant grain-size, I. Y2O3, J. Am. Ceram. Soc. 89 (2) (2006) 431–437. Suppression of grain growth in sub-micrometer alumina via two step sin-
[13] H. Palmour III, D.R. Johnson, Phenomenological model for rate controlled tering method, J. Eur. Ceram. Soc. 29 (8) (2009) 1371–1377.
sintering, in: G.C. Kutzynski, et al., (Eds.), Sintering and Related Phenomena, [45] C.-J. Wang, C.-Y. Huang, Y.-C. Wu, Two step sintering of fine alumina–zirconia
Gordon and Breach, New York, 1967, pp. 779–791. ceramics, Ceram. Int. 35 (4) (2009) 1467–1472.
[14] M.L. Huckabee, H. Palmour III, Rate controlled sintering of fine grained alu- [46] D. Galusek, K. Ghillányová, J. Sedláček, J. Kozánková, P. Šajgalík, The influence
mina, Am. Ceram. Soc. Bull. 51 (7) (1972) 574–576. of additives on microstrucutre of sub-micron alumina ceramics prepared by
[15] H. Palmour III, M.L. Huckabee, T.M. Hare, Microstructural development dur- two-stage sintering, J. Eur. Ceram. Soc. 32 (9) (2012) 1965–1970.
ing optimized rate controlled sintering, in: R.M. Fulrath, J.A. Pack (Eds.), [47] S.-Y. Lee, Sintering behavior and mechanical properties of injection-molded
Ceramic Microstructure, 76, Westview Press, Boulder, Colorado, 1977, zirconia powder, Ceram. Int. 30 (4) (2004) 579–584.
pp. 308–319. [48] M. Mazaheri, A. Simchi, F. Golestani-Fard, Densification and grain growth of
[16] M.L. Huckabee, T.M. Hare, H. Palmour III, Rate controlled sintering as a nanocrystalline 3Y-TZP during two step sintering, J. Eur. Ceram. Soc. 28 (15)
processing method, in: H. Palmour III, R.F. Davis, T.M. Hare (Eds.), Processing (2008) 2933–2939.
of crystalline ceramics, Material Science Research, Plenum Press, New York, [49] G. Suárez, Y. Sakka, T.S. Suzuki, T. Uchikoshi, E.F. Aglietti, Effect of bead-
1978, pp. 205–215. milling treatment on the dispersion of tetragonal zirconia nanopowder and
[17] H. Palmour III, M.L. Huckabee, T.M. Hare, Rate controlled sintering: principles improvements of two step sintering, J. Ceram. Soc. Jpn. 117 (1364) (2009)
and practice, in: M.M. Ristic (Ed.), Sintering-New Developments, Elsevier Sci. 470–474.
Publ. Co., Amsterdam, 1979, pp. 46–56. [50] G. Suárez, Y. Sakka, T.S. Suzuki, T. Uchikoshi, X. Zhu, E.F. Aglietti, Effect of
[18] H. Palmour III, T.M. Hare, Rate controlled sintering revisited, in: Sintering’85, starting powders on the sintering of nanostructured ZrO2 ceramics by col-
Plenum Press, New York, 1987, pp. 17–34. loidal processing, Sci. Technol. Adv. Mater. 10 (2) (2009) 1–8.
[19] H. Palmour III, Rate controlled sintering of a whiteware porcelain, Am. Cer- [51] Y. Xiong, J. Hu, Z. Shen, Dynamic pore coalescence in nanoceramic con-
am. Soc. 7 (11–12) (1986) 1203–1212. solidated by two step sintering procedure, J. Eur. Ceram. Soc. 33 (11) (2013)
[20] H. Palmour III, Rate controlled sintering for ceramics and selected powder 2087–2092.
metals, in: D.P. Uskokovic, et al., (Eds.), Science of Sintering, Plenum Press, [52] A. Ghosh, A.K. Suri, B.T. Rao, T.R. Ramamohan, Low-temperature sintering
New York, 1989, pp. 337–356. and mechanical property evaluation of nanocrystalline 8 mol% yttria fully
[21] A.V. Ragulya, V.V. Skorokhod, Rate-controlled sintering of ultrafine nickel stabilized zirconia, J. Am. Ceram. Soc. 90 (7) (2007) 2015–2023.
powder, Nanostruct. Mater. 5 (7–8) (1995) 835–843. [53] M.A.M. Lourenço, Aplicabilidade da Sinterização em Duas Etapas na Obten-
[22] G. Agarwal, R.F. Speyer, W.S. Hackenberger, Microstructural development of ção de Cerâmicos com Tamanho de Grão Submicrométrico, Universidade De
ZnO using a rate-controlled sintering dilatometer, J. Mater. Res. 11 (3) (1996) Aveiro, Aveiro, Portugal, 2009.
