Вы находитесь на странице: 1из 19

Journal

J. Am. Ceram. Soc., 94 [1] 119 (2011)


DOI: 10.1111/j.1551-2916.2010.04210.x
r 2010 The American Ceramic Society

Electric Current Activation of Sintering: A Review of the Pulsed Electric


Current Sintering Process
Zuhair A. Munirw and Dat V. Quach
Department of Chemical Engineering and Materials Science, University of California, Davis, California 95616

Manshi Ohyanagi
Department of Materials Chemistry, Ryukoku University, Ohtsu 520-2134, Japan

The phenomenal increase during the past decade in research


utilizing pulsed electric current to activate sintering is attributed
generally to the intrinsic advantages of the method relative to
conventional sintering methods and to the observations of the
enhanced properties of materials consolidated by this method.
This review focuses on the fundamental aspects of the process,
discussing the reported observations and simulation studies in
terms of the basic aspects of the process and identifying the
intrinsic benets of the use of the parameters of current (and
pulsing), pressure, and heating rate.

I. Introduction

INTERING as a process to consolidate powders is an ancient


art that has been practiced for more than 6000 years in the
making of bricks and pottery1 and in the consolidation of precious metals in pre-Columbian South America.2 The ability to
achieve consolidation without melting is made possible by the
thermal activation of mass transport processes driven by reduction of surface and grain boundary energies. To optimize thermal activation and attain high density with concomitant
strength, sintering is carried out at high temperatures, relative
to the melting point of the material. For practical as well as
economic reasons, signicant efforts have been, and continue to
be directed, toward other means of activation to achieve high
density at lower temperatures or in shorter times. Among these
is the use of an electric current to activate sintering. The recent
widespread use of this form of activation has been referred to
variously as spark plasma sintering (SPS), pulsed electric current
sintering (PECS), eld-activated sintering technique, and current-activated pressure-assisted densication. Research using
eld activation in sintering has increased dramatically in the
past decade and has drawn attention to this process at both the
fundamental and the applied levels.

D. J. Greencontributing editor

Manuscript No. 28308. Received September 20, 2010; approved September 21, 2010
w
Author to whom correspondence should be addressed. e-mail: zamunir@ucdavis.edu

In a previous review, we provided a historical perspective


for the use of a current to activate sintering.3 A recent review
of patents on activated sintering attributes the rst use of
current in sintering to Bloxam, who obtained a patent in
1906.4 However, little and typically unnoticed work on
current activated sintering was carried out during the next eight
decades. Between 1900 and the rst half of 2008, more than 640
patents were issued worldwide4; the majority of these (86%)
were issued since 1990. The topics of these patents cover a wide
range of properties and utility of materials, as can be seen in
Fig. 1.4 The most dominant coverage in these patents deals with
the functional aspect of materials, including magnetic, thermoelectric, and electronic properties, Fig. 1(a), while coverage for
structural properties and use is dominated by utilization as cutting tools and composites, Fig. 1(b). The marked increase in the
number of patents issued since 1990 is mirrored by a corresponding increase in the number of publications on eld-activated sintering. Figure 2(a) shows the number of published
papers since 1993; statistics for earlier years are reported in
the previous review.3 The increase in the number of published
papers worldwide has an exponential trend, reecting the strong
interest in and the utilization of this method of sintering. Nearly
450 papers were published in 2009. Initially, most of the publications came from Japanese investigators, a reection of the
fact that until relatively recently the equipment for eld-activated sintering was manufactured exclusively in Japan. However, as Fig. 2(b) shows, other countries have become more
active in this area, with the largest number of recent publications
now coming from China.
Several reviews of this process, with different emphasis on
aspects of the process or on specic materials, have been published previously.39 In this review, we will attempt to emphasize
the fundamental aspects related to the PECS. In doing so, it will
neither be possible nor advisable to provide a comprehensive
discussion of all published papers on PECS. Aside from being a
daunting undertaking, inclusion of all papers on PECS would
add little to the emphasis on fundamentals, which is the aim of
this review.
The rapid increase in the use of PECS can be attributed
largely to two broad considerations: (a) the intrinsic advantages of the method relative to conventional sintering methods
and (b) the observations of enhanced properties of materials

Feature

Journal of the American Ceramic SocietyMunir et al.

II. Reported Advantages of PECS

(a)
Magnetic

71

Thermoelectric

60

Electronics

30

Biomaterials

26

Electric

26

CNT

12

Sputtering

11

FGM

10

Optical

(b)
Cutting

35

Composite

26

Structural

24
16

Mechanical
Metal

15

Abrasive

13

High temp.

11

Porous

7
0

10

15

20

25

30

35

40

Fig. 1. Number of pulsed electric current sintering patents published


from 1900 to 2008 applied to (a) functional and (b) structural materials.4

2007

(a) 500
450

2008

consolidated by this method. Here, we refer to selected examples


of both categories and plan to discuss some in more detail in the
subsequent sections of this paper.

2001

1999

1996

1995

1994

50

1993

100

1998

150

2000

200

2002

250

2006

2005
2003

300

2004

350

1997

Number of Papers

400

1000

Japan

(b) 1200

China

800

Singapore

Germany

Sweden

Italy

200

France

400

USA

600
S. Korea

Number of Papers

Vol. 94, No. 1

0
Fig. 2. Number of published papers (a) from 1994 to 2009 and (b) by
country.

The literature contains numerous publications demonstrating


the advantages of PECS over other methods of consolidation. A
common advantage is the shorter time needed to consolidate
powders relative to conventional methods, including hot pressing. For example, to obtain a density of B95% for ultrane Ni
powders (100 nm), it took 150 min at 7001C when hot pressing
was used, while it took only 1 min at an even lower temperature
(5001C), using approximately the same pressure.10 Similar
observations of a shorter sintering time and lower sintering
temperatures have been made in other studies.11,12 Attainment
of higher densities at the same temperature has also been
reported.13,14 Thus, the difference between PECS and other
methods has ramications in process efciency and energy savings as well as microstructural and compositional implications.
The energy efciency of PECS relative to hot pressing has been
demonstrated in the sintering of composites.15 With respect to
compositional and microstructural changes, sintering at lower
temperatures and for shorter times minimizes material loss due
to vaporization,1619 undesirable phase transformation,20 and
suppression of grain growth.21,22
While the benets cited above provided a signicant impetus
for the increased interest in PECS, the claims of better or improved properties of materials sintered by this method have
generated an even stronger push for its use. Improvements in a
variety of properties have been reported,2329 including cleaner
grain boundaries in sintered ceramic materials,30 a remarkable
increase in the superplasticity of ceramics,31,32 higher permittivity in ferroelectrics,33 improved magnetic properties,34 reduced impurity segregation at grain boundaries,32 higher
chemical stability,25 higher hydrogen storage capacity in BCC
solid metallic solutions,35 better thermoelectric properties,36,37
improved mechanical properties,38 and better optical properties.39 In the most recent work on impurity segregation, Mitzuguchi et al.12 showed that ZrB2 sintered by PECS had lower
levels of impurities in the grain boundaries and in the grains
than samples sintered by hot pressing. Furthermore, PECS samples had a higher density and a lower grain size than those consolidated by a hot press.
In addition to the above, many reports have indicated unusual accomplishments in materials processing using the PECS
method. The consolidation of mechanically alloyed amorphous
Al-based AlNiTi intermetallic was investigated recently using
three methods: PECS, hot pressing, and pressureless sintering.40
The work demonstrated that the consolidation was best accomplished with PECS to produce a uniform distribution of intermetallic nanoparticles in an amorphous matrix. As will be
discussed in a later section in this paper, a related observation
was made earlier in a study on the effect of an electric eld on
the crystallization of bulk metallic glass.41 Also, recently, Ericksson et al.16 succeeded in consolidating lead-free ferroelectric
niobate ceramics using the PECS method avoiding volatilization, and demonstrated that the ceramic had improved ferroelectric properties with an unusually high remnant polarization.
In a recent investigation, Wang et al.42 produced a glass phase of
zeolite from crystalline powders using eld-activated sintering.
The approach is based on an orderdisorder transformation
under a pulsed-current heating and a uniaxial pressure. In a series of related studies, disorderorder transformation was shown
to enhance the densication of SiC under similar conditions.43
Similar observations were also made in the consolidation of
carbon with an amorphous-graphite transformation.44 Under
PECS conditions, the transformation resulted in an unconventional alignment of lattice planes in the resulting graphite (normally the c-axis of sintered graphite aligns parallel with the
pressure direction, but that resulting from the transformation
aligns with the c-axis perpendicular to the pressure direction).
The combination of prior mechanical activation (high energy
milling) on powders with a subsequent sintering or reactive
sintering by the PECS process has been utilized to simultaneously synthesize and densify nanostructured, intermetallic,

January 2011

Electric Current Activation of Sintering

and composite materials.4547 For example, using this approach,


it was shown that MoSi2 could be microalloyed successfully with
Mg,48 a long sought after goal, facilitating a reduction in the
ductile-brittle transition temperature in accordance with theoretical predictions.49 Accomplishments were also made in the
formation of functionally graded materials (FGMs)5052 and in
the joining (bonding) of materials.5355
At the end of the Section I, we pointed that increased
interest in the PECS process was generally motivated by the
intrinsic advantages of the method relative to conventional
sintering methods and by the observations of enhanced properties of materials consolidated by this method. It should be
pointed out, however, that these considerations are not independent of each other and that the observed property enhancement is most likely (in a large segment of these observations)
related to the reduction in the sintering temperature. Lower
sintering temperatures affect composition and microstructure
and these changes in turn result in different properties. However, it should be emphasized that not all the observations of
property enhancement could be attributed solely to the effect of
lower temperature. As we will discuss in more detail in a subsequent section, it is likely that other parameters of the process
can play a role, specically the current and pressure.

III. The PECS Process


In its simplest form, the PECS process involves the consolidation of powders under the simultaneous action of a current and
a unixial pressure. The current provides the heat to achieve the
desired sintering temperature and its application constitutes the
main difference between hot pressing and the PECS process. As
will be seen later, the application of the current gives rise to the
high heating rates that can be accomplished in the PECS and to
other nonthermal contributions including current effect on mass
transport. Figure 3 shows a schematic of the PECS apparatus.
Typically, a pulsed DC current is applied with a relatively low
voltage (B10 V). The pulsing pattern is made up of a sequence
of pulses with the current, followed by the absence of a current.
Thus, a pulse pattern of 122 means that 12 pulses are applied,
followed by a duration of two pulses where the current is
not applied. Pulses in a typical PECS apparatus are 3.3 ms in
duration.
The simultaneous output of temperature and displacement
(shrinkage) makes it possible to gain an insight into sintering

Fig. 3. Schematic of a pulsed electric current sintering apparatus.

kinetics or the reaction mechanism in the PECS method. However, the observed displacement represents an overall characterization of shrinkage as it also includes contributions from the die
and system. With calibration (to obtain baseline displacement)
and accurate measurements of temperature, it is possible to
obtain valid shrinkage data, as is demonstrated in Fig. 4(a) for
the densication of zirconia.56 Using such an analysis, the PECS
method can provide information on reactivity, as has been
demonstrated in Anselmi-Tamburini et al.56 This is shown in
Fig. 4(b), where the small increase in temperature represents the
exothermic reaction of the formation of B4C from elemental
powders. As the measuring thermocouple also provides feedback to the system, the dip in the power curve represents the
control step taken by the system to account for the small, albeit
transient, increase in temperature due to the reaction. Similarly,
the decrease in the displacement (i.e., shrinkage) signies the
decrease in the molar volume accompanying the reaction.
It is the application of the pulsed current that has been
claimed to be the main advantage of the PECS process. A
more specic discussion on this will be presented below. But the
concept of using an electric spark in the sintering of powders is
not new. As was detailed in a previous review paper,3 in a recent
comprehensive review,8 and as has been discussed in a paper on
PECS patents,4 the concept of electric spark was utilized in various forms for more than a century. However, the recent surge in
the use of this approach is, in large part, the consequence of the
availability of commercial units, manufactured initially by companies in Japan. More recently, PECS equipment has been manufactured in Germany, the U.S., Korea, and China.

(1) Nature and Inuence of Pulsing


A persisting source of controversy regarding the benets of the
PECS process is the oft-repeated claim that the pulsing of the
current creates a plasma that activates the surfaces of the powder particles, through the removal of surface layers (e.g., oxides).
Conicting results have been provided to argue for the existence
or absence of the plasma,5761 but the more convincing evidence
points to its absence under PECS conditions.
However, aside from the concept of plasma, the role of pulsing pattern in sintering and reactivity in the PECS has been the
subject of several investigations. Nanko and coinvestigators
reported earlier an absence of pulse current effects on the sintering of cast iron and Ni-20Cr powders.61,62 Xie et al.63 investigated the effect of the frequency of sintering of Al powders.
They densied the powder in the PECS under different pulse
frequencies: 0 and 300 Hz, and 10 and 40 kHz, with patterns
shown in Fig. 5, and concluded that pulse frequency had no
effect on the densication and microstructure of the sintered
powders. Similarly, Dang et al.64 studied the effect of the waveform of the pulsed current on the sintering of a-Al2O3 using 300
Hz and 16 kHz, with pulse patterns (on-off) of 122 and 26 for
the former frequency and 4010 and 1020 for the latter. Their
results are summarized in Figs. 6(a) and (b), which show the
effect on density and grain size, respectively. As the gures
show, neither the frequency nor the pulse pattern had an effect
on the densication or the grain growth of alumina.
As part of a series of investigations on the fundamental aspects of the SPS process, we studied the nature and effect of the
pulsing pattern.65,66 Figure 7(a) shows a pulse pattern of 82,
i.e., eight pulses of 3.3 ms on, followed by two pulses off. As can
be seen from this gure, the peaks do not correspond to one
voltage and in fact the exact pulse number is not always followed, as is evident in the second sequence of on pulses, where
there are nine instead of eight pulses. A Fourier transform of the
pattern of Fig. 7(a) is shown in Fig. 7(b). The transform exhibits
a peak at about 350 Hz and smaller ones at higher frequencies.
However, it is seen that the bulk of the power in the PECS is
provided by the component at zero frequency, i.e. DC power. As
the contribution of a given frequency to the heating is proportional to the square of its amplitude, we plot this as a function of

Journal of the American Ceramic SocietyMunir et al.

Vol. 94, No. 1

Fig. 4. (a) Corrected displacement during the densication of nanometric zirconia under a pressure of 106 MPa and (b) temperature, power, and
displacement proles during the synthesis of B4C from the elements in PECS (pressure 5 50 MPa and heating rate 5 1001C/min).56

frequency in Fig. 7(c). From this, it can be seen that most of the
heating in the PECS is generated by frequencies of o100 Hz.
As we have seen above, various studies have shown that pulse
pattern had no effect on densication or grain growth.6164 To
investigate the effect of pulse pattern on reactivity (and hence

mass transport) in the PECS, a study was carried out using a


three-layer sample to determine the effect on product formation
at the interfaces between the layers. A wafer of p-type Si was
placed between two foils of Mo and annealed at constant temperatures under different pulsing patterns.65 The use of a three-

Fig. 5. Measured waveforms during the pulsed electric current sintering process of aluminum powder with different pulse frequencies.63

January 2011

Electric Current Activation of Sintering

Fig. 6. Relative density (a) and grain size (b) of alumina pulsed electric
current sinteringed at different temperatures with various frequencies
and pulse patterns.64

IV. PECS Parameters and their Effect on Processing


The parameters that are typically associated with the PECS
process include the current, the applied uniaxial pressure, and
heating rate. Typically, the current and sintering temperature
are dependent parameters as Joule heating is the source of thermal activation, whether in the graphite die only (when the sample is nonelectrically conducting) or in the die and sample (when
the sample is electrically conducting). The maximum (DC) current available depends on the specic PECS apparatus; in the
case of the Sumitomo Model 2050 apparatus, for example, the
maximum current is 5000 A. As seen later, it is possible to make
the current and temperature independent parameters in the SPS
with certain design modications. While the current is commonly (but not accurately) associated with Joule heating only, it
can also play another intrinsic role in enhancing mass transport,
as will be shown subsequently.
The pressure has been shown recently to play a crucial role
in the consolidation of materials, particularly nanostructured

Fig. 7. (a) Typical pulse pattern of 8:2 (on:off) in the PECS; (b) Fourier
transform of the pattern of (a); (c) Fourier transform plotted as the
square of the magnitude versus frequency.66

powders. Until recently, the maximum pressure that can be uniaxially applied in the PECS was limited by the mechanical property of the graphite die. A theoretical upper limit of about
5000
1270 C, pulses 12:2
1270 C, pulses 7:7
1170 C, pulses 12:2
1170 C, pulses 8:2
1070 C, pulses 8:2
1070 C, pulses 2:8

4500
Square of Layer Thickness ( m2)

layer system was planned to determine the possible effect of the


direction of the DC current. As can be seen from Fig. 8, the
pulse pattern had no effect on the thickness of the product (primarily MoSi2) formed at the two interfaces. Moreover, the
direction of the current also had no effect. As will be seen
later in this paper, this latter nding is not surprising in interactions where a compound is formed.
The results discussed above lead to the conclusion that pulsing pattern and pulse frequency have no measurable effect on
densication, grain growth, and mass transport.

4000
3500
3000
2500
2000
1500
1000
500
0
0

1000

2000
3000
Time (s)

4000

5000

Fig. 8. The growth of MoSi2 layer in the pulsed electric current sintering at different temperatures under different pulse patterns.65

Journal of the American Ceramic SocietyMunir et al.

Graphite
Tungsten carbide
Silicon carbide
Sample

Fig. 9. Schematic of a double acting die that allows the use of much
higher pressures.67

140 MPa is used as a guide, but in practice, the limit may be lower.
Recent modications of die design (Fig. 9) have made it possible
to achieve much higher pressures.67 With this design, it is possible
to apply pressures as high as 1 GPa. The importance of this will be
seen when we discuss the role of the pressure, below.
Heating rate is another parameter that distinguishes the
PECS process from hot pressing. Because of the nature of heating, much higher heating rates can be achieved in the PECS, as
high as about 20001C/min. The advantage of higher heating
rates is the bypassing of the nondensifying sintering mechanisms, e.g., surface diffusion. However, for nonelectrically conducting samples, the heating rate can play a role in the
occurrence of thermal gradients, depending on the thermal conductivity and the size of the sample.68

(1) Inuence of the Current


(A) General Observations of Current Effect in Materials
Processing: As was stated above, the notion of the presence of
plasma (and hence the common name of the process) has not
been demonstrated adequately and thus the (nonthermal) role of
the current is likely to be in its effect on mass transport. That the
current could have an inuence on mass transport has been
shown clearly by numerous investigations. The current enhances
mass transport through electromigration,69 point defect generation,70 and enhanced defect mobility.71
Electromigration studies on multilayer systems have shown
that the current increases the rate of product layer formation
and decreases the incubation time for the nucleation of the new
phase.7275 The imposition of a current has also been shown to
have other effects. It was shown to increase the solubility in liquid metals and inuence the resulting microstructure of the
solidied product.76,77 In a study on the crystallization of
bulk metallic glasses, the imposition of a current was shown to
inuence the grain size and fraction of the resulting nanocrystallites.41 Conrad and colleagues have carried out extensive
investigations on the effect of an electric eld (current) on a
variety of materials-related processes. In a recent investigation,

Vol. 94, No. 1

they showed that the imposition of a modest electric eld reduced the tensile ow stress of MgO, Al2O3, and tetragonal
ZrO2.78 They attributed the decrease to a reduction in the electrochemical potential for the formation of vacancies. They also
reported a retardation of grain growth as a consequence of the
eld. It was proposed that grain growth retardation might be
attributed to either a eld effect on solute ion segregation (Y in
the case of ZrO2), a decrease in grain boundary energy, or a
decrease in ion mobility. The grain growth retardation was consistent with earlier observations on the inuence of a eld on
copper and tetragonal zirconia, as can be seen in Figs. 10 and 11,
respectively.79,80 Similar observations of grain growth retardation in tetragonal yttria-stabilized zirconia (YSZ) were recently
reported by Ghosh et al.81 with the application of an electric
eld of about 4 V/cm. While these observations of eld effects
on grain growth have important implications in the processing
of materials, their underlying cause is not yet clear, a circumstance that does not diminish their importance but clearly suggests that more research needs to be carried out.
(B) Current Effects in PECS Processing: Various observations made while processing materials on the PECS have
been directly or indirectly attributed to the role of the pulsed
current. Using a starting powder of Pb(Mg1/3Nb2/3)O3-PbTiO3
with a grain size of 110 mm, Chen et al.82 obtained a sintered
body with a grain size in the range 20100 nm. This unique
observation of making a nanostructured ceramic from microstructured powders during sintering in the PECS was attributed
to the role of the current. It is proposed that the pulsed current
induced thermo-mechanical fatigue, which resulted in the
breakup of the microstructured grains into nanograins. This
observation is of signicant interest and its application to other
materials is needed before it can be considered as a general phenomenon in PECS processing. Nagae et al.83 studied the sintering of aluminum powders by the PECS and hot-pressing
methods and reported a lower electrical resistivity of samples
sintered by the PECS, which they attributed to the effect of the
pulsed current on the destruction of the surface oxide through
localized Joule heating at the contact points between particles.
Another observation, which was attributed to the role of the
eld (current), was made during the sintering of the spinel
MgAl2O4. Mussi et al.84 observed a different (inversion) occupation of the tetrahedral and octahedral sites when powders
were sintered in the PECS relative to observations when the
spinel was sintered by pressureless sintering and hot-isostatic
pressing. They explained their nding on the basis of the effect
of the PECS process (presumably the electric eld) on the
space charge. While there is no direct evidence of changes in
the space charge layer during SPS sintering, the occurrence of
space charge and its associated distribution of defects in magnesium aluminum spinel and alumina have been demonstrated.85,86 As in the case of the observation of grain growth
retardation, the concept of a change in the space charge in ceramics due to exposure to PECS conditions remains to be demonstrated directly.
The effect of current on reactivity in the PECS has been
investigated.56,8789 The reaction between Mo foils and Si wafers
was investigated under PECS conditions with and without a
current passing through the multilayer ensemble.56 As was
stated above,65 the reaction between these elements was not
affected by the pulse pattern. However, the reaction rate to produce the interface product, primarily MoSi2 with minor
amounts of Mo5Si3, was markedly inuenced by the presence
of a current, as can be seen in Fig. 12. The kinetics of growth of
MoSi2 were determined from the results under both conditions,
as can be seen in Fig. 13, in which the rate constant is plotted
against the reciprocal of the absolute temperature. The calculated activation energies for the growth of MoSi2 for the cases
with and without a current (B600 A/cm2) are 168 and 175 kJ/
mol, respectively. The effect of the direction of the DC current
on the growth of the product was made possible by the threelayer geometry (Si is sandwiched between two Mo foils). The
results, depicted in Fig. 14, show that there is no effect of current

January 2011

Electric Current Activation of Sintering

Fig. 11. Mean linear intercept grain size d of tetragonal zirconia versus
temperature in the grip tab (e  0) and near the fracture surface (e  1.0)
with and without an electric eld.80

the electromigration study on the CuNi system that forms continuous solid solutions.75 In that study, a clear dependence on
the direction of the current on interdiffusivity was seen.
Other observations on the effect of current on reactivity under PECS conditions include reactions between carbon and
Nb,87 Mo,88 and Ti and Zr.89 Figure 15 shows the time dependence of the growth of the product layer thickness (b-Mo2C) on
current density for samples annealed at 1842 K.88 The effect of
the current density is best shown in Fig. 16, in which the
annealing time is constant at 20 min. The gure shows an apparent threshold value of current before an effect is observed; in
this case, the value is approximately 500 A/cm2. In contrast to
the case of the reaction between Mo and Si,56 the activation energy for the growth of the product has a strong dependence on
current density, as can be seen in Fig. 17. Relative to annealing
without a current, a decrease of about 44% was seen when the
samples were annealed at 1476 A/cm2. In the cases of the growth
of TiC and ZrC layers, the activation energy was basically unaltered with the application of a current, although the growth
rate was enhanced in the presence of a current.89 Similar results
were obtained with the system NbC, except that in this case,
two carbide phases were formed: NbC and Nb2C.87 Again,
growth enhancement was observed, but no change in activation
energy. The difference in behavior between the MoC case and
the others (MoSi, TiC, ZrC, and NbC) is not well understood. It may be the consequence of the kinetics of the ab phase
transformation of Mo2C relative to similar transformations in
the other binary systems.

Fig. 10. Log grain size D versus log time t for annealing of a Cu foil
(thickness 5 18 mm) at 15011951C with and without an electric eld:
(a) top side of the foil and (b) bottom side.79

direction on the growth of the product layer. This nding is not


surprising as the growth of the product, MoSi2 in this case,
requires the diffusion of both Mo and Si. Thus, if one is
enhanced by electromigration, the growth rate will still depend
on the slower diffusing species. A validation of this is seen in
cases where no product forms. This was demonstrated clearly by

Square of Layer Thickness (m2)

2500
1270 C current
1270 C no current
1200 C current
1200 C no current
1150 C current
1150 C no current
1100 C current

2000

1500

1000

500

0
0

1000

2000
3000
Time (s)

4000

5000

Fig. 12. Growth rates of the MoSi2 layer at different temperatures in


the presence and absence of current owing through the sample.65

8
0.0
0.2

with current

0.4

without current

Square of Layer Thickness (108m2)

5.0

0.6
ln (k)

0.8
1.0
1.2
1.4
1.6
1.8
2.0

6.4

6.6

6.8

7.2

7.4

The results presented above clearly show the effect of the


current on reactivity (mass transport) under PECS conditions.
However, as the PECS method is used primarily for the consolidation of powders, a more convincing investigation would be
the direct demonstration of current effect on sintering. Such an
investigation was carried out using the sintering of copper
spheres to copper plates as the model.90 Figure 18 shows a schematic of the arrangement used in the SPS. In order to demonstrate the effect of the current, it was necessary to devise an
experimental setup in which the current can be varied while the
temperature is held constant.
This is important as under a normal PECS operation, the
current and temperature are interdependent parameters. The use
of carbon foil layers made it possible to achieve the desired goal.
The current (I) dependence on the number of graphite foil layers, x, is derived as90:
1
x1=2

P
4Rco 2Rgf

4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0

Fig. 13. Arrhenius plot of the temperature dependence of the growth


rate of MoSi2 in the presence and absence of current owing through the
sample.65

0 [A.cm2]
521 [A.cm2]
732 [A.cm2]
1145 [A.cm2]
1476 [A.cm2]

4.5

0.0

104/ T (K1)

1=2
(1)

where P is the power and R is the resistance, where the subscript


co refers to contact resistance and gf refers to graphite foil
resistance. Figure 19 shows the experimentally determined current as a function of the number of graphite foil layers for
T 5 9001C and 15 min. The gure also shows the predicted
relationship from Eq. (1) as a solid line. The effect of current on

60
50
40
30
20

500

1000
1500
Holding Time (s)

2000

Fig. 15. Inuence of current density on the growth of the b-Mo2C layer
at 1842 K.88

the sintering of the copper spheres to copper plates is shown


qualitatively in Fig. 20. The gure shows SEM images of the
fracture surface of the necks that formed between them on sintering at 9001C for 60 min under different current values, ranging
from 0 to1040 A. These images clearly show the effect of the
current by the increase in the diameter of the neck with an increase in current. Quantitative interpretations of these results
were made from measurements of the radii of the neck, x, and
expressing them in the form,
 x n
R

Bt
Rm

(2)

where R is the radius of the sphere, B is a constant that contains


the diffusion coefcient, and the exponents n and m are mechanism-dependent constants. Eq. (2) is for the initial stages
of sintering with x/Rr0.3. A plot of (x/R) versus t is shown in
Fig. 21, in which the effect of the current is clearly demonstrated.
An important observation made in that study is the presence
of features on the surface of the copper plates indicative of
evaporation. The features appear as white rings surrounding
neck regions under optical microscopy (Fig. 22) but are shown
to be surface ledge structures when observed by scanning electron microscopy, Fig. 23. The extent of these areas is directly
200

Product Layer Thickness (m)

70
Bottom Layer Thickness (m)

Vol. 94, No. 1

Journal of the American Ceramic SocietyMunir et al.

150

100

10
50

0
0

10

20
30
40
50
Top Layer Thickness (m)

60

70

Fig. 14. Comparison of MoSi2 layer thickness at the two (MoSi and
SiMo) interfaces relative to the direction of the current.65

500

1000

1500

Current Density (A.cm2)


Fig. 16. Inuence of current density on the thickness of the b-Mo2C
layer (T 5 1842 K; t 5 20 min).88

January 2011

Electric Current Activation of Sintering

Fig. 17. Calculated values of the activation energy for the growth of
b-Mo2C as a function of current density.88

related to the strength of the imposed current, as can be seen


from Fig. 22. The occurrence of these ledge structures and their
proximity to the neck regions suggest an effect of the current on
evaporation. While there are no published accounts on the effect
of a current on evaporation, there is indirect evidence from previous electromigration studies on the system AgZn.91
A comparative study showing the contribution of the PECS
conditions to sintering relative to hot pressing was carried out by
Fu et al.92 In this study, the sintering between spheres (CuNi
and FeCu) was investigated under the two conditions. In contrast to the study on copper spheres and copper plates,90 in this
study, the difference was not related directly to the level of the
current, but only to a specic temperature and hold time of annealing. However, the results unambiguously show an effect of
the PECS conditions on sintering, as can be seen from Table I.92
The table shows the effect of temperature and hold time on the
(x/R) ratio for samples sintered in PECS and in the hot press.
Also shown in Table I are calculated diffusion coefcients based
on neck formation.93 The results show a signicant sintering enhancement under PECS conditions relative to hot pressing, both
in terms of neck formation and in terms of the diffusion coefcient. The diffusion coefcients of Ni, calculated from concentration proles across necks between nickel and copper spheres
for samples sintered under the two conditions, are shown in
Fig. 24.92 The values obtained when sintering in PECS
are greater by a factor of 23 than those obtained during hot
pressing.
More recently, N. Toyofuku, T. Kuramoto, T. Imai, M.
Ohyanagi and Z. A. Munir (unpublished results) investigated
the effect of current on the sintering of W wires to W plates, in
an arrangement similar to that described above for the copper
spheres and plates. The time dependence of the neck radius is
shown in Fig. 25 for the cases of sintering with and without

Fig. 19. Dependence of total current on the number of graphite foil


layers at 9001C.90

current at 17001C. After 30 min of sintering, the neck radius with


a current is 1.5 times that of the neck formed in the absence of a
current. Evidence of a possible effect of the current on evaporation is also provided. However, in this case, the effect is likely
due to the reduction of surface oxide, as has been observed in the
sintering of tungsten.94,95
The investigations discussed briey above (N. Toyofuku,
T. Kuramoto, T. Imai, M. Ohyanagi and Z. A. Munir unpublished results)90,92 provide evidence of sintering enhancement
under PECS conditions, with a direct correlation to the level of
the current, as was shown in the sintering of copper spheres to
plates. The enhancement of mass transport is believed to be the
consequence of electromigration. The increase in the ux, Ji, of a
diffusing species, i, is a result of the momentum transfer from
the electron wind effect, as can be seen from the following
relationship96:
Ji



Di Ci RTq ln Ci
Fz E
RT
@x

(3)

where Ji is the ux of the diffusing ith species, Di is the diffusivity


of the species, Ci is the concentration of the species, F is Faradays constant, z is the effective charge on the diffusing species,
E is the eld, R is the gas constant, and T is the temperature.
The above studies show the effect of current on neck formation, i.e. during the initial stage of sintering. We are not aware of
any study showing such an effect directly during the intermediate
and nal stages of sintering.

(2) The Effect of Pressure


Experimental observations showing the benet of pressure in
densication are numerous, with hot pressing being a common
example.97 The role of pressure in sintering has been investigated extensively. The effect of pressure on various applicable
mechanisms in sintering has been discussed in a review by German.98 In the present review, we will focus primarily on the
effects of pressure in the PECS process. In typical PECS experiments, using graphite dies, there is a practical upper limit for the
applied pressure (B140 MPa) dictated by the mechanical properties of graphite. Higher pressures were possible through a
modication of die design.66 As was indicated in a previous
review,3 the pressure has intrinsic and extrinsic effects on sintering; at a fundamental level, the former involves an increase in
the chemical potential, as indicated by the following99:
mI moi  sn OI

Fig. 18. Schematic of a sample of a Cu sphere to plate sintering geometry in the pulsed electric current sintering apparatus.90

(4)

where mI is the chemical potential at a particle interface under


stress, moi is the standard chemical potential, sn is the normal

10

Journal of the American Ceramic SocietyMunir et al.

Vol. 94, No. 1

Fig. 20. SEM images showing the effect of current on the neck formation between copper spheres and copper plates sintered at 9001C for 60 min:
(a) zero current, (b) 700 A, (c) 850 A, and (d) 1040 A.90

stress at the interface, and OI is the atomic volume of the diffusing species. In addition to inuencing diffusion-related mass
transport, the pressure inuences other processes intrinsically,
including viscous ow, plastic ow, and creep. Extrinsically, the
pressure inuences particle rearrangement and the destruction
of agglomerates in powders, the latter playing an important role
in the consolidation of nanopowders, as will be seen below.
Makino et al.100 investigated the effect of pressure on the
sintering of ultrane a-alumina powders under PECS conditions. They showed that densication under a low pressure (30
MPa) was inuenced by the grain size of the starting powder
(100 and 230 nm powders) but that such an inuence was suppressed when densication was carried out under a high pressure

Fig. 21. Time dependence of neck growth between copper spheres and
copper plates at 9001C under different currents. The neck size at zero
time refers to the value obtained during ramp up to temperature.90

(100 MPa). In addition, the authors reported an inuence of the


pressure on crystallite growth; grain growth suppression increased when the powders were sintered at the high pressure,
as can be seen in Fig. 26.
Guillard et al.101 investigated the role of the PECS parameters
in the densication of SiC and showed that the effect of pressure
depended on the temperature at which it is applied. They sintered SiC under two different conditions: in the rst, they
applied the pressure (75 MPa) at the ultimate temperature of
sintering (18001C), and in the second case, the pressure was
applied at a lower temperature (10001C). Their results show that
samples in which the pressure was applied at the lower temperature (i.e., case 2) had lower densities, which they attributed to
the difculty in removing closed porosity after the application of
the pressure. In a similar investigation, Chaim and Shen102
showed no effect on the density of the temperature at which
the pressure was applied in the sintering of Ndyttriumaluminum garnet (YAG) nanopowders. However, the effect on grain
size was more complex: the grain size was lower when the pressure was applied at the sintering temperature if the latter (T) was
less than about 13751C but it was larger when sintering was
carried out at higher temperatures. The grain size was independent of the sintering temperature when the pressure was applied
at a lower temperature (12001C). The authors explain their
results in terms of the role of particle coarsening. They suggest
that the process of coarsening during the heating up period has a
strong inuence on grain growth. This process may lead to a
variation in the particle size distribution, which, in turn, affects
the grain boundary curvature before pressure application. They
conclude that application of pressure before signicant coarsening of the nanoparticles would be benecial for the suppression of grain growth in the dense compact.
To demonstrate the effect of the applied pressure (as well as
other sintering parameters), Chaim and Margulis103 developed
SPS densication maps for nanocrystalline MgO using hot-isostatic pressing as a model. As can be seen in Fig. 27, an increase
in pressure at a constant relative density changes the mechanism

January 2011

11

Electric Current Activation of Sintering

Fig. 23. SEM image near the edge of a neck showing the formation of
ledges (AIP).90

was pointed out above, in order to reach high pressures in the


PECS, we have designed a die conguration, as shown in Fig. 9.
The results of a more detailed determination of the effect of
pressure on density and grain size for YSZ are depicted in
Fig. 29.105 When sintering was carried out at a relatively low
temperature (9801C), the pressure had a marked effect on density; the density increased from about 78% to 96% as the pressure was increased from 150 to 700 MPa. However, when
sintering was carried out at a relatively high temperature
(11801C), the pressure had a marginal effect.
The difference in behavior seen in these results reects the role
of temperature relative to that of the pressure, as can be seen in
the following relationship:
 g

dr
B g P
1  r dt
x

Fig. 22. Optical micrographs of neck images on a copper plate showing


the halo formation around the perimeters of necks: (a) zero current,
(b) 700 A, and (c) 1040 A.90

from diffusion to plastic ow control. Or for any given pressures, the sintering is dominated initially by plastic ow and
nally by diffusion for the case of MgO. Agreement with
experimental observations was taken as support for the validity
of the assumed model.
The effect of pressure on the densication of nanopowders
was demonstrated in the case of sintering of 8 mol% YSZ.104
Figure 28 shows the effect of the pressure on the temperature
needed to obtain 95% dense YSZ in 5 min. The gure shows an
exponential decrease in temperature from about 13501 to about
8901C as the uniaxial pressure is increased from 30 to 800 MPa.
The grain size decreased initially from about 200 to 15 nm and
then remained constant for pressures higher than 150 MPa. As

(5)

where r is the fractional density, B is a parameter that includes


the diffusion coefcient (of the slowest species) and temperature,
g is a geometric constant, g is the surface energy, x is a parameter representing a size scale that is related to particle size, t is the
time, and P is the effective pressure. The effective pressure
exerted on pores varies according to the pore geometry and
the stage of sintering, but we can assume its value to be proportional to the macroscopic applied pressure. At the lower
temperature, mass transport through diffusion is less signicant
(B in Eq. (1) is relatively small) and the pressure plays a dominant role. At the higher temperature, the relative contribution
of the pressure becomes less signicant. This was demonstrated
by Quach et al.105 through simple comparative calculations.
An important, extrinsic, contribution of the pressure relates
to its effect on agglomerates in powders, especially for nanopowders. Nanopowders are susceptible to the formation of
agglomerates due to van der Waal bonds between particles.106
When compacted, agglomerates produce an inhomogeneous
structure in the green body and this leads to a low green density.107 The role of the pressure in the particle rearrangements
and the breakup of agglomerates is illustrated schematically in

Table I. Neck Ratios (x/R) and Calculated Diffusion


Coefcients of Ni in the Pulsed Electric Current Sintering
(PECS) and Hot-Press Sintering of Ni/Cu Spheres92
Sintering process

T (1C)

Hold time (s)

x/R

DNi (  108) (m2/s)

PECS

1000
1100
1000
1100

300
300
2700
2700

0.595
0.705
0.548
0.626

3.564
9.239
1.111
2.142

Hot press

12

Journal of the American Ceramic SocietyMunir et al.


65

PECS
HP
6
Crystallite Size (nm)

60

5
DNi (1010 cm2/s)

Vol. 94, No. 1

4
3
2

AA-alumina (100MPa)
AA-alumina (30 MPa)
TM-alumina (100 MPa)
TM-alumina (30 MPa)

55

50

45

40
900

0
1060

1080

1100
1120
1140
Temperature (K)

1160

1180

Fig. 24. Temperature dependence of the diffusion coefcients of


Ni at the interface for pulsed electric current sintering (PECS) and hot
pressing (HP).92

Fig. 30.108 The effect of pressure on the pore size distribution of


a reactive mixture of ZrO2 and Y2O3 is shown in Fig. 31.108 As
will be discussed below, the destruction of agglomerates through
the application of high pressure was the key to producing, for
the rst time, dense bulk cubic YSZ with a grain size o20 nm.104
Another aspect of the inuence of pressure relates to the
temperature difference between the sample and the die.109
Grasso et al.109 investigated the effect of the applied pressure
on the difference in the temperature at these two locations
experimentally and through simulation for a graphite die and
sample. They showed that an increase in pressure resulted in a
markedly lower difference in temperature at the two locations,
and attributed this to a decrease in electrical and thermal contact
resistances at the punch/die interface due to Poisson deformation of the punch with a higher pressure. Contact resistance has
been identied previously as a reason for the temperature differences between the surface of the graphite die and the center of
the sample.110,111 Another aspect of the pressure, the rate at
which it is applied, was investigated by Xu et al.112 in the densication of YSZ. They found that the displacement rates
affected the densication rates and the nal density of the sam-

60

1000
1100
1200
1300
1400
Consolidating Temperature (C)

1500

Fig. 26. Dependence of crystallite size on pulsed electric current sintering (PECS) temperature and applied pressure for two different commercial powders of alumina (TM-alumina and AA-alumina).100

Fig. 27. Densication map for 20 nm particle size nanocrystalline MgO


at 7501C in pulsed electric current sintering (PECS).103

ples signicantly. Higher rates resulted in higher densication


rates, as can be seen from Fig. 32.

(3) The Effect of Heating Rate


One of the main differences between the PECS process and hot
pressing is the heating rate. Heating rates as high as about

55
With current
Neck radius (m)

50
45
40
Without current
35
30
25
20
5

10

15

20
Time (min)

25

30

35

Fig. 25. Time dependence of neck growth between W wires and plates
annealed at 17001C with and without current (N. Toyofuku, T. Kuramoto,
T. Imai, M. Ohyanagi and Z. A. Munir unpublished results).

Fig. 28. Relationship between hold temperature and the applied pressure required to obtain samples with a relative density of 95% in the case
of nanometric fully stabilized zirconia (8% YSZ). Hold time: 5 min. The
grain size of the materials is also shown.104

January 2011

Electric Current Activation of Sintering

105
1180 C

% Theoretical Density

100
95
90

980 C

85
80
75
70
0

200
400
600
Applied Pressure, MPa

800

Fig. 29. Effect of the applied pressure on the nal density of c-YSZ
samples pulsed electric current sinteringed at 9801 and 11801C for
5 min.105

20001C/min can be achieved in the PECS. High heating rates


reduce the time that powders dwell at the lower temperatures,
where nondensifying (grain coarsening) mechanisms (e.g., surface diffusion) are dominant. In addition, higher heating rates

13

create an additional driving force due to large thermal gradients.113 However, despite theoretical expectations and the results
of simulations,114 experimental observations on the effect of the
heating rate in the SPS have provided conicting results.
For example, in the case of alumina, in one study, it was
found that the heating rate (507001C/min) had no effect on the
nal density,115 and in another study, the effect was negative
(i.e., the density decreased with an increase in the heating rate)
when the heating rates were high (43501C/min).116 In both of
these studies, the heating rate had an effect on grain size; higher
heating rates resulted in smaller grain sizes. Similar observations
were made more recently by other investigators in a study on
alumina117 and on cubic zirconia.118 Figure 33 shows the effect
of the heating rate on the densication and grain size of 8 mol%
YSZ sintered under a high pressure (500 MPa).105 As can be
seen in the gure, the heating rate had no effect on density but
had an effect on grain size, consistent with earlier observations.
The inconsistency in some of the results on the effect of heating rate on densication in the PECS is likely the consequence of
differences in the materials properties and also experimental
uncertainties. The latter include differences in the effective thermal and electrical conductivities of the samples (and thus
temperature gradients) on the contact resistances between the
sample and the die and between parts of the die assembly, and
the timing and rate of pressure application. And the effect of the
heating rate on grain size relates to the bypassing of grain coarsening processes and on the effective time for sintering: higher

Fig. 30. Schematic on the role of applied pressure in particle re-arrangements and breakup of agglomerates in the reactive mixture of ZrO2 and Y2O3
when the pressure is (a) 4 MPa (b) 8 MPa (c) 95 MPa (d) 400 MPa.108

14

Journal of the American Ceramic SocietyMunir et al.

Vol. 94, No. 1

Fig. 31. Pore size distribution curves determined from gas desorption measurements for a reactive mixture of ZrO2 and Y2O3 obtained by compaction
under (a) 4 MPa (b) 8 MPa (c) 95 MPa and (d) 400 MPa.108

V. Simulation Studies on the PECS Process

% Theoretical Density

Simulation studies on the PECS process have been carried out


generally for two purposes: (1) to verify the role of the assumed
parameters of the process and to make predictions on their
effects and (2) to explain observations made in the processing of
various materials by this method. Thus, an important contribution of all these studies is providing a basic understanding of the
PECS process, and in doing so, removing the black box
stigma that, we believe, has contributed to the slow establishment of this method in the United States.
A signicant number of simulation investigations have been
conducted in recent years. Matsugi and colleagues investigated
the voltage, temperature, and density distributions of titanium.
They found that the largest heat source was in the punch of the

die and thus heat ow was mainly from the punch to the sample.119 McWilliams and Zavaliangos120 investigated the density
evolution in relation to the conduction path of the current and
showed that variation in the local density in sample during the
PECS process can inuence its sintering behavior.
Olevsky and Froyen have made signicant contributions in
the area of simulations. They examined the role of the Ludwig
Soret effect of thermal diffusion during SPS sintering and
reported it to be signicant, especially for small particle sizes.121
Using a model that incorporates thermal diffusion, they found
their predicted results on the sintering of Al2O3 to be in qualitative agreement with experimental observations,116 as can be
seen in Fig. 34. Olevsky et al.114 also investigated the effect of
heating rate on densication and, as was pointed out above,
their results show an enhancement of consolidation with an increase in the heating rate. Figure 35 shows the predicted effect of
heating rate on shrinkage of aluminum powders. In another
study, Olevsky and Froyen122 incorporated electromigration
into a constitutive model for PECS sintering and concluded
102

80

100

70

98

60

96

50

94

40

92

30

90
0
Fig. 32. Time evolution of the relative density of eight YSZ samples
processed using varying displacement control rates at 12001C.112

100
200
300
Heating Rate, C/min.

Grain Size, nm

heating rates have a shorter dwell time and are thus expected
to result in smaller grain growth. Nevertheless, this area of
investigation has not received adequate attention experimentally. Simulations studies have been performed with clear predictions, as will be discussed below.114

20
400

Fig. 33. Effect of the heating rate on the density and grain size of eight
YSZ heated up to 11801C with no holding time under 500 MPa.105

January 2011

15

Electric Current Activation of Sintering

Fig. 36. (a) Sample of 1-mm-thick YSZ 8% densied at 10001C under a


pressure of 600 MPa; (b) sample of 1-mm-thick YSZ 3% densied at
10001C under a pressure of 800 MPa. For both samples, the hold time
was 5 min.135

Fig. 34. Porosity kinetics during pulsed electric current sintering of


alumina powder.121 Comparison of the developed model taking into
account the impact of thermal diffusion with the experimental data of
Shen et al.116

that mass transport due to this effect can provide a signicant


contribution to the sintering process. This provides analytical
support for the experimental observations discussed above:
(a) the direct sintering enhancement due to the action of a current for the case of copper spheres to copper plates90 and tungsten wires to tungsten plates (N. Toyofuku, T. Kuramoto,
T. Imai, M. Ohyanagi and Z. A. Munir unpublished results)
and (b) for the observed reactivity enhancements in the PECS by
the action of the current.8789
Numerous other simulation studies have focused on the temperature and current distributions for conducting and nonconducting materials under PECS conditions.65,110,111,123129 In
addition to elucidating the temperature and electrical potential
distributions, Wang et al.130 also simulated the stress distribution in SPS experiments and found that stress gradients, which
depended on materials properties (coefcient of thermal expansion, CTE, and modulus), were larger than thermal gradients.
Stress gradients, both in the vertical and in the radial directions,
were found to be signicant, especially for materials with a high
CTE, which indicates that simple calculations of stress on the
basis of the force and sample area may not be valid. The inuence of percolation on the sintering of ZrO2TiN composites by
SPS was investigated experimentally and by simulation by Vanmeensel et al.131,132 The inuence of adding TiN to zirconia on
the temperature and current distribution was simulated.
The effect of pressure on the homogeneity of SPS-sintered
tungsten carbide was investigated experimentally and by simulation by Grasso et al.133 Taking into account the electrothermal
contact resistance change due to sample shrinkage and the concomitant punch sliding, and using a moving-mesh nite element
model,129 the authors showed the effect of pressure on grain
growth, residual porosity, and hardness distributions along the

sample radius. Increasing sintering pressure resulted in a reduction in the sintering temperature, a conclusion consistent with
previously reported observations on oxide ceramics.104

VI. Consolidation of Functional Materials


In Section II,, we provided examples of the reported advantages of the PECS process. In this section, we highlight selected
accomplishments focusing on functional materials, including
transparent materials, electroceramics, and porous materials.

(1) Transparent Materials


Efforts aimed at obtaining transparent materials focus on the
parameters of density and grain size. Pores and grain boundaries
are light-scattering regions, but as has been shown, porosity
plays a more determining role.134 A decrease in pore size (to
nanometric scale) leads to a decrease in scattering, a circumstance that has led to the motivation to prepare nanostructures
as a means of obtaining transparency.134 Success in obtaining
transparent nanostructured tetragonal and cubic zirconia was
demonstrated, Fig. 36.135 The effect of sintering pressure on
transparency for both modications of zirconia is shown in
Fig. 37; for any given temperature, increasing the pressure
changed the samples from translucent to transparent.
Success has been shown in the preparation of a variety of
transparent materials by PECS processing. A transparent
MgAl2O4 spinel was prepared by reducing porosity and grain
size by controlling the heating rate (o101C/min).39 Morita and
colleagues found that high heating rates enhanced the formation
of closed porosity during the heating process, with the closed
pores remaining at grainboundary junctions. Figure 38 depicts
the SEM images showing the effect of the heating rate on the
microstructure of samples heated at 100 and 101C/min.39 Other
transparent materials have also been prepared successfully by
using the PECS method. These include alumina, AlN ceramics,
mullite, YAG, cubic zirconia, spinel, and others.136144
(a) 900

(b) 900
YSZ 3%

YSZ 8%

400
300

Fig. 35. Simulations on the shrinkage kinetics of aluminum powder.114

600
500
400
300

Translucent

500

700
Pressure (MPa)

600

Translucent

Pressure (MPa)

700

Transparent

800
Transparent

800

200

200

100
850 900 950 1000 1050
Temperature (C)

100
900 950 1000 1050 1100
Temperature (C)

Fig. 37. Combined temperature and pressure conditions required for


optical transparency in (a) YSZ 3% and (b) YSZ 8% sintered powders.
Transparency limit set at 10% of transmittance at a wavelength of 600
nm for a 1-mm-thick sample.135

16

Vol. 94, No. 1

Journal of the American Ceramic SocietyMunir et al.

Fig. 38. SEM images of MgAl2O4 spinel pulsed electric current sintered at 13001C with no hold time under different heating rates of (a) 1001C/min and
(b) 101C/min.39

T (C)
500 400

300

200

100

101
102
103
tT/ (S cm1K)

(2) Porous Materials


Typically, the use of the PECS method is directed toward
obtaining dense materials; however, the method has also been
used to obtain porous materials. Kun et al.145 obtained porous
stainless steel and found it to have compressive strength superior
to that of samples prepared by hot pressing. Through modication of die design, it is possible to obtain a temperature gradient along the vertical axis of the die, and with this, obtain
FGMs. Suk et al.146 used this approach to obtain porous structures of tungsten FGM with porosity and pore size distribution,
as shown in Fig. 39. Using TiH2 powder as an additive to form
pores upon decomposition, Zhao and Taya147 prepared porous
NiTi alloys by PECS. Other nanocrystalline alloys of Ti were
prepared in porous form for biomedical applications,148 and
other porous materials for biomedical (implant) applications
have also been prepared successfully using the PECS methods.149,150 Porous alumina, boron nitride, and other materials
have also been prepared successfully.151,152

104
105
106
107
108
109
1010

Grain size
(nm)
13
22
31
50
100

1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4

(3) Electroceramics
The concept of the size effect in electroceramics has been the
focus of a number of investigations. The aim is to demonstrate
the effect of grain size on the electrical properties of various
oxide ceramics. Okamoto et al.22 investigated the effect of grain
size in suppressing the cubic-rhombic phase transformation in
(12 mol%) scandia-stabilized zirconia, and thus the elimination
of a discontinuity in electrical conductivity. Their work showed
that ne grain size, which can be obtained by the PECS, lowered
the rhombic-cubic transformation point to room temperature
and thus prevented the formation of the less conductive
rhombohedral phase (Zr7Sc2O17).
45

8
Porosity
Pore Size

7
6

35
5
30
4
25

Pore Size (m)

Porosity (%)

40

20

2
0

4
6
8 10 12 14 16
Distance from the Top (mm)

18

20

Fig. 39. Porosity and pore size at various locations in the pure W
sample146

103T 1 / K1
Fig. 40. Temperature dependence of conductivity of YSZ for nanometric samples in the range 13100 nm (pH2 O 5 23 000 ppm).153

In a more recent series of investigations on YSZ and samaria


and gadolinia-doped ceria, new and unexpected behavior was
discovered. For the rst time, through the application of high
pressure in the PECS, these ceramics were prepared in a dense
form (98%1) with a grain size of o20 nm.104 It was discovered
that in this grain size, YSZ is a protonic conductor in the presence of moisture.153 This can be seen in Fig. 40, in which the
decrease in the oxygen ion conductivity with decreasing temperature is reversed at low temperature due to the contribution
of protonic conductivity.
The examples briey discussed above highlight the success of
the PECS process in the preparation of highly dense materials
with a very small grain size, an accomplishment that opens the
door for application in fuel cells at low temperatures.

References
1

J. E. Burke, A History of the Development of a Science of Sintering, Ceram.


Civilizat., 1, 31533 (1985).
2
J. A. Pero-Sanz Alroz, J. I. Verdeja-Gonzalez, J. P. Sanch-Martinez, and N.
Vilela, Melting and Sintering Platinum in the 18th Century: The Secret of the
Spanish, JOM, 51 [10] 913 (1999).
3
Z. A. Munir, U. Anselmi-Tamburini, and M. Ohyanagi, The Effect of Electric
Field and Pressure on the Synthesis and Consolidation of Materials: A Review of
the Spark Plasma Sintering Method, J. Mater. Sci., 41, 76377 (2006).
4
S. Grasso, Y. Sakka, and G. Maizza, Electric Current Activated/Assisted
Sintering (ECAS): A Review of Patents 19062008, Sci. Technol. Adv. Mater., 10,
053001, 24pp (2009).
5
M. Omori, Sintering, Consolidation, Reaction and Crystal Growth by the
Spark Plasma System (SPS), Mater. Sci. Eng., A287, 1838 (2000).

January 2011
6

Electric Current Activation of Sintering

M. Nygren and Z. Shen, Spark Plasma Sintering: Possibilities and Limitations, Key Eng. Mater., 264268, 71924 (2004).
7
T. Hungria, J. Galy, and A. Castro, Spark Plasma Sintering as a Useful
Technique to the Nanostructuration of Piezo-ferroelectric Materials, Adv. Eng.
Mater., 11 [8] 61531 (2009).
8
R. Orru, R. Licheri, A. M. Locci, A. Cincotti, and G. Cao, Consolidation/
Synthesis of Materials by Electric Current Activated/Assisted Sintering, Mater.
Sci. Eng. Reports, R63 [46] 127287 (2009).
9
J. E. Garay, Current-activated, Pressure-assisted Densication of Materials,
Annu. Rev. Mater. Res., 40, 44568 (2010).
10
J. Gubicza, H-Q. Bui, F. Fellah, and G. F. Dirras, Microstructure and Mechanical Behavior of Ultrane-grained Ni Processed by Different Powder Metallurgy Methods, J. Mater. Res., 24 [1] 21726 (2009).
11
R. Ritasalo, M. E. Cura, X. W. Liu, O. Soderberg, T. Ritvonen, and S.-P.
Hannula, Spark Plasma Sintering of Submicron-sized Cu-powder Inuence of
Processing Parameters and Powder Oxidization on Microstructure and Mechanical Properties, Mater. Sci. Eng. A, A527, 27337 (2010).
12
T. Mizuguchi, S. Guo, and Y. Kagawa, Transmission Electron Microscopy
Characterization of Spark Plasma Sintered ZrB2 Ceramic, Ceram. Int., 36, 9436
(2010).
13
K. Kanamori, T. Kineri, R. Fukuda, K. Nishio, M. Hashimoto, and H. Mae,
Spark Plasma Sintering of Sol-gel Derived Amorphous ZrW2O8, J. Am. Ceram.
Soc., 92, 325 (2009).
14
S-Q. Guo, T. Nishimura, Y. Kagawa, and J-M. Yang, Spark Plasma Sintering of Zirconium Diborides, J. Am. Ceram. Soc., 91, 284855 (2010).
15
C. Musa, R. Licheri, A. M. Locci, R. Orru, G. Cao, M. A. Rodriguez, and L.
Jaworska, Energy Efciency During Conventional and Novel Sintering Processes: The Case of TiAl2O3TiC Composites, J. Cleaner Prod., 17, 87782
(2009).
16
M. Eriksson, H. Yan, M. Nygren, M. J. Reece, and Z. Shen, Low Temperature Consolidated Lead-free Ferroelectric Niobate Ceramics with Improved Electrical Properties, J. Mater. Res., 25 [2] 2407 (2010).
17
S. Le Gallet, L. Campayo, E. Courtois, S. Hoffmann, Y. Grin, F. Bernard,
and F. Bart, Spark Plasma Sintering of Iodine-bearing Apatite, J. Nuclear
Mater., 400, 2516 (2010).
18
L. Campyao, S. Le Gallet, Y. Grin, E. Courtois, F. Bernard, and F. Bart,
Spark Plasma Sintering of Lead Phosphovanadate Pb3(VO4)1.6(PO4)0.4, J. Eur.
Ceram. Soc., 29, 147784 (2009).
19
Y. Kan, P. Wang, T. Xu, G. Zhang, D. Yan, Z. Shen, and Y-B. Cheng,
Spark Plasma Sintering of Bismuth Titanate Ceramics, J. Am. Ceram. Soc., 88
[6] 16313 (2005).
20
C. Drouet, C. Largeot, G. Raimbeaux, C. Estournes, G. Dechambre, C.
Combes, and C. Rey, Bioceramics: Spark Plasma Sintering (SPS) of Calcium
Phosphates, Adv. Sci. Technol., 49, 4550 (2006).
21
A. Bassano, M. T. Buscaglia, V. Buscaglia, and P. Nanni, Particle Size and
Morphology Control of Perovskite Oxide Nanopowders for Nanostructured
Materials, Integrated Ferroelectr., 109, 117 (2009).
22
M. Okamotoa, Y. Akimunea, K. Furuyab, M. Hatanoc, M. Yamanakac, and
M. Uchiyamac, Phase Transition and Electrical Conductivity of Scandia-stabilized Zirconia Prepared by Spark Plasma Sintering Process, Solid State Ionics,
176, 67580 (2005).
23
M. Nygren and Z. Shen, Novel Assemblies via Spark Plasma Sintering,
Silicates Ind., 69 [78] 2118 (2004).
24
B. Basu, J-H. Lee, and D-Y. Kim, Development of Nanocrystalline Wearresistant Y-TZP Ceramics, J. Am. Ceram. Soc., 87 [9] 17714 (2004).
25
M. Yue, J. X. Zhang, W. Q. Liu, and G. P. Wang, Chemical Stability
and Microstructure of NdFeB Magnet Prepared by Spark Plasma Sintering,
J. Magn. Magn. Mater., 271 [23] 3648 (2004).
26
X. Su, P. Wang, W. Chen, Z. Shen, M. Nygren, Y. Cheng, and D. Yan, Optical Properties of SPS-ed Y- and (Dy,Y)-a-Sialon Ceramics, J. Mater. Sci., 39
[20] 625762 (2004).
27
Y. H. Han, M. Nagata, N. Uekawa, and K. Kakegawa, Eutectic Al2O3
GdAlO3 Composite Consolidated by Combined Rapid Quenching and Spark
Plasma Sintering Technique, Br. Ceram. Trans., 103 [5] 21922 (2004).
28
Y.-S. Kwon, D. V. Dudina, M. A. Korchagin, and O. I. Lomovsky, Microstructure Changes in TiB2Cu Nanocomposite Under Sintering, J. Mater. Sci.,
39 [1617] 532531 (2004).
29
L. Zhou, Z. Zhao, A. Zimmermann, F. Aldinger, and M. Nygren, Preparation and Properties of Lead Zirconate Stannate Titanate Sintered by Spark Plasma
Sintering, J. Am. Ceram. Soc., 87 [4] 60611 (2004).
30
S. H. Risbud, J. R. Groza, and M. J. Kim, Clearn Grain Boundaries in
Aluminum Nitride Ceramics Densied Without Additives by a Plasma-Activated
Sintering Process, Philos. Mag., B 69 [3] 52533 (1994).
31
Z. Shen, H. Peng, and M. Nygren, Formidable Increase in the Superplasticity
of Ceramics in the Presence of an Electric Field, Adv. Mater., 15 [12] 10069
(2003).
32
X. J. Chen, K. A. Khor, S. H. Chan, and L. G. Yu, Overcoming the Effect of
Contaminant in Solid Oxide Fuel Cell (SOFC) Electrolyte: Spark Plasma Sintering
(SPS) of 0.5 wt.% Silica-doped Yttria-stabilized Zirconia (YSZ), Mater. Sci.
Eng., A374 [12] 6471 (2004).
33
T. Takeuchi, E. Betourne, M. Tabuchi, H. Kageyama, Y. Kobayashi, A.
Coats, F. Morrison, D. C. Sinclair, and A. R. West, Dielectric Properties of
Spark-plasma-sintered BaTiO3, J. Mater. Sci., 34 [5] 91724 (1999).
34
M. Yue, J. Zhang, Y. Xiao, G. Wang, and T. Li, New Kind of NdFeB
Magnet Prepared by Spark Plasma Sintering, IEEE Trans. Magn., 39 [6] 35513
(2003).
35
P. Pei, X. P. Song, J. Liu, M. Zhao, and G. L. Chen, Improving Hydrogen Storage Properties of Laves Phase Related BCC Solid Solution
Alloy by SPS Preparation Method, Intern. J. Hydrogen Storage, 34, 8597602
(2009).

36

17

X. B. Zhao, S. H. Yang, Y. Q. Cao, J. L. Mi, Q. Zhang, and T. J. Zhu,


Synthesis of Nanocomposites with Improved Thermoelectric Properties, J. Electron. Mater., 38 [7] 101724 (2009).
37
J. G. Noudem, M. Prevel, A. Veres, D. Chateigner, and J. Galy, Thermoelectric Ca3Co4O9 Ceramics Consolidated by Spark Plasma Sintering, J. Electroceram., 22, 917 (2009).
38
K. Amezawa, Y. Nishikawa, Y. Tomii, and N. Yamamoto, Electrical and
Mechanical Properties of Sr-doped LaPO4 Prepared by Spark Plasma Sintering,
J. Electrochem. Soc., 152, A10607 (2005).
39
K. Morita, B-N. Kim, H. Yoshida, and K. Hiraga, Spark-Plasma-Sintering
Condition Optimization for Producing Transparent MgAl2O4 Spinel Polycrystal,
J. Am. Ceram. Soc., 92 [6] 120816 (2009).
40
S. Mula, K. Mondal, S. Ghosh, and S. K. Pabic, Structure and Mechanical
Properties of AlNiTi Amorphous Powder Consolidated by Pressure-less, Pressure-assisted and Spark Plasma Sintering, Mater. Sci. Eng., A527, 375763
(2010).
41
T. B. Holland, J. F. Lofer, and Z. A. Munir, Crystallization of Metallic
Glasses Under the Inuence of High-density d.c. Current, J. Appl. Phys., 95 [5]
28969 (2004).
42
L. Wang, W. Jiang, L. Chen, and Z. Shen, Formation of a Unique Glass by
Spark Plasma Sintering of a Zeolite, J. Mater. Res., 24 [10] 32415 (2009).
43
M. Ohyanagi, T. Yamamoto, H. Kitaura, Y. Kodera, T. Ishii, and Z. A. Munir, Consolidation of Nanostructured b-SiC with Disorder-Order Transformation, Scr. Mater., 50, 1114 (2004).
44
N. Toyofuku, M. Nishimoto, K. Arayama, Y. Kodera, M. Ohyanagi, and
Z. A. Munir, Consolidation of Carbon with the Amorphous-graphite Transformation by SPS, Ceram. Trans., 212, 3140 (2010).
45
R. Orru, J. Woolman, G. Cao, and Z. A. Munir, Synthesis of Dense Nanometric MoSi2 Through Mechanical and Field Activation, J. Mater. Res., 16 [5]
143948 (2001).
46
F. Bernard, F. Charlot, E. Gaffet, and Z. A. Munir, One-step Synthesis and
Consolidation of Nanophase Iron Aluminide, J. Am. Ceram. Soc., 84 [5] 9104
(2001).
47
J. W. Lee, Z. A. Munir, M. Shibuya, and M. Ohyanagi, Synthesis of Dense
TiB2TiN Nanocrystalline Composites Through Mechanical and Field Activation, J. Am. Ceram. Soc., 84 [6] 120916 (2001).
48
J. N. Woolman, J. J. Petrovic, and Z. A. Munir, Incorporating Mg into the Si
Sub-Lattice of Molybdenum Disilicide, Scr. Mater., 48 [6] 81924 (2003).
49
U. V. Waghmare, V. Bulatov, E. Kaxiras, and M. S. Duesbery, Microalloying for Ductility in Molybdenum Disilicide, Mater. Sci. Eng., A261 [12] 1475
(1999).
50
M. Tokita, Development of Square-Shaped Large-Size WC/Co/Ni System
FGM Fabricated by Spark Plasma Sintering (SPS) Method and its Industrial
Applications, Mater. Sci. Forum, 4923, 7118 (2005).
51
E. M. Heian, J. C. Gibeling, and Z. A. Munir, Synthesis and Characterization of Nb5Si3/Nb Functionally Graded Composites, Mater. Sci. Eng., A368,
16874 (2004).
52
Q. S. Meng, S. P. Chen, J. F. Zhao, and Z. A. Munir, Synthesis and Characterization of TiB2NiNi3AlCrNi Alloy Graded Material by Field-activated
Combustion, J. Alloys Compd., 476, 88993 (2009).
53
T. Sui, J-F. Li, and S. Jin, Joining CoSb3 to Metal Surface of FGM Electrode
for Thermoelectric Modules by SPS, Key Eng. Mater., 36872, 185861 (2008).
54
A. Wang and O. Ohashi, Titanium Mesh/rod Joined by Pulse Electric Current Sintering: Effect of Heating Rate, Mater. Trans., 47, 234852 (2006).
55
T. Nakamura, K. Hayakawa, S. Tanaka, H. Imaizumi, and Y. Nakagawa,
Bonding Characteristics of Various Metals by DC Pulse Resistance Heat Pressure
Welding, Mater. Trans., 46 [2] 2927 (2005).
56
U. Anselmi-Tamburini, J. E. Garay, and Z. A. Munir, Fundamental Investigations on the Spark Plasma Sintering/synthesis Process III. Current Effect on
Reactivity, Mater. Sci. Eng., A407, 2430 (2005).
57
M. Tokita, Development of Advanced Spark Plasma Sintering (SPS) Systems
and its Industrial Applications, Ceram. Trans., 194, 519 (2006).
58
T. Misawa, N. Shikani, Y. Kawakami, T. Enjoji, and Y. Ohtsu, Inuence of
Internal Pulsed Current on the Sintering Behavior of Pulsed Current Sintering
Process, Mater. Sci. Forum, 63842, 210914 (2010).
59
D. M. Hulbert, A. Anders, D. V. Dudina, J. Andersson, D. Jiang, C. Unuvar,
U. Anselmi-Tamburini, E. J. Lavernia, and A. K. Mukherjee, The Absence of
Plasma in Spark Plasma Sintering, J. Appl. Phys., 104 [3] 033305, 7pp (2008).
60
T. Misawa, N. Shikatani, Y. Kawakami, T. Enjoji, Y. Ohtsu, and H. Fujita,
Observation of Internal Pulsed Current Flow Through the ZnO Specimen in the
Spark Plasma Sintering Method, J. Mater. Sci., 44, 16415 (2009).
61
M. Nanko, T. Maruyama, and H. Tomino, Neck Growth on Initial Stage of
Pulse Current Pressure Sintering for Coarse Atomized Powder Made of CastIron, J. Jpn. Inst. Metals, 63 [7] 91723 (1999).
62
M. Nanko, T. Oyaidu, and T. Maruyama, Densication of Ni-20Cr Alloy
Coarse-Powder by Pulse Current Pressure Sintering, J. Jpn. Inst. Metals, 66 [2]
8793 (2002).
63
G. Xie, O. Ohashi, K. Chiba, N. Yamaguchi, M. Song, K. Furuya, and T.
Noda, Frequency Effect on Pulse Electric Current Sintering Process of Pure
Aluminum Powder, Mater. Sci. Eng., A359, 38490 (2003).
64
K. Q. Dang, M. Kawahara, S. Takei, and M. Nanko, Effects of Pulsed Current Waveforms on Sample Temperature and Sintering Behavior in PECS of
Alumina, J. Jpn. Soc. Powder Metall., 56 [12] 7807 (2009).
65
W. Chen, U. Anselmi-Tamburini, J. E. Garay, J. R. Groza, and Z. A. Munir,
Fundamental Investigations on the Spark Plasma Sintering/Synthesis Process,
Mater. Sci. Eng., A394 [12] 1328 (2005).
66
U. Anselmi-Tamburini, S. Gennari, J. E. Garay, and Z. A. Munir, Fundamental Investigations on the Spark Plasma Sintering/Synthesis Process II.
Modeling of Current and Temperature Distributions, Mater. Sci. Eng., A394
[12] 13948 (2005).

18
67

Journal of the American Ceramic SocietyMunir et al.

U. Anselmi-Tamburini, Z. A. Munir, and J. E. Garay, Preparation of Dense


Nanostructured Oxide Ceramics with Fine Crystal Size by High-Pressure Spark
Plasma Sintering; U.S. Patent No. 7,601,403, October 13, 2009, 10pp.
68
A. Wang and O. Ohashi, Preparation of Dense Nanostructured Functional
Oxide Materials with Fine Crystallite Size by Field Activation Sintering, Mater.
Trans., 47 [9] 234852 (2006).
69
H. T. Orchard and A. L. Greer, Electromigration Effects on Compound
Growth at Interfaces, Appl. Phys. Lett., 86 [23] 231906, 3pp (2005).
70
P. Asoka-Kumar, K. OBrien, K. G. Lynn, P. J. Simpson, and K. P. Rodbell,
Detection of Current-Induced Vacancies in Thin Aluminum-copper Lines Using
Positrons, Appl. Phys. Lett., 68 [3] 4068 (1996).
71
J. E. Garay, S. C. Glade, U. Anselmi-Tamburini, P. Asoka-Kumar, and Z. A.
Munir, Electric Current Enhanced Defect Mobility in Ni3Ti Intermetallics,
Appl. Phys. Lett., 85 [4] 5735 (2004).
72
N. Bertolino, J. Garay, U. Anselmi-Tamburini, and Z. A. Munir, Electromigration Effects in AlAu Multilayers, Scr. Mater., 44 [5] 73742 (2001).
73
N. Bertolino, J. Garay, U. Anselmi-Tamburini, and Z. A. Munir, High-ux
Current Effects in Interfacial Reactions in AuAl Multilayers, Philos. Mag. B, 82
[8] 96985 (2002).
74
J. E. Garay, U. Anselmi-Tamburini, and Z. A. Munir, Enhanced Growth of
Intermetallic Phases in the NiTi System by Current Effects, Acta Mater., 51 [15]
448795 (2003).
75
J. Zhao, J. E. Garay, U. Anselmi-Tamburini, and Z. A. Munir, Directional
Electromigration-Enhanced Interdiffusion in the CuNi System, J. Appl. Phys.,
102 [11] 114902, 7pp (2007).
76
J. F. Zhao, C. Unuvar, U. Anselmi-Tamburini, and Z. A. Munir, Kinetics of
Current-enhanced Dissolution of Nickel in Liquid Aluminum, Acta Mater., 55
[16] 5592600 (2007).
77
J. Zhao, C. Unuvar, U. Anselmi-Tamburini, and Z. A. Munir, Microstructural Evolution During the Dissolution of Nickel in Liquid Aluminum Under the
Inuence of an Electric Field, Acta Mater., 56 [8] 18408 (2008).
78
H. Conrad and D. Yang, Inuence of an Applied dc Electric Field on the
Plastic Deformation Kinetics of Oxide Ceramics, Philos. Mag., 90 [9] 114157
(2010).
79
K. Jung and H. Conrad, Retardation of Grain Growth in Electrodeposited
Cu by an Electric Field, J. Mater. Sci., 42, 39944003 (2007).
80
S. Starnes and H. Conrad, Grain Size Distribution in Ultrane-grained
Yttria-stabilized Zirconia Deformed Without and With an Electric Field, Scr.
Mater., 59, 11158 (2008).
81
S. Ghosh, A. H. Chokshi, P. Lee, and R. Raj, A Huge Effect of Weak dc
Electrical Fields on Grain Growth in Zirconia, J. Am. Ceram. Soc., 92 [8] 18569
(2009).
82
K. Chen, X. Zhang, H. Wang, L. Zhang, J. Zhu, F. Yang, and L. An, Making Nanostructured Ceramics from Micrometer-sized Powders via Grain Renement During SPS Sintering, J. Am. Ceram. Soc., 91 [8] 247580 (2008).
83
T. Nagae, M. Yokota, M. Nose, S. Tomida, T. Kamiya, and S. Saji, Effects
of Pulse Current on an Aluminum Powder Oxide Layer During Pulse Current
Pressure Sintering, Mater. Trans., 43 [6] 13907 (2002).
84
A. Mussi, G. Bernard Granger, A. Addad, N. Benameur, F. Beclin, and A.
Bataille, Inversion Defects in MgAl2O4 Elaborated by Pressureless Sintering,
Pressureless Sintering Plus Hot Isostatic Pressing, and Spark Plasma Sintering,
Scr. Mater., 61, 5169 (2009).
85
N. Nuns, F. Beclin, and J. Crampon, Grain-Boundary Characterization
in a Nonstoichiometric Fine-Grained Magnesium Aluminate Spinel: Effects of
Defect Segregation at the Space-charge Layers, J. Am. Ceram. Soc., 92 [4] 8705
(2009).
86
A. Bataille, A. Addad, C. Courtois, T. Duhoo, and J. Crampon, Solute and
Defect Segregation at the Space Charge Layers of Fe-Doped Fine-Grained Al2O3:
Effect on the Creep Rate, J. Eur. Ceram. Soc., 28, 112934 (2008).
87
T. Kondo, M. Yasuhara, T. Kuramoto, Y. Kodera, M. Ohyanagi, and Z. A.
Munir, Effect of Pulsed dc Current on Atomic Diffusion of NbC Diffusion
Couple, J. Mater. Sci., 43, 64005 (2008).
88
T. Kondo, T. Kuramoto, Y. Kodera, M. Ohyanagi, and Z. A. Munir, Enhanced Growth of Mo2C Formed in MoC Diffusion Couple by Pulsed dc Current, J. Jpn. Soc. Powder Metall., 55, 64350 (2008).
89
T. Kondo, T. Kuramoto, Y. Kodera, M. Ohyanagi, and Z. A. Munir, Inuence of Pulsed dc Current and Electric Field on Growth of Carbide Ceramics
During Spark Plasma Sintering, J. Ceram. Soc. Jpn., 116, 118792 (2008).
90
J. M. Frei, U. Anselmi-Tamburini, and Z. A. Munir, Current Effects on
Neck Growth in the Sintering of Copper Spheres to Copper Plates by the Pulsed
Electric Current Method, J. Appl. Phys., 101, 114914, 8pp (2007).
91
J. R. Friedman, J. E. Garay, U. Anselmi-Tamburini, and Z. A. Munir, Modied Interfacial Reactions in AgZn Multilayers Under the Inuence of High d.c.
Currents, Intermetallics, 12 [6] 58997 (2004).
92
Z. Fu, K. Wang, T. Tan, Y. Xiong, D. He, Y. Wang, and Z. A. Munir, Study
on the Process Mechanism in Spark Plasma Sintering, Ceram. Trans., 194, 321
(2006).
93
J. E. Burke and D. Turnbull, Recrystallization and Grain Growth, Progr. In
Metal Phys., 3, 22092 (1952).
94
C. Shearwood and H. B. Ng, Spark Plasma Sintering of Wire Exploded
Tungsten Nano-Powder; pp. 67981B-110 in Microelectronics: Design, Technology, and Packaging III, Proc. SPIE 6798, Edited by A. J. Hariz, and
V. K. Varadan, Canberra, Australia, 2007.
95
D. Y. Kim, G. Gladel, and A. Accary, Morphological Study of Tungsten
Powder Obtained by Hydrogen Reduction of Tungsten Trioxide Powder at 700
9001C, from Prepr. Eu. Symp. Powder Metall., 5th, 2, 185193 (1978).
96
H. B. Hungtinton, Diffusion in Solids; in Edited by A. S. Nowick, and J. J.
Burton. Academic Press, New York, NY, 1975.
97
R. M. German, Sintering Theory and Practice. Wiley, New York, NY, 1996,
p. 170.

98

Vol. 94, No. 1

R. M. German, High Density Powder Processing Using Pressure-Assisted


Sintering, Rev. Particular Mater., 2, 11772 (1994).
99
J. Jamnik and R. Raj, Space-Charge-Controlled Diffusional Creep: Volume
Diffusion Case, J. Am. Ceram. Soc., 79 [1] 1938 (1996).
100
Y. Makino, M. Sakaguchi, J. Terada, and K. Akamatsu, Consolidation of
Ultrane Alumina Powders with SPS Method, J. Jpn. Soc. Powder Metall., 54,
21925 (2007).
101
F. Guillard, A. Allemand, J-D. Lulewicz, and J. Galy, Densication of SiC
by SPS Effects of Time, Temperature and Pressure, J. Eur. Ceram. Soc., 27,
27258 (2007).
102
R. Chaim and Z. Shen, Grain Size Control by Pressure Application Regime
During Spark Plasma Sintering of NdYAG Nanopowders, J. Mater. Sci., 43,
50237 (2008).
103
R. Chaim and M. Margulis, Densication Maps for Spark Plasma Sintering
of Nanocrystalline MgO Ceramics, Mater. Sci. Eng., A407, 1807 (2005).
104
U. Anselmi-Tamburini, J. E. Garay, and Z. A. Munir, Fast Low-temperature Consolidation of Bulk Nanometric Ceramic Materials, Scripta Mater., 54,
8238 (2006).
105
D. V. Quach, H. Avila-Paredes, S. Kim, M. Martin, and Z. A. Munir, Pressure Effects and Grain Growth Kinetics in the Consolidation of Nanostructured
Fully Stabilized Zirconia by Pulsed Electric Current Sintering, Acta Mater., 58,
502230 (2010).
106
R. J. Stokes and D. Evans, Fundamentals of Interfacial Engineering, pp.
2433. Wiley, New York, NY, 1996.
107
G. Y. Onoda and J. Toner, Fractal Dimensions of Model Particle Packings
Having Multiple Generations of Agglomerates, J. Am. Ceram. Soc., 69, C2789
(1986).
108
M. A. C. G. Van de Graaf, J. H. H. Ter Maat, and A. J. Burgraaf, Microstructure and Sintering Kinetics of Highly Reactive ZrO2Y2O3, J. Mater. Sci.,
20, 140718 (1985).
109
S. Grasso, Y. Sakka, and G. Maizza, Pressure Effects on Temperature Distribution During Spark Plasma Sintering with Graphite Sample, Mater. Trans.,
50 [8] 21114 (2009).
110
A. Zavaliangos, J. Zhang, M. Krammer, and J. R. Groza, Temperature
Evolution During Field Activated Sintering, Mater. Sci. Eng., A379, 2182
(2004).
111
K. Vanmeensel, A. Laptev, J. Hennicke, J. Vleugels, and O. Van der Biest,
Modelling of the Temperature Distribution During Field Assisted Sintering,
Acta Mater., 53, 437988 (2005).
112
J. Xu, S. R. Casolco, and J. E. Garay, Effect of Varying Displacement Rates
on the Densication of Nanostructured Zirconia by Current Activation, J. Am.
Ceram. Soc., 92 [7] 150613 (2009).
113
R. M. German, Sintering Theory and Practice, p. 482. Wiley, New York, NY,
1996.
114
E. A. Olevsky, S. Kandukuri, and L. Froyen, Consolidation Enhancement
in Spark-plasma Sintering: Impact of High Heating Rates, J. Appl. Phys., 102,
114913, 12pp (2007).
115
L. A. Stanciu, V. Y. Kodash, and J. R. Groza, Effects of Heating Rate on
Densication and Grain Growth During Field-assisted Sintering of a-Al2O3 and
MoSi2 Powders, Metall. Mater. Trans. A, 32A, 26338 (2001).
116
Z. Shen, M. Johnsson, Z. Zhao, and M. Nygren, Spark Plasma Sintering of
Alumina, J. Am. Ceram. Soc., 85 [8] 19217 (2002).
117
Y. Zhou, K. Hirao, Y. Yamauchi, and S. Kanzaki, Effects of Heating Rate
and Particle Size on Pulse Electric Current Sintering of Alumina, Scr. Mater., 48,
16316 (2003).
118
U. Anselmi-Tamburini, J. E. Garay, Z. A. Munir, A. Tacca, F. Maglia, and
G. Spinolo, Spark Plasma Sintering and Characterization of Bulk Nanostructured Fully Stabilized Zirconia: Part I. Densication Studies, J. Mater. Res., 19
[11] 325562 (2004).
119
K. Matsugi, H. Kuramoto, O. Yanagisawa, and M. Kiritani, A Case Study
for Production of Perfectly Sintered Complex Compacts in Rapid Consolidation
by Spark Sintering, Mater. Sci. Eng., A354, 23442 (2003).
120
B. McWilliams and A. Zavaliangos, Multi-Phenomena Simulation of Electric Field Assisted Sintering, J. Mater. Sci., 43, 50315 (2008).
121
E. A. Olevsky and L. Froyen, Impact of Thermal Diffusion on Densication
During SPS, J. Am. Ceram. Soc., 92 [S1] S12232 (2009).
122
E. Olevsky and L. Froyen, Constitutive Modeling of Spark-plasma Sintering
of Conductive Materials, Scr. Mater., 55, 11758 (2006).
123
J. Rathel, M. Herrmann, and W. Beckert, Temperature Distribution for
Electrically Conductive and Non-conductive Materials During Field-Assisted
Sintering (FAST), J. Eur. Ceram. Soc., 29, 141925 (2009).
124
D. Tiwari, B. Basu, and K. Biswas, Simulations of Thermal and Electric
Field Evolution During Spark Plasma Sintering, Ceram. Int., 35, 699708
(2009).
125
A. Cincotti, A. M. Locci, R. Orru, and G. Cao, Modeling of SPS Apparatus: Temperature, Current and Strain Distribution with no Powders, A. I. Ch.
E. J., 53 [3] 70319 (2007).
126
R. S. Dobedoe, G. D. West, and M. H. Lewis, Spark Plasma Sintering
of Ceramics: Understanding Temperature Distribution Enables More Realistic
Comparison with Conventional Processing, Adv. Appl. Ceram., 104 [3] 1106
(2005).
127
N. Chennou, G. Majkic, Y. C. Chen, and K. Salama, Temperature, Current, and Heat Loss Distributions in Reduced Electrothermal Loss Spark Plasma
Sintering, Metall. Mater. Trans. A, 40A, 24019 (2009).
128
X. Liu, X. Song, J. Zhang, and S. Zhao, Temperature Distribution and
Neck Formation of WC-Co Combined Particles During Spark Plasma Sintering,
Mater. Sci. Eng., A488, 17 (2008).
129
G. Maizza, S. Grasso, and Y. Sakka, Moving Finite-element Mesh Model
for Aiding Spark Plasma Sintering in Current Control Mode of Pure Ultrane WC
Powder, J. Mater. Sci., 44, 121936 (2009).

January 2011

Electric Current Activation of Sintering

130

19

142

X. Wang, S. R. Casolco, G. Xu, and J. E. Garay, Finite Element Modeling


of Electric Current-activated Sintering: The Effect of Coupled Electrical Potential,
Temperature and Stress, Acta Mater., 55, 361122 (2007).
131
K. Vanmeensel, A. Laptev, O. Van der Biest, and J. Vleugels, Field Assisted
Sintering of Electro-Conductive ZrO2-based Composites, J. Eur. Ceram. Soc., 27,
97985 (2007).
132
K. Vanmeensel, A. Laptev, O. Van der Biest, and J. Vleugels, The Inuence
of Percolation During Pulsed Electric Current Sintering of ZrO2TiN Powder
Compacts with Varying TiN Content, Acta Mater., 55, 18011 (2007).
133
S. Grasso, Y. Sakka, G. Maizza, and C. Huz, Pressure Effect on the Homogeneity of Spark Plasma-Sintered Tungsten Carbide Powder, J. Am. Ceram.
Soc., 92 [10] 241821 (2009).
134
R. Apetz and P. B. van Bruggen, Transparent Alumina: A Light-scattering
Model, J. Am. Ceram. Soc., 86 [3] 4806 (2003).
135
U. Anselmi-Tamburini, J. N. Woolman, and Z. A. Munir, Transparent
Nanometric Cubic and Tetragonal Zirconia Obtained by High-pressure Pulsed
Electric Current Sintering, Adv. Funct. Mater., 17, 326773 (2007).
136
B-N. Kim, K. Hiraga, K. Morita, H. Yoshida, T. Miyazaki, and Y. Kagawa,
Microstructure and Optical Properties of Transparent Alumina, Acta Mater.,
57, 131926 (2009).
137
Y. Xiong, Z. Fu, Y. Wang, and F. Quan, Fabrication of Transparent AlN
Ceramics, J. Mater. Sci., 41, 25379 (2006).
138
D. Jiang, D. M. Hulbert, U. Anselmi-Tamburini, T. Ng, D. Land, and A. K.
Mukherjee, Optically Transparent Polycrystalline Al2O3 Produced by Spark
Plasma Sintering, J. Am. Ceram. Soc., 91 [1] 1514 (2008).
139
G. Zhang, Y. Wang, Z. Fu, H. Wang, W. Wang, J. Zhang, S. W. Lee, and K.
Nihara, Transparent Mullite Ceramic from Single-Phase Gel by Spark Plasma
Sintering, J. Eur. Ceram. Soc., 29, 270511 (2009).
140
R. Chaim, Z. Shen, and M. Nygren, Transparent Nanocrystalline MgO by
Rapid and Low-Temperature Spark Plasma Sintering, J. Mater. Res., 19 [9]
252731 (2004).
141
K. Morita, B-N. Kim, K. Hiraga, and H. Yoshida, Fabrication of Highstrength Transparent MgAl2O4 Spinel Polycrystals by Optimizing Spark-PlasmaSintering Conditions, J. Mater. Res., 24 [9] 286372 (2009).

R. Chaim, R. Marder, and C. Estournes, Optically Transparent Ceramics by


Spark Plasma Sintering of Oxide Nanoparticles, Scripta Mater., 63, 2114 (2010).
143
R. Chaim, M. Kalina, and J. Z. Shen, Transparent Yttrium Aluminum
Garnet (YAG) Ceramics by Spark Plasma Sintering, J. Eur. Ceram. Soc., 27,
33317 (2007).
144
J. E. Alaniz, F. G. Perez-Gutierrez, g. Aguilar, and J. E. Garay, Optical
Properties of Transparent Nanocrystalline Yttria Stabilized Zirconia, Opt.
Mater., 32, 628 (2009).
145
W. Kun, F. Zhengyi, W. Weimin, W. Yucheng, Z. Jinyong, and Z. Qingjie,
Study on Fabrication and Mechanism in of Porous Metals by Spark Plasma
Sintering, J. Mater. Sci., 42, 3026 (2007).
146
M-J. Suk, W-S. Seo, and Y-S. Kwon, Fabrication of Graded Porous Structure with Pore size Distribution by SPS Process, Mater. Sci. Forum, 534536,
9658 (2007).
147
Y. Zhao and M. Taya, Processing of Porous NiTi by Spark Plasma Sintering, Proc. SPIE, 6170, 617013, 6pp (2006).
148
R. Nicula, F. Luethen, M. Stir, B. Nebe, and E. Burkel, Spark Plasma
Sintering Synthesis of Porous Nanocrystalline Titanium Alloys for Biomedical
Applications, Biomol. Eng., 24, 5647 (2007).
149
D. Kawagoe, R. Sawai, and T. Ishiduka, Preparation of Porous Hydroxyapatite Ceramics by Spark Plasma Sintering, Trans. Mater. Res. Soc. Jpn., 33 [4]
9114 (2008).
150
F. Zhang, K. Lina, J. Changa, J. Lua, and C. Ning, Spark Plasma Sintering
of Macroporous Calcium Phosphate Scaffolds from Nanocrystalline Powders,
J. Eu. Ceram. Soc., 28, 53945 (2008).
151
D. Chakravarty, H. Ramesh, and T. N. Rao, High Strength Porous
Alumina by Spark Plasma Sintering, J. Eur. Ceram. Soc., 29, 13619 (2009).
152
P. Dibandjo, L. Bois, C. Estournes, B. Durand, and P. Miele, Silica, Carbon
and Boron Nitride Monoliths with Hierarchical Porosity Prepared by Spark
Plasma Sintering Process, Micropor. Mesopor. Mater., 111, 6438 (2008).
153
S. Kim, U. Anselmi-Tamburini, H. J. Park, M. Martin, and Z. A. Munir,
Unprecedented Room-temperature Electrical Power Generation Using
Nanoscale Fluorite-structured Oxide Electrolytes, Adv. Mater., 20, 5569
(2008).
&

Zuhair A. Munir is a Distinguished Research Professor in the


Department of Chemical Engineering and Materials Science
and former Dean of the College
of Engineering at the University of
California, Davis (UC Davis). He
received all his degrees from the
University of California, Berkeley.
His research is in the general area
of kinetics and thermodynamics of
materials processing with emphasis on eld-activated processes. Professor Munir is the recipient
of numerous awards and honors including the Societys John
Jeppson Award and the Outstanding Educator Award. He
received the Gold Medal from the Russian Academy of Sciences
for contribution to the science of self-propagating high-temperature synthesis (SHS), and received the UC Davis Prize, the
campus highest award for scholarly distinction and outstanding
teaching. Professor Munir has published more than 475 papers
and holds 13 U.S. patents. He is listed as a Highly Cited Author
in Materials Science. His service for the Society includes Associate Editor and member of the Task Force on Globalization and
the Phase Equilibria Subcommittee. He also served as Editor-inChief for the Journal of Materials Synthesis and Processing as
Principal Editor for the Journal of Materials Research, and as
Editor for the Journal of Materials Science.

interested in research on alternative energy sources and has


worked on template-based synthesis of thin lms and nanowires
of skuterrudites for thermoelectric applications. Quach was a
bronze medalist at the 29th International Chemistry Olympiad
in Montreal, Canada. He graduated with a combined B.S. in
Mechanical Engineering + Materials Science and recently
received a Ph.D. in Materials Science and Engineering.

Dr. Dat V. Quach is a Postdoctoral Research Fellow in the Department of Chemical Engineering
and Materials Science at the University of California, Davis. His
main research focuses on chemical
changes and phase stability of nanostructured materials processed
under electric eld applications
and on the inuence of external
elds on consolidation. He is also

Manshi Ohyanagi is Dean of the


Faculty of Science and Technology and Director of Innovative
Materials and Processing Research Center at Ryukoku University in
Ohtsu, Japan. He
received his doctorate in Applied
Chemistry at Waseda University
in 1988 and spent a year as a
postdoctoral fellow at Caltech.
After that, he joined the Department of Materials Chemistry at
Ryukoku University as Assistant Professor. In 2001, he was
promoted to Full Professor. In 2002-2003, he was a Visiting
Professor at the University of California, Davis. His current
research focuses on Spark Plasma Sintering (SPS) of nanostructured ceramics, with emphasis on the sintering of high temperature materials with stacking disorder-order transformation. His
research also includes the processing of hydrogen storage materials. He served as coorganizer of several symposia on SPS,
associated with Pac Rim conferences. He has published 4110
papers and holds 10 U.S. patents.

Вам также может понравиться