Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Spark Plasma Sintering

Umberto Anselmi-Tamburini, University of Pavia, Pavia, Italy


r 2019 Elsevier Inc. All rights reserved.

Introduction

The term Spark Plasma Sintering (SPS) is generally used to identify a sintering technique involving the contemporaneous use of
uniaxial pressure and high-intensity, low-voltage, pulsed current. In general terms, SPS can be considered a modification of hot
pressing, where the furnace is replaced by the mold containing the sample, that is heated by a current flowing directly through it
and eventually through the sample. However, since there is no general agreement on the details of the technique, an unequivocal
definition of the SPS and its associated procedures cannot be defined (Grasso et al., 2009; Orrù et al., 2009). The name SPS itself
has been often disputed as, despite several attempts, the presence of plasma and electric discharges during the process it has never
been proved unequivocally.
SPS has been receiving growing attention in the last two decades due to its remarkable effectiveness, allowing to obtain fast
sintering and densification, particularly in the case of materials considered hard to sinter, such as extremely refractory materials,
metastable phases or nanomaterials. For some of these materials, SPS has already become the sintering technique of choice.
SPS belongs to a quite extensive group of techniques involving the use of electric current for the sintering of materials. Some of
these techniques have been introduced, and found widespread application, since the beginning of the last century, particularly for
metallic materials. More details on this historical evolution can be found in recent reviews (Olevsky and Dudina, 2018; Guillon
et al., 2014; Munir et al., 2011; Orrù et al., 2009). The most distinctive characteristics of SPS, represented by the use of pulsed DC,
appeared originally in a patent by Inoue in the mid ‘60 of the last century (Inoue, 1966), while the term SPS was introduced later
on, by the Japanese producers of the first commercial machines. In the beginning, the technique found application mostly in Japan
and in a few other far-east countries. The diffusion in western countries, mainly in research institutions, started in the mid ‘90 of
the last century and spread rapidly in the industrial environment.
Machines similar to SPS, but using electric currents presenting a time pattern different from pulsed DC have also been proposed
over the years. In particular, the use of plain DC or AC has been suggested. Although there is no generally recognized nomen-
clature, such techniques are usually referred to as ‘Field Assisted Sintering’ (acronym FAST) (Orrù et al., 2009). However, a
thorough investigation on the role played by the current pattern in determining the characteristics of the sintering process has
never been presented. Available indications are suggesting that the current pattern might play a significant role only in a few
specific cases, while the main characteristics remain quite similar. For this reason, the coupled acronym SPS/FAST is often used in
recent literature to indicate all pressure-assisted, current-assisted, rapid sintering techniques.

General Characteristics

When compared to other more conventional sintering techniques, such as pressureless sintering, hot-pressing, and hot isostatic
pressing, SPS is characterize by some distinctive features, represented by the combination of faster heating rates, shorter sintering
cycles, lower sintering temperatures, and reduced grain growth. SPS systems can typically operate using heating rate in the range of
hundreds of degrees per minute, but values above 10001C min1 can in principle be achieved. These values are between one and
three orders of magnitude higher than in conventional sintering and allow for a drastic reduction in the length of the sintering
cycles. Besides faster heating rates, the technique also allows shorter sintering times. The two effects combined allow the reduction
of typical sintering cycles to a few minutes, which compares with the few hours generally used in conventional sintering. This
increase in sintering efficiency allows also reducing the sintering temperatures, sometimes hundreds of degrees, particularly in the
case of very refractory materials. This decrease in the sintering temperature produces a general reduction in grain growth, producing
materials characterized by smaller grain size and better mechanical properties.

Experimental Setup

Machine Layout
As mentioned before, the general layout of SPS systems presents several similarities with that of a hot-pressing. A simplified
schematic is reproduced in Fig. 1. The sample is placed in a mold, or die. No preliminary forming of the material to be sintered is
required, as it is generally introduced in the mold in powder form. The mold is warmed up through the passage of a high intensity/
low voltage electric current directly through it. A uniaxial pressure is applied by an hydraulic system. These pistons also act as
electrodes, supplying the high-intensity current that must flow through the die and the sample. The hydraulic pistons are water-
cooled in order to protect them from the heat deriving from the die. The pressure is controlled by a PID feedback system in order
to keep it constant during the sintering process. In fact, the sample and the die experience a considerable change in volume during

Encyclopedia of Materials: Technical Ceramics and Glasses doi:10.1016/B978-0-12-803581-8.11730-8 1


2 Spark Plasma Sintering

Fig. 1 Schematic of a typical SPS system. Reproduced from Munir, Z.A., Quach, D.V., Ohyanagi, M., 2011. Electric current activation of sintering:
A review of the pulsed electric current sintering process: Electric current activation ofsintering. J. Am. Ceram. Soc. 94, 1–19.

the process, that, if not properly compensated, can result in significant changes in the applied pressure. An example of the overall
variations in the linear dimension of the die/sample assembly during a typical SPS sintering cycle is presented in Fig. 2. The actual
sample volume contraction due to the densification cannot be directly obtained from these curves unless a calibration based on
the thermal expansion characteristics of the tool and of the fully densified samples is performed first.
The most distinctive feature of SPS systems is represented by the use of pulsed electric current. As mentioned in the intro-
duction, this current can present different characteristics. Typically low-voltage (o10 V) and high-intensity currents (between 1
and 50 kA) are used. Different time patterns can be used. In early SPS systems, the current was always pulsed DC, presenting pulses
of constant duration (about 3 ms) that could be assembled in predefined sequences (Fig. 3). More recent machines offer more
flexibility, allowing to define the duration of the pulses, the delay time between them, the total number of pulses in the sequence,
and the number of pulses on and off. In some cases, plain DC or AC can also be used.
The temperature control is generally realized using a thermocouple or a pyrometer, depending on the maximum required
temperature. The thermocouple is usually positioned in a small hole drilled halfway through the lateral wall of the die body. A
larger hole (few mm in diameter), acting as a black-body cavity, is used in the case of pyrometer reading. Alternative positioning
for the pyrometer and the thermocouples have been suggested. The most common involves positioning on the central axis of one
of the punches of the die. This axial positioning is more accurate but it is less used because more complicated to setup.