671–679. [54] M.A. Lourenço, G.G. Cunto, F.M. Figueiredo, J.R. Frade, Model of two step
[23] A.V. Ragulya, Rate-controlled synthesis and sintering of nanocrystalline sintering conditions for yttria-substituted zirconia powders, Mater. Chem.
barium titanate powder, Nanostruct. Mater. 10 (3) (1998) 349–355. Phys. 126 (1) (2011) 262–271.
[24] O.B. Zgalat-Lozinskii, V.N. Bulanov, I.I. Timofeeva, A.V. Ragulya, V. [55] M. Mazaheri, M. Valefi, Z.R. Hesabi, S.K. Sadrnezhaad, Two step sintering of
V. Skorokhod, Sintering of refractory compounds nanocrystalline powders.II. nanocrystalline 8Y2O3 stabilized ZrO2 synthesized by glycine nitrate process,
Non-isothermal sintering of titanium nitride powder, Powder Metall. Met. Ceram. Int. 35 (1) (2009) 13–20.
Ceram. 40 (11–12) (2001) 573–581. [56] M. Mazaheri, A.M. Zahedi, M.M. Hejazi, Processing of nanocrystalline 8 mol%
[25] I.W. Chen, Grain boundary kinetics in oxide ceramics with the cubic fluorite yttria-stabilized zirconia by conventional, microwave-assisted and two step
crystal structure and its derivatives, Interface Sci. 8 (2–3) (2000) 147–156. sintering, Mater. Sci. Eng. A 492 (1) (2008) 261–267.
[26] J.-L. Huang, L.-M. Din, H.-H. Lu, W.-H. Chan, Effects of two step sintering on [57] M. Mazaheri, Z.R. Hesabi, F. Golestani-Fard, S. Mollazadeh, S. Jafari, S.
the microstructure of Si3N4, Ceram. Int. 22 (2) (1996) 131–136. K. Sadrnezhaad, The effect of conformation method and sintering technique
[27] F.J.T. Lin, L.C. De Jonghe, M.N. Rahaman, Initial coarsening and micro- on the densification and grain growth of nanocrystalline 8 mol% yttria-sta-
structural evolution of fast-fired and MgO-doped Al2O3, J. Am. Ceram. Soc. 80 bilized zirconia, J. Am. Ceram. Soc. 92 (5) (2009) 990–995.
(11) (1997) 2891–2896. [58] S. Schwarz, O. Guillon, Two step sintering of cubic yttria stabilized zirconia
[28] E. Sato, C. Carry, Effect of powder granulometry and pre-treatment on sin- using field assisted sintering technique/spark plasma sintering, J. Eur. Ceram.
tering behavior of submicron-grained α-alumina, J. Eur. Ceram. Soc. 15 (1) Soc. 33 (4) (2013) 637–641.
12572 N.J. Lóh et al. / Ceramics International 42 (2016) 12556–12572

[59] M. Mazaheri, Z.R. Hesabi, S.K. Sadrnezhaad, Two step sintering of titânia [89] M. Esnaashary, M. Fathi, M. Ahmadian, The effect of the two step sintering
nanoceramics assisted by anatase-to-rutile phase transformation, Scr. Mater. process on consolidation of fluoridated hydroxyapatite and its mechanical
59 (2) (2008) 139–142. properties and bioactivity, Int. J. Appl. Ceram. Technol. 11 (1) (2013) 47–56.
[60] M. Mazaheri, A.M. Zahedi, M. Haghighatzadeh, S.K. Sadrnezhaad, Sintering of [90] M. Lukić, Z. Stojanović, S.D. Škapin, M. Maček-Kržmanc, M. Mitrić,
titania nanoceramic: densification and grain growth, Ceram. Int. 35 (2) S. Marković, D. Uskoković, Dense fine-grained biphasic calcium phosphate
(2009) 685–691. (BCP) bioceramics designed by two step sintering, J. Eur. Ceram. Soc. 31 (1)
[61] Y.I. Lee, J.-H. Lee, S.-H. Hong, D.-Y. Kim, Preparation of nanostructured TiO2 (2011) 19–27.
ceramics by spark plasma sintering, Mater. Res. Bull. 38 (6) (2003) 925–930. [91] A. Nadernezhad, F. Moztarzadeh, M. Hafezi, H. Barzegar-Bafrooei, Two step
[62] B.-R. Li, D.-Y. Liu, J.-J. Liu, S.-X. Hou, Z.-W. Yang, Two step sintering assisted sintering of a novel calcium magnesium silicate bioceramic: sintering para-
consolidation of bulk titania nano-ceramics by spark plasma sintering, Cer- meters and mechanical characterization, J. Eur. Ceram. Soc. 34 (15) (2014)
am. Int. 38 (5) (2012) 3693–3699. 4001–4009.
[63] H.T. Kim, Y.H. Han, Sintering of nanocrystalline BaTiO3, Ceram. Int. 30 (7) [92] C. Zhou, P. Xie, Y. Chen, Y. Fan, Y. Tan, X. Zhang, Synthesis, sintering and
(2004) 1719–1723. characterization of porous nano-structured CaP bioceramics prepared by a
[64] X.-H. Wang, X.-Y. Deng, H.-L. Bai, H. Zhou, W.-G. Qu, L.-T. Li, I.-W. Chen, Two two step sintering method, Ceram. Int. 41 (3) (2015) 4696–4705.
step sintering of ceramics with constant grain-size, II: BaTiO3 and Ni–Cu–Zn [93] A. Polotai, K. Breece, E. Dickey, C. Randall, A. Ragulya, A novel approach to
Ferrite, J. Am. Ceram. Soc. 89 (2) (2006) 438–443. sintering nanocrystalline barium titanate ceramics, J. Am. Ceram. Soc. 88 (11)
[65] T. Karaki, K. Yan, M. Adachi, Barium titanate piezoelectric ceramics manu- (2005) 3008–3012.
factured by two step sintering, Jpn. J. Appl. Phys. 46 (10S) (2007) [94] C.-Y. Chang, H.-I. Ho, T.-Y. Hsieh, C.-Y. Huangn, Y.-C. Wu, Effects of extra Ba2 þ
S7035–S7038. and two step sintering on the crystal structure, microstructure, and dielectric
[66] X.-H. Wang, X.-Y. Deng, H. Zhou, L.-T. Li, I.-W. Chen, Bulk dense nanocrys- properties of (Ba, Ca)(Ti, Zr)O3, Ceram. Int. 39 (7) (2013) 8245–8251.
talline BaTiO3 ceramics prepared by novel pressureless two step sintering [95] P. Balaya, M. Ahrens, L. Kienle, J. Maier, B. Rahmati, S.B. Lee, W. Sigle,
method, J. Electroceram. 21 (1–4) (2008) 230–233. A. Pashkin, C. Kuntscher, M. Dressel, Synthesis and characterization of na-
[67] Y. Huan, X. Wang, J. Fang, L. Li, Grain size effects on piezoelectric properties nocrystalline SrTiO3, J. Am. Ceram. Soc. 89 (9) (2006) 2804–2811.
and domain structure of BaTiO3 ceramics prepared by two step sintering, J. [96] A. Rafferty, T. Prescott, D. Brabazon, Sintering behaviour of cobalt ferrite
Am. Ceram. Soc. 96 (11) (2013) 3369–3371. ceramic, Ceram. Int. 34 (1) (2008) 15–21.
[68] Y. Huan, X. Wang, J. Fang, L. Li, Grain size effect on piezoelectric and ferro- [97] Z.-H. Chen, J.-T. Li, J.-J. Xu, Z.-G. Hu, Fabrication of YAG transparent ceramics
electric properties of BaTiO3 ceramics, J. Eur. Ceram. Soc. 34 (5) (2014) by two step sintering, Ceram. Int. 34 (7) (2008) 1709–1712.
1445–1448. [98] X. Li, B. Zheng, T. Odoom-Wubah, J. Huang, Co-precipitation synthesis and
[69] Y.-I. Lee, Y.-W. Kim, M. Mitomo, D.-Y. Kim, Fabrication of dense nanos- two step sintering of YAG powders for transparent ceramics, Ceram. Int. 39
tructured silicon carbide ceramics through two step sintering, J. Am. Ceram. (7) (2013) 7983–7988.
Soc. 86 (10) (2003) 1803–1805. [99] A.M.D. Souza, Sinterização em Euas Etapas Duas Etapas de Pós Ultra Finos de
[70] Y.-I. Lee, Y.-W. Kim, M. Mitomo, Effect of processing on densification of na- Alumina, Universidade Estadual De Ponta Grossa, Ponta Grossa, Brasil, 2011.
nostructured SiC ceramics fabricated by two step sintering, J. Mater. Sci. 39 [100] A.S.A. Chinelatto, A.L. Chinelatto, C.L. Ojaimia, J.A. Ferreira, E.M.J.A. Pallone,
(11) (2004) 3801–3803. Effect of sintering curves on the microstructure of alumina–zirconia nano-
[71] G. Magnani, A. Brentari, E. Burresi, G. Raiteri, Pressureless sintered silicon
composites, Ceram. Int. 40 (9) (2014) 14669–14676.
carbide with enhanced mechanical properties obtained by the two step
[101] H.S. Kim, S.T. Oh, Y.D. Kim, Effects of the two step sintering process on the
sintering method, Ceram. Int. 40 (1) (2014) 1759–1763.
optical transmittance and mechanical strength of polycrystalline alumina
[72] G. Magnani, G. Sico, A. Brentari, P. Fabbri, Solid-state pressureless sintering of
ceramics, Ceram. Int. 40 (9) (2014) 14471–14475.
silicon carbide below 2000 °C, J. Eur. Ceram. Soc. 34 (15) (2014) 4095–4098.
[102] V. Pouchly, K. Maca, Z. Shen, Two-stage master sintering curve applied to two
[73] J. Ma, S. Zhu, C. Ouyang, Two step hot-pressing sintering of nanocomposite
step sintering of oxide ceramics, J. Eur. Ceram. Soc. 33 (12) (2013) 2275–2283.
WC-MgO compacts, J. Eur. Ceram. Soc. 31 (10) (2011) 1927–1935.
[103] H.M. Lee, D.K. Kim, High-strength AlN ceramics by low-temperature sintering
[74] H. Qu, S. Zhu, Two step hot pressing sintering of dense fine grained WC-Al2O3
with CaZrO3–Y2O3 co-additives, J. Eur. Ceram. Soc. 34 (15) (2014) 3627–3633.
composites, Ceram. Int. 39 (5) (2013) 5415–5425.
[104] M.S. Hejazi, M. Ahmadian, M. Meratian, M.H. Fathi, Effect of alumina contents
[75] Z. Li, Z. Li, A. Zhang, Y. Zhu, Two step sintering behavior of sol–gel derived
on phase stability and mechanical properties of magnesium fluorapatite/
nanocrystalline corundum abrasive with MgO–CaO–SiO2 additions, J. Sol-Gel
alumina composites, J. Mech. Behav. Biomed. Mater. 40 (2014) 95–101.
Sci. Technol. 48 (3) (2008) 283–288.
[105] H.R. Khosroshahi, H. Ikeda, K. Yamada, N. Saito, K. Kaneko, K. Hayashi,
[76] Z. Li, Z. Li, A. Zhang, Y. Zhu, Synthesis and two step sintering behavior of sol–
K. Nakashima, Effect of cation doping on mechanical properties of yttria
gel derived nanocrystalline corundum abrasives, J. Eur. Ceram. Soc. 29 (8)
prepared by an optimized two step sintering process, J. Am. Ceram. Soc. 95
(2009) 1337–1345.
(10) (2012) 3263–3269.
[77] J. Fang, X.H. Wang, Z.B. Tian, C.F. Zhong, L.T. Li, R. Zuo, Two step sintering: an
[106] S. Wang, L. Zhang, L. Zhang, K. Brinkman, F. Chen, Two step sintering of ul-
approach to broaden the sintering temperature range of alkaline niobate-
trafine-grained barium cerate proton conducting ceramics, Electrochim. Acta
based lead-free piezoceramics, J. Am. Ceram. Soc. 93 (11) (2010) 3552–3555.
[78] Y. Li, Y.-J. Dai, H.-Q. Wanga, T. Sun, W. Xu, X.-W. Zhang, Microstructures and 87 (2013) 194–200.
electrical properties of KNbO3 doped (Li,Ta,Sb) modified (K,Na)NbO3 lead- [107] S. Zhang, X. Wang, H. Wang, L. Li, Grain boundary region and local piezo-
free ceramics by two step sintering, Mater. Lett. 89 (2012) 70–73. electric response of BiScO3–PbTiO3 nanoceramics prepared by combination of
[79] X. Pang, J. Qiu, K. Zhu, J. Du, (K, Na)NbO3-based lead-free piezoelectric SPS and two step sintering, J. Eur. Ceram. Soc. 34 (10) (2014) 2317–2323.
ceramics manufactured by two step sintering, Ceram. Int. 38 (3) (2012) [108] J. Li, Q. Chen, G. Feng, W. Wu, D. Xiao, J. Zhu, Optical properties of the
2521–2527. polycrystalline transparent Nd:YAG ceramics prepared by two step sintering,
[80] J. Hao, W. Bai, B. Shen, J. Zhai, Improved piezoelectric properties of Ceram. Int. 38 (2012) S649–S652.
(KxNa1  x)0.94Li0.06NbO3 lead-free ceramics fabricated by combining two step [109] C.M. Lapa, D.P.F. de Souza, F.M.L. Figueiredo, F.M.B. Marques, Two step sin-
sintering, J. Alloy. Compd. 534 (2012) 13–19. tering ceria-based electrolytes, Int. J. Hydrog. Energy 35 (7) (2010)
[81] M.R. Bafandeh, R. Gharahkhani, J.-S. Lee, Comparison of sintering behavior 2737–2741.
and piezoelectric properties of (K,Na)NbO3-based ceramics sintered in con- [110] J. Ding, Y. Liu, Y. Lu, H. Qian, H. Gao, H. Chen, C. Ma, Enhanced energy-storage
ventional and microwave furnace, Mater. Chem. Phys. 143 (3) (2014) properties of 0.89Bi0.5Na0.5TiO3–0.06BaTiO3–0.05K0.5Na0.5NbO3 lead-free
1289–1295. anti-ferroelectric ceramics by two step sintering method, Mater. Lett. 114
[82] M.R. Bafandeh, R. Gharahkhani, J.-S. Lee, Sintering behavior, dielectric and (2014) 107–110.
piezoelectric properties of sodium potassium niobate-based ceramics pre- [111] D. Zhou, Y. Ren, J. Xu, Y. Shi, G. Jiang, Z. Zhao, Fine grained Nd3 þ :Lu2O3
pared by single step and two step sintering, Ceram. Int. 41 (1) (2015) transparent ceramic with enhanced photoluminescence, J. Eur. Ceram. Soc. 34
163–170. (8) (2014) 2035–2039.
[83] M.H. Fathi, M. Kharaziha, Two step sintering of dense, nanostructural for- [112] D. Zhou, Y. Shi, J.J. Xie, P. Yun, Fabrication and luminescent properties of
sterite, Mater. Lett. 63 (17) (2009) 1455–1458. Nd3 þ -doped Lu2O3 transparent ceramics by pressureless sintering, J. Am.
[84] S.M. Mirhadi, Synthesis and characterization of nanostructured forsterite Ceram. Soc. 92 (10) (2009) 2182–2187.
scaffolds using two step sintering method, J. Alloy. Compd. 610 (2014) [113] Z. Khakpour, A.A. Yuzbashi, A. Maghsodipour, K. Ahmadi, Electrical con-
399–401. ductivity of Sm-doped CeO2 electrolyte produced by two step sintering, Solid
[85] S.M. Mirhadi, A.A. Nourbakhsh, N. Lotfian, B. Hosseini, Strength develop- State Ion. 227 (2012) 80–85.
ment, bioactivity and biodegradability of forsterite nanostructure scaffold, [114] T. Isobe, A. Ooyama, M. Shimizu, A. Nakajima, Pore size control of Al2O3
Ceram. Int. 41 (1) (2015) 1361–1365. ceramics using two step sintering, Ceram. Int. 38 (1) (2012) 787–793.
[86] M. Mazaheri, M. Haghighatzadeh, A.M. Zahedi, S.K. Sadrnezhaad, Effect of a [115] S. Li, C. Song, X. Qin, Z. Zeng, Densification and grain growth behavior of
novel sintering process on mechanical properties of hydroxyapatite cera- highly dense MgO ceramics in pressureless sintering, Ceram. Int. 41 (8)
mics, J. Alloy. Compd. 471 (1) (2009) 180–184. (2015) 10148–10151.
[87] M.J. Lukić, S.D. Škapin, S. Marković, D. Uskoković, Processing route to fully [116] Y.J. Eoh, H.J. Ahn, E.S. Kim, Effect of two step sintering on the microwave
dense nanostructured HAp bioceramics: from powder synthesis to sintering, dielectric properties of Zn1.8SiO3.8 ceramics, Ceram. Int. 41 (1) (2015)
J. Am. Ceram. Soc. 95 (11) (2012) 3394–3402. S544–S550.
[88] S. Marković, M.J. Lukić, S.D. Škapin, B. Stojanović, D. Uskoković, Designing, [117] S.N. Hosseini, H.S. Jazi, M. Fathi, Novel electrophoretic deposited nanos-
fabrication and characterization of nanostructured functionally graded HAp/ tructured forsterite coating on 316L stainless steel implants for biocompat-
BCP ceramics, Ceram. Int. 41 (2) (2015) 2654–2667. ibility improvement, Mater. Lett. 143 (2015) 16–19.

View publication stats

S-ar putea să vă placă și