Experimental Procedure
A typical SPS process involves the following sequence of operations. The starting material, generally in the form of fine powders, is
placed in the die. The characteristics of this die must be selected on the basis of considerations that will be discussed in the
following sections. The die is then positioned between the two hydraulic pistons and the uniaxial pressure is applied. Different
kinds of pressure cycles can be used. In most cases, the pressure is applied at the beginning of the process and is maintained
constant throughout the sintering cycles. A common alternative is to apply only a minimum pressure at the beginning and the full
pressure only when the sample reachs the target sintering temperature. The system is evacuated to a low vacuum (generally few Pa).
The main purpose of the vacuum is just to protect the graphite tools from oxidation and to reduce the thermal conduction. An
atmosphere of inert gas can be used when overheating is not an issue. The use of an inert atmosphere might also help to reduce the
evaporation of volatile samples. Air, on the other hand, requires an oxidation-resistant tool material, such as SiC. The sintering cycle
is then started. The temperature profile generally involves a linear ramp to the target temperature with a heating rate between 50 and
2001C min1. More complex temperature cycles can be used in the case of reacting or outgassing materials. At the end of the sintering
process the heating is turned off and the pressure released. In most cases, the sample is let to cool naturally. This involves a very rapid
initial cooling, in the range of several hundreds of degrees per minute. The sample can be removed after 20–60 min, depending on
the type of machine and the size of the sample.
Spark Plasma Sintering 3

Fig. 2 Example of the variation in the linear dimension of the SPS tool during a typical densification experiment. The experimental curve indicates the
overall modification. The baseline indicates the changes due only to the thermal expansion of the die and the fully dense sample. The corrected line is
obtained subtracting the two previous curves and indicates the actual change in the dimension of the sample due to the densification process.
Reproduced from Anselmi-Tamburini, U., Garay, J.E., Munir, Z.A., 2005. Fundamental investigations on the spark plasma sintering/synthesis process.
Mater. Sci. Eng. A 407, 24–30.

Temperature, Current and Stress Distribution


A crucial aspect of SPS technology is represented by the possible presence of inhomogeneities in the final product deriving from an
uneven distribution in temperature, current density, and mechanical stress during the densification process. This point was largely
underestimated in the early applications of the technique, but has been clarified in subsequent investigations, thanks mostly to
numerical modeling (Muñoz and Anselmi-Tamburini, 2013; Olevsky and Froyen, 2006; Olevsky et al., 2012; Pavia et al., 2013).
The most critical aspect is represented by the presence of temperature inhomogeneities. These inhomogeneities represent the
downside of otherwise positive characteristics of SPS, represented by its fast response time and low thermal inertia. The absence of
a furnace makes the control of the temperature distribution much more challenging. In SPS systems the heating of the sample
derives only from the current flowing through the die and the sample itself. The current distribution, however, could be quite
complex, depending primarily on the die geometry and the electrical properties of the sample. The portion of the punches sticking
out of the die, for instance, usually represents the part of the tool with the smaller cross-section and the higher current density; as a
result, these parts tend to heat up faster and to remain hotter throughout the process. Within the die body, the current densities can
vary considerably, depending on the sample conductivity. In the case of samples presenting an elevated electrical conductivity
most of the current flows through the sample itself, with only a minimal amount of current flowing in the graphite walls of the die
(see Fig. 4). However, since the joule heating is dependent also on the resistivity of the material, regions with low resistivity give a
little contribution in terms of heat generation (Fig. 5). On the other hand, in the case of non-conducting samples, a much more
complex current distribution is expected, with the higher current densities being observed within the die walls and on the edges of
the sample.
Heat losses contribute also in determining the temperature distribution. Since SPS experiments are generally performed in a
vacuum, at low temperatures the only significant heat loss is due to the conduction along the die punches towards the cooler
hydraulic cylinders. At higher temperatures, above 700–8001C, radiation losses tend to dominate. To further complicate the
picture, it must be considered that the electrical properties of the die and of the sample change with temperature and the extent of
the densification process. It is not unusual for metallic samples, for instance, to present very low electrical conductivity at the
beginning of the densification process, when still in powder form, and become very conductive when the densification is
completed. All these factors contribute to producing complex and difficult to predict temperature distributions. Experimental
investigations on these aspects have been quite limited, but extensive numerical modeling has been used to clarify the influence of
4 Spark Plasma Sintering

Fig. 3 Example of a typical current pattern used in SPS machines of the first generation. The pulses width is constant (3 ms), while the number
of pulses on and off in the sequence can be modified (see boxes). Reproduced from Chen, W., Anselmi-Tamburini, U., Garay, J.E., Groza, J.R.,
Munir, Z.A., 2005. Fundamental investigations on the spark plasma sintering/synthesis process. Mater. Sci. Eng. A394, 132–138.

the various parameters on the temperature uniformity within the samples (Muñoz and Anselmi-Tamburini, 2013; Olevsky and
Froyen, 2006; Olevsky et al., 2012; Pavia et al., 2013). These studies evidenced, for instance, as the temperature distribution is
strongly dependent on the size of the sample. Samples with small diameter, up to 1–2 cm, always present quite limited tem-
perature inhomogeneities. In these samples the radial temperature distribution is generally characterized by a maximum in the
center of the sample, which can be more or less evident depending on the electrical properties of the sample. Obtaining a uniform
temperature distribution becomes challenging in the case of larger samples, presenting a diameter of 5 cm or higher. For very large
samples temperature differences in the range of tens of even hundreds of degrees between the center and the edges of the samples
have been reported. Some actions can be taken to mitigate such inhomogeneities. First and foremost, an appropriate design of the
die must be optimized through numerical modeling. In some cases, a reduction in the heat losses can be helpful, particularly if
very high temperatures are involved. It is a quite common practice, for instance, to wrap the dies with carbon or alumina felt in
order to reduce heat losses by radiation. Although this might help in the case of radial temperature distribution presenting a
maximum in the center, it might be detrimental when, as in the case of non-conducting samples, the temperature maximum tends
to be located on the sample edges.
Besides temperature, stress might also present non-uniform distribution within the sample. This point has been less investi-
gated also because it is very difficult, if not impossible, to monitor the stress distribution within the die and the sample during the
process. The available information derives only form numerical modeling. A relevant point evidenced by these studies is the role of
the difference in thermal expansion between the sample and the die. Materials with thermal expansion larger than graphite, such
as metals, experience a significant increase in the stress in the proximity of the sample edges. The role of these stress distribution on
the microstructure of the densified materials has never been thoroughly investigated, although, in principle, might be relevant. On
the other hand, differences in thermal expansion might induce large stress in the peripheral region of the sample during the
cooling, which is generally fast and uncontrolled in typical SPS processes, producing fracture in brittle materials.
Spark Plasma Sintering 5

Fig. 4 Distribution of the current flow in the case of (a) conductive powder and die, (b) conductive powder, insulating die, (c) non-conductive
powder, conductive die (note the vertical hole in the punch to enable temperature measurement by an axial pyrometer). Reproduced from Guillon, O.,
Gonzalez-Julian, J., Dargatz, B., et al., 2014. Field-assisted sintering technology/spark plasma sintering: Mechanisms, materials, and technology
developments: FAST/SPS: Mechanisms, materials, and technology developments. Adv. Eng. Mater. 16, 830–849.
6 Spark Plasma Sintering

Fig. 5 Infrared thermal images of open die containing: (a) Alumina sample. (b) Copper sample. Reproduced from Manière, C., Pavia, A., Durand, L.,
et al., 2016. Finite-element modeling of the electro-thermal contacts in the spark plasma sintering process. J. Eur. Ceram. Soc. 36, 741–748.

Tooling
The choice of the material used as a die represents a critical point of SPS technology. In order to be effective, the die must present
the appropriate combination of thermo-mechanical-electrical properties within the temperature range used for the densification
process. Regarding the mechanical stresses, it must be considered that while the punches experience only compressive stress, the
body of the die experience a complex, mostly tensile stress. The die must also present low reactivity towards the sample and
suitable electrical properties. This last requirement implies an electrical conductivity that is high enough to allow high-intensity
current to flow, but low enough to produce a significant Joule heating. Although ideal for most applications, graphite presents
several limitations. Even the best graphites, characterized by high density, low porosity, and small grain size, present a quite
limited compressive strength (o150 MPa). This limitation reduces the maximum uniaxial pressure that can be applied to the
sample. Another obvious limitation of graphite is related to the possible reaction/contamination of the sample. Several transition
metals can, in fact, form very stable carbides. Besides altering the chemical composition of the surface of the sample, this might
result in the formation of a strong bond, that in some cases can make even impossible to remove the sample from the die without
breaking it. In these cases some sort of a superficial lining is necessary. Graphite contamination can also derive from the release of
graphite powders during the sample loading or from high-temperature fast diffusivity of graphite in some materials. Finally, it
must be considered that graphite dies can be used only in vacuum or in an inert atmosphere. This environment, on the other hand,
can be damaging towards some kind of samples, as they result exposed to a very strong reducing environment during the sintering
process. Low-stability oxides represent a typical case, as they can be partially or totally reduced during SPS processing. Unfortu-
nately, alternatives to graphite are quite limited. When high temperatures are involved, the only materials presenting mechanical
and chemical properties better then graphite are silicon carbide and tungsten carbide. Silicon carbide also represents the only
alternative when oxidizing atmosphere must be used. SiC and WC dies, however, are very expensive, particularly when large
samples are involved. Furthermore, being a semiconductor, the electrical resistance of SiC decreases significantly with temperature,
complicating the temperature control. Metallic dies, such as high-strength, low-conductivity steel, can be used only for low-temperature
Spark Plasma Sintering 7

applications, where plastic deformation and creep are not a concern. The same considerations hold for cemented carbides (hard metals)
presenting much higher compressive strength, but only at temperatures below 7001C. A combination of ceramic dies with conducting
punches is sometimes used, although this requires the sample to be conductive.

Microscopic Mechanisms

The remarkable ability of SPS to produce fast densification generated, from the beginning, an intense debate about the underlying
microscopic mechanisms. Several excellent reviews are available in the literature discussing this topic in detail (Anselmi-Tamburini
et al., 2013; Cao et al., 2019; Olevsky and Dudina, 2018; Garay, 2010; Grasso et al., 2009; Guillon et al., 2014; Munir et al., 2011,
2006). It must be noted, however, that the number of proposed microscopic mechanisms is quite large and single out the
importance of each one of them proved to be very difficult. The number of available experimental and modeling approaches that
can be used to test their relevance is, in fact, limited. In general, it can be concluded that there is probably not a single process that
is responsible for all the SPS characteristics, but a number of them, that vary depending on the thermal, electrical and chemical
characteristics of the material to be sintered. It also often forgotten that the growth of interest towards SPS coincided with the
increased availability of materials in the form of nanopowders. Due to the fast heating rates, SPS can take better advantage than
conventional sintering of the intrinsic faster sinterability of the nanopowders. However, the understanding of the basic
mechanisms is still largely a work in progress, which is crucial for bringing this technique on a solid basis and it is expected to
continue receiving large attention in future investigations.

Arc and Plasma Discharges


In early works it was suggested that the unique SPS characteristics derived entirely from the type of pulsed current used during the
process (Tokita, 1999). Short pulses of high-intensity current were, in fact, believed to generate local conditions enhancing the atomic
mobility through the formation of plasma, or electric discharges, in the small volumes between the grains. Although it was believed
that these phenomena were active mostly in the early stages of the process, their beneficial effects were supposed to influence also the
late stages of the sintering process. This interpretation enjoyed a large fortune and is still often presented today when discussing the
SPS characteristics. Later investigations, however, questioned this interpretation evidencing some major inconsistencies it implies
(Hulbert et al., 2008a). Most types of vacuum discharges, such as arc, spark and glow discharges, require high voltages, that are not
compatible with the SPS experimental conditions, generally characterized by voltages below 10 V. Only arc discharges can in
principle be sustained under low voltages, but they require high electrical conductivity, while SPS is very effective even in the case of
insulating ceramic materials. More importantly, the presence of plasma or electrical discharges was never substantiated by any sound
experimental evidence, with few exceptions limited to some specific situation (Bonifacio et al., 2013; Saunders et al., 2015; Zhang
et al., 2014). So, it became more and more difficult to justify the characteristics of the SPS considering only a single mechanism.

Applied Electric Fields


The possible influence of the applied electric fields plays a central role in the discussion on the mechanisms responsible for the SPS
characteristics. This interest is suggested by the consideration that SPS processes involve the use of very high current intensities, up
to several kA. In reality, it must be considered that the intensity of the electric field and the currents experienced by the sample are
usually quite modest. As evidenced by Fig. 1, the graphite die and the sample can be seen as two electric loads in parallel to each
other. Since the electrical resistance of the die is quite low, the voltage drop across the die and the sample is always limited to a few
mV, regardless of the voltage generated by the power supplier (usually below 10 V) and of the sample resistivity. For the same
reason, the current flow is usually mostly confined within the die components, except in the case of very conducting samples, such
as metals in the late stages of densification. These considerations drive to the conclusion that the influence of the electric fields and
currents on the densification mechanisms involved in SPS processes, has probably been often overstated.
However, applied electric fields can indeed modify considerably the atomic mobility in solid materials. The most direct
evidence is observed in ionic materials, where applied electric fields can induce the movement of the constitutive ions (electro-
migration), or produce polarization when the electric charges are not free to move. Both phenomena have been largely investi-
gated in solid state science and have the potential to produce profound modifications in the sintering mechanisms. Despite that,
the possibility to enhance or decrease the rate of sintering and densification of ionic materials through the application of an
electric field is still largely unexplored. However, it must be considered that the influence of electrotransport can be quite complex
and depends strongly on the characteristics of the material under consideration. Not in all cases, in fact, the ionic flux produced by
the applied electric field can produce the kind of material transport required in sintering. Complete decomposition and recon-
struction of crystallographic sites are required, not just the movement of one component. However, in ionic materials intense
fluxes of single charged elements might induce mobility of other elements through correlation effects. A typical example is
represented by the correlation between electrons and ions movements in mixed conductors. In the case of dielectric materials,
polarization might play a significant role, particularly when very small grains are considered. In the point of contact between small
particles the polarization effect might produce an amplification of the applied electric field that might extend to several orders of
magnitude, producing extreme localized conditions that might enhance atomic mobility.
8 Spark Plasma Sintering

Another phenomenon associated with applied electric fields often mentioned in SPS literature is electromigration. This term
generally relates to the transport of material in metals when elevated electronic current densities (41000 A/cm2) are present.
Electromigration has been extensively investigated for its implication in microelectronics but has never been well characterized in
materials presenting complex microstructures, such as powder compacts in the early stages of sintering. Its association with SPS is
suggested by the intense electric currents involved with this technique although, as mentioned earlier, the local current densities
experienced by SPS samples are usually orders of magnitude below the threshold required for the electromigration phenomena to
become evident. Direct experimental evidence on the role played by electromigration in SPS sintering, is limited to model systems
where all the current is forced to flow through the sample (Frei et al., 2007).
Other physical effects that are often associated with the presence of applied electric fields in SPS are elastoplasticity, ponderomotive
forces, electromagnetic ‘pinch’ effect, and Peltier effect. The relevance of these effects on the actual sinter process has still largely to be
determined.

Applied Pressure
SPS belongs to the group of the pressure-assisted sintering techniques, as hot-pressing and hot-isostatic pressing. Similar arguments
relative to the influence of the pressure on densification apply to all these techniques. The pressure is generally expected to enhance
densification through plastic deformation and grain boundary sliding. These processes reduce the pore volume and the dependence
of the densification from bulk diffusion. As a result, also grain growth is generally reduced applying pressure. As noted before, in SPS
the maximum applied pressure is limited by the compressive strength of the tooling. Graphite dies do not allow pressures above
100–150 MPa. High-pressure modification of the conventional technique has been presented and be described later on.

Heating Rates
Another typical feature often associated with SPS superior sintering capability is represented by the possibility to use very high
heating rates. As mentioned before, in the SPS practice it is quite common to use heating rates in the range of hundreds of degrees
per minute. These values are between one and three orders of magnitude higher than in hot-pressing, hot-isostatic pressing, or
even in conventional pressureless sintering. Such a relevant increase in the heating rate has a profound influence on the sintering
process. Although in general the heating rate does not modify the densification mechanisms, it can influence the relative
importance of competing processes presenting different activation energy. During sintering, in fact, several mass transport pro-
cesses are active at the same time. We can mention, in order of increasing values of activation energy, surface diffusion, grain
boundary diffusion, power-law creep, and bulk diffusion. Surface diffusion is a non-densifying process, as it cannot contribute to
the growth of the intergranular necks. It contributes significantly, on the other hand, on grain growth. As a result, long exposure of
the green compact at temperatures where surface diffusion is dominating (0.25–0.4 Tm) might result in a microstructure char-
acterized by poorly sintered large grains and large pores, reducing the effectiveness of the following high-temperature densification
process, as the driving force for bulk diffusion is reduced. In this respect, high heating rates, allowing to reduce or skip the lower
temperature processes entirely, increase the effectiveness of the densifying processes. This argument is particularly relevant when
the starting powder is nanometric. Nanopowders, in fact, present an intrinsic higher sinterability, that might be lost in the presence
of long, low-temperature annealing.
It has also been suggested that a higher heating rate might enhance densification through thermomigration or thermotransport
effect, driven by the presence of macroscopic and microscopic temperature gradients generated when high heating rates are
involved. The details of these phenomena are quite complex and strongly dependent on the thermo-electrical characteristics of the
materials. We already discussed the macroscopic temperature gradients before. For conducting materials, the possibility of
microscopic temperature gradients at the point of contact between the grains has often been suggested in the literature. A
quantitative analysis of such a phenomenon, however, shows that it might become relevant only in the case of very large grains of
materials presenting low thermal conductivity. When nano or submicrometric powders are involved, the build-up of temperature
gradients within each grain is impossible, as the thermal conduction is too fast, even in the presence of large currents flowing
between the grains.

Examples of Application

In the last three decades, SPS enjoyed a growing success as a method for the densification of functional and structural materials.
There are countless examples of applications including almost any class of material. A comprehensive list of these applications can
be found in the available reviews and in the extensive original literature (Grasso et al., 2009; Guillon et al., 2014; Orrù et al., 2009).
Most of these applications involve materials that can also be obtained using more conventional sintering techniques, but that
benefit from the reduction in sintering times, in sintering temperatures, and grain growth offered by SPS. Materials that are
traditionally considered harder to sinter received more attention. Among them, there are surely refractory metals, such as W, Ta,
and Re, that present outstanding mechanical and thermal properties, but are difficult to densify completely due to their high
melting point. The use of SPS allows, in this case, to reduce the maximum required temperature and sintering time, with a
consequent reduction in contamination, from oxygen and graphite, and a reduction in the grain size. Similar results have been
Spark Plasma Sintering 9

obtained in the case of intermetallic phases for advanced high-temperature structural applications. Typical examples are repre-
sented by the intermetallic phases present in the Nb/Al and Ti/Al systems. In this case, reactive sintering has often been applied
using SPS. In this approach, mixtures of the basic metals have been treated, performing synthesis and densification in the same
process. SPS is generally quite effective in enhancing solid-state reactions kinetics, as the applied uniaxial pressure helps main-
taining the ideal contact between the reacting phases. Smaller grain size has been obtained using this approach, with beneficial
effects on the mechanical properties. Another typical area of application is represented by ultra-high-temperature ceramics, such as
borides, carbides, and nitrides of Zr, Ti, and Hf. These materials, with their extreme thermal characteristics, have always represented
a challenge for material processing. The interest towards them, on the other hand, has been growing significantly in the last two
decades. Long sintering cycles using HP or HIP have always been the only viable approach for their densification. However, this
approach produced significant grain growth and contamination, reducing the possibility of obtaining full densification. The use of
SPS allowed reducing the processing time drastically, with an overall beneficial effect on the material properties and a significant
reduction in the production costs. Another field where SPS attracted large interest is represented by transparent ceramics for optical
applications. The possibility of obtaining transparent polycrystalline ceramics, characterized by lower costs and better mechanical
properties than single crystals, has driven this interest. Many applications can be found in the literature, with particular emphasis
on yttrium-aluminum-garnet (YAG), spinels, zirconia, and alumina. However, contamination from graphite and formation of
crystallographic defects, mostly oxygen vacancies, deriving from the strongly reducing conditions typical of SPS, have been found
to reduce the in-line transmission in most cases.
Besides these more conventional applications, SPS opened the possibility to realize materials that are problematic or even
impossible to obtain using conventional sintering approaches. In the following sections we present a brief review of these
applications.

Nanostructured Bulk Materials


The nanostructure has a profound influence on the physical and chemical properties of materials. For this reason, the possibility of
obtaining bulk materials preserving the unique properties deriving from the nanostructure is considered particularly attractive.
Bulk materials presenting a grain size between 10 and 50 nm present, in fact, superior mechanical properties and, in some cases,
unique functional characteristics. The nanometric grain size can also alter the phase equilibria in polymorphic materials, allowing
to stabilize materials in unusual crystallographic forms. Sintering and grain growth, however, share the same mechanisms and
obtaining full densification without the loss of the nanostructure is a challenging feat. Conventional sintering approaches have
largely failed in this respect. The introduction of SPS offered a valuable alternative, that proved to be particularly effective. Because
of this, SPS has become the technique of choice for the production of bulk nanostructured materials. Fig. 6 shows how the use of
short sintering times (5 min) and high uniaxial pressure allow reducing drastically the grain growth in the case of SPS sintering of
nano-YSZ.

Functionally Graded Materials


FGM represents a class of materials characterized by a gradient in composition or microstructure along one direction. Although
these materials have been introduced well before the SPS technology became available, the use of SPS opened new possibilities for
their development, particularly regarding the materials presenting a gradient of microstructure. In early acceptation, FGMs were

Fig. 6 Relationship between the sintering temperature, applied pressure and final grain size in the case of nanometric fully-stabilized zirconia
(8% YO1.5). Reproduced from Anselmi-Tamburini, U., Garay, J.E., Munir, Z.A., 2006. Fast low-temperature consolidation of bulk nanometric
ceramic materials. Scr. Mater. 54, 823–828.
10 Spark Plasma Sintering

characterized only by the presence of compositional gradients. Typical examples were represented by materials presenting a
gradual transition from metallic to ceramics (see Fig. 7). Even in the case of these conventional FGM, the SPS approach offers some
advantages, as the shorter sintering times allow reducing the possibility of interdiffusion and chemical reaction between the
individual components during the process. Obtaining FGM presenting a gradient in porosity or in grain size is much more difficult
using conventional sintering methods. The realization of these materials, in fact, requires the presence of steep temperature
gradients. The localized heat generation mechanism, typical of SPS, is in principle well suited for obtaining these conditions. It has
been shown that significant temperature gradients can be obtained quite simply just modifying the assembly of the die or
changing its shape, as shown in Fig. 8. A simple offset of the die body, leaving only one of the punches outside can produce strong
axial temperature gradients, in the range of hundreds of degrees per cm. More complex temperature gradients can be obtained
modifying the shape of the die. An example of an innovative material that can be obtained using this approach is shown in Fig. 9.
Here a sample of B4C presents an axial gradient of porosity that was then backfilled with molten aluminum under vacuum. This
allowed realizing a material that is extremely hard and wear-resistant on one end and presenting a high toughness on the other.
Radial instead that an axial gradient of porosity and composition can also be obtained. The proper control of the temperature
gradients requires careful design of the die, which must be guided by appropriate numerical modeling.

Fig. 7 The microstructure of an eight-layer FGM produced by SPS. Reproduced from Fujii, T., Tohgo, K., Isono, H., Shimamura, Y., 2017.
Fabrication of a PSZ-Ti functionally graded material by spark plasma sintering and its fracture toughness. Mater. Sci. Eng. A 682, 656–663.

Fig. 8 Tool configurations generating different temperature distributions: (a) symmetric configuration, (b) asymmetric location of the die body,
and (c) asymmetric graphite die. Reproduced from Guillon, O., Gonzalez-Julian, J., Dargatz, B., et al., 2014. Field-assisted sintering technology/
spark plasma sintering: Mechanisms, materials, and technology developments: FAST/SPS: Mechanisms, materials, and technology developments.
Adv. Eng. Mater. 16, 830–849.
Spark Plasma Sintering 11

Fig. 9 The microstructure of a B4C–Al FGM along with an EDS profile of the constituent elements relative to the cross-section distance. The
darker region is B4C rich, while the lighter areas contain Al. Al infiltration occurred from left to right. Reproduced from Hulbert, D.M., Jiang, D.,
Anselmi-Tamburini, U., Unuvar, C., Mukherjee, A.K., 2008b. Experiments and modeling of spark plasma sintered, functionally graded boron
carbidealuminumcomposites. Mater. Sci. Eng. Struct. Mater. Prop. Microstruct. Process A488, 333–338.

Non-Equilibrium Materials
The short sintering times involved in the SPS technology offer the unique possibility to realize the densification of materials that are
out of equilibrium. This is generally impossible to obtain with conventional sintering, as the long high-temperature treatments
involved with these techniques drive necessarily the materials towards their most stable state. Over the last few years, a large number of
densifications of non-equilibrium materials using SPS have been reported. A typical example, already mentioned before, is represented
by nanometric materials. The intrinsic instability of most nanophases towards high-temperature treatment makes them impossible to
densify using conventional sintering methods. Other non-equilibrium materials are represented by amorphous phases or glasses. A
very interesting example of application in this area is represented by the densification of metallic glasses (Fig. 10). These materials are
generally obtained by rapid solidification, an approach that limits strongly the possibility to produce objects with complex shapes and
in bulk form. The possibility to densify powders deriving from gas atomization or by mechanical alloying opens new and interesting
possibilities of application. Numerous examples have been reported (Choi et al., 2007; Chu et al., 2012). Another interesting
application is represented by the possibility to produce bulk materials presenting an unstable or metastable crystallographic structure.
Several examples have been reported, particularly in the case of metastable crystallographic forms stabilized by the nanostructure, as
for undoped cubic/tetragonal zirconia and titania in anatase crystallographic form (Maglia et al., 2010; Masahashi, 2007). The
densification of supersaturated solid solutions represents another field of particular interest. The densification of Al alloys with a
content of other metals above their saturation limit has been reported (Zhang et al., 2013). Particularly interesting is the possibility to
maintain local compositional gradients, in the micro or nanometric range, as it opens the possibility to realize materials with unique
functional or transport properties not accessible to homogeneous materials. A typical example is represented by the densification of
core-shell powders (Airimioaei et al., 2017), avoiding the homogenization by interdiffusion. Recently the possibility to obtain doping
localized at the grain boundary, taking advantage of the low-temperature surface diffusion has also been demonstrated. Other
examples of technologically relevant metastable material are represented by the formation of non-equilibrium composites containing
diamond or cubic boron nitride (Herrmann et al., 2012).

Innovative Aspects and Future Evolutions

SPS technology has grown in the last two decades to become one of the most relevant innovations in sintering. However, the area of
field-assisted sintering is still evolving, and in the last few years some new trends emerged, that might challenge the more ‘conventional’
applications. Some of these evolutions try to answer unresolved limitations of the technique; others are just attempting to expand its
application limits. In the following sections, some of the most interesting new evolutions will be briefly described.

High and Ultra-High Pressures


As already noted before, the maximum uniaxial pressure achievable with a standard SPS tooling, made out of high-density
graphite, falls in the range between 100 and 150 MPa. In the last few years, however, attempts to extend such limit has been
12 Spark Plasma Sintering

Fig. 10 SEM images of the specimens of Cu46Zr42Al7Y5 metallic glass sintered at 593K (a) and 653K (b); EDS spectra taken from the specimen
sintered at 653K (C); high-resolution TEM micrograph taken from the specimen sintered at 653K (d); Bright-field TEM image of the specimen
sintered at 653K (e) and the corresponding SAD patterns (f)–(h) taken from the indicated by “A”, “B” and “C” in (e), respectively. Reproduced
from Chu, Z.H., Kato, H., Xie, G.Q., et al., 2012. Consolidation and mechanical properties of Cu46Zr42Al7Y5 metallic glass by spark plasma
sintering. J. Non-Cryst. Solids 358, 1263–1267.

reported several times. The possibility to use much higher pressure can be particularly beneficial for applications involving the
densification of nanocrystalline materials. As reported in Fig. 6, the temperature required to obtain a certain level of densification
decrease drastically with the increase of the uniaxial pressure. As a result, high-pressure densification is particularly beneficial when
the grain growth must be controlled, as in the case of the densification of nanopowders. High or ultra-high pressure may also
induce phase transformation in some solid materials.
To achieve uniaxial pressures in the range of the hundreds of MPa or even of few GPa, the typical SPS tooling must be
properly modified. An example of a high-pressure setup is reported in Fig. 11 (Anselmi-Tamburini et al., 2006). The external
part of the die is very similar to a standard SPS tool and is made out of high-density graphite. This part is where most of the
current flows and the heating takes place. The internal part of the die is where the sample is placed and the high pressure is
realized. This part must be made with materials with the appropriate compressive strength, such as silicon carbide, tungsten
carbide, or cemented carbide. This setup allows achieving 3–4 GPa on samples that are few mm in diameter. Although this
pressure is still at least an order of magnitude below the limit achievable with diamond anvil systems, it is enough to observe
some pressure-induced phase transformations. The highest temperature achievable is limited to 700–8001C when cemented
carbides are used, but can be much higher in the case of silicon carbide or tungsten carbide tools.
Spark Plasma Sintering 13

Fig. 11 Two stages SPS die for high-pressure applications. Reproduced from Anselmi-Tamburini, U., Garay, J.E., Munir, Z.A., 2006. Fast low-temperature
consolidation of bulk nanometric ceramic materials. Scr. Mater. 54, 823–828.

Fig. 12 Die design presenting a reduction of the cross-section allowing a reduction in the electric power required to reach the target temperature
and improved temperature uniformity. (a) View of the sample/die assembly; (b) details of the punches presenting holes (left) and rings (right).
Reproduced from (a) Giuntini, D., Raethel, J., Herrmann, M., Michaelis, A., Olevsky, E.A., 2015. Advancement of tooling for spark plasma sintering.
J. Am. Ceram. Soc. 98, 3529–3537. (b) Giuntini, D., Raethel, J., Herrmann, M., et al., 2016. Spark plasma sintering novel tooling design:
Temperature uniformization during consolidation of silicon nitride powder. J. Ceram. Soc. Jpn. 124, 403–414.
14 Spark Plasma Sintering

Tooling Optimization
Despite a large number of successful applications, SPS is often considered a technique with limited applicability to large
scale industrial processes. This conviction derives in part from few misconceptions developed in early applications, but is
also based on some real issues that are limiting its scaling up. In this respect SPS presents two main limitations: there is a
limit to the maximum size of the samples that can be produced and it allows to obtain only one sample at a time. Despite
their relevance, little effort has been paid in trying to mitigate these two limitations. Both limitations are ultimately related
to the tooling design. The maximum sample dimension is mostly controlled by two factors: (1) the maximum current that
can be delivered by the power supplier used in SPS machines, and (2) the difficulty of maintaining a uniform temperature
distribution in very large samples. Regarding the first point, it must be noted that current intensities above 40–50 kA are
impractical for economic and logistic reasons. This threshold limits the maximum achievable sample dimension to disks of
30–40 cm in diameter, as the graphite tools used in SPS require very high current to produce enough joule heating to warm
up large samples. In this respect, hot pressing and HIP can produce even larger samples with lower investment and
operating costs. However, the geometry of large dies is never usually optimized, as it is generally obtained by a simple
scaling up of dies used for the smaller samples. Recently it has been shown that a proper design of the tools can save
considerable electric power through an optimization of the current fluxes (Giuntini et al., 2016, 2015). A careful design of
the tools, achieved through numerical modeling, also allows optimizing the temperature distribution in the case of large
samples. An example of this approach is shown in Fig. 12, where the electrical cross-section of the punches has been
properly reduced in order to obtain a substantial reduction in the electric power required to reach high temperatures and
uniform temperature distribution within the sample.
Analogous considerations can be applied to the other limitation of SPS, represented by the possibility to produce only one
sample for each sintering cycle. Also in this case, the limitation is mostly related to a lack of evolution in the tooling design. In
fact, it is possible, at least in principle, to produce several samples with a single sintering cycle using an appropriate tool. In
Fig. 13 few examples of possible setups are reported, presenting parallel, serial, or parallel-serial alignment of tool cavities. This
approach, however, has received very limited attention in the SPS practice. Such complex tools, in fact, are difficult to make and
to control, particularly in terms of temperature uniformity, as the current fluxes are in this case even more complex and difficult
to predict without an accurate numerical modeling (Fig. 14).
Finally, it is worth mentioning the possibility of realizing objects characterized by complex shapes. Traditionally, SPS samples
present a disk-like or cylindrical shape. Recently, however, it has been presented the possibility of realizing more complex shapes
through near-net-shape approaches involving the use of the so-called “deformable interfaces” or “controllable interfaces” method
(Manière et al., 2019). In this approach, a layer of deformable material, usually a graphite foil, is shaped to define a complex
geometry. The foil is then placed within a traditional SPS tool, and the two volumes separated by the foil are filled with similar or
different powders. At the end of the SPS densification, the two regions can be easily separated, producing one or two complex
shape elements.

Fig. 13 Tools configuration for the production of multiple parts in a single sintering cycle. Reproduced from Guillon, O., Gonzalez-Julian, J.,
Dargatz, B., et al., 2014. Field-assisted sintering technology/spark plasma sintering: Mechanisms, materials, and technology developments:
FAST/SPS: Mechanisms, materials, and technology developments. Adv. Eng. Mater. 16, 830–849.
Spark Plasma Sintering 15

Fig. 14 Main steps of the spark plasma sintering (SPS) process involving the ‘controllable interface’ approach. (a) loading; (b) assembly after
SPS sintering; (c) releasing; and (d) final components after polishing. Reproduced from Manière, C., Torresani, E., Olevsky, E., 2019. Simultaneous
spark plasma sintering of multiple complex shapes. Materials 12, 557.

Fig. 15 Schematic of a hybrid SPS system including an induction heater. Reproduced from Yushin, D.I., Smirnov, A.V., Pinargote, N.W.S.,
Peretyagin, P.Y., Millan, R.T.S., 2015. Modeling process of spark plasma sintering of powder materials by finite elementmethod. Mater. Sci. Forum
834, 41–50.

Hybrid Heating
As mentioned before, one of the main challenges in SPS technology is represented by the possibility to control the temperature
distribution during the sintering process; a problem that is particularly relevant in the case of very large samples. We have seen as a
way to mitigate this problem lays in the optimization of the tool design through numerical modeling. Such an approach, however,
is time-consuming and must be repeated every time a change in the sample size or geometry is required. A more general-purpose
approach involves the use of a secondary heating source, allowing a more uniform temperature distribution. The use of a
secondary heating element also allows mitigating the other factor affecting the scaling up of SPS, represented by the maximum
power allowed by SPS power supplier. Machines allowing to use two heating methods at the same time are usually identified by
16 Spark Plasma Sintering

the term ‘hybrid SPS’ and have become commercially available in the last few years. Two different methods of secondary heating
are generally used: conventional resistance furnace and induction heating. The first approach is less expensive, but it goes
somehow in conflict with the most typical characteristic of SPS, represented by the possibility to use fast heating rates. In this
respect induction heating is more congruent and received more attention. An example of the typical setup used in a hybrid SPS
with induction heating is shown in Fig. 15. Despite their advantages, the use of a hybrid SPS is still very limited, in part due to the
high cost of the apparatus, but appear very promising when scaling up towards very large sample dimensions is crucial for the
application.

References

Airimioaei, M., Buscaglia, M.T., Tredici, I., et al., 2017. SrTiO3–BaTiO3 nanocomposites with temperature independent permittivity and linear tunability fabricated using field-
assisted sintering from chemically synthesized powders. J. Mater. Chem. C 5, 9028–9036.
Anselmi-Tamburini, U., Spinolo, G., Maglia, F., et al., 2013. Field assisted sintering mechanisms. In Sintering Mechanisms of Convention Nanodensification and Field Assisted
Processes. Engineering Materials. Springer. pp. 159–191.
Anselmi-Tamburini, U., Garay, J.E., Munir, Z.A., 2006. Fast low-temperature consolidation of bulk nanometric ceramic materials. Scr. Mater. 54, 823–828.
Bonifacio, C.S., Holland, T.B., van Benthem, K., 2013. Evidence of surface cleaning during electric field assisted sintering. Scr. Mater. 69, 769–772.
Cao, G., Estournes, C., Garay, J., Orrù, R., 2019. Spark Plasma Sintering: Current Status, New Developments and Challenges. Elsevier.
Choi, P.P., Kim, J.S., Nguyen, O.T.H., et al., 2007. Al-La-Ni-Fe bulk metallic glasses produced by mechanical alloying and spark-plasma sintering. Mater. Sci. Eng. A
449–451, 1119–1122.
Chu, Z.H., Kato, H., Xie, G.Q., et al., 2012. Consolidation and mechanical properties of Cu46Zr42Al7Y5 metallic glass by spark plasma sintering. J. Non-Cryst. Solids 358,
1263–1267.
Olevsky, E.A., Dudina, D.V., 2018. Field-Assisted Sintering: Science and Applications. New York, NY: Springer Science þ Business Media.
Frei, J.M., Anselmi-Tamburini, U., Munir, Z.A., 2007. Current effects on neck growth in the sintering of copper spheres to copper plates by the pulsed electric current method.
J. Appl. Phys. 101, 114914/1–114914/8.
Garay, J.E., 2010. Current-activated, pressure-assisted densification of materials. Annu. Rev. Mater. Res. 40, 445–468.
Giuntini, D., Raethel, J., Herrmann, M., et al., 2016. Spark plasma sintering novel tooling design: Temperature uniformization during consolidation of silicon nitride powder. J.
Ceram. Soc. Jpn. 124, 403–414.
Giuntini, D., Raethel, J., Herrmann, M., Michaelis, A., Olevsky, E.A., 2015. Advancement of tooling for spark plasma sintering. J. Am. Ceram. Soc. 98, 3529–3537.
Grasso, S., Sakka, Y., Maizza, G., 2009. Electric current activated/assisted sintering (ECAS): A review of patents 1906–2008. Sci. Technol. Adv. Mater. 10, 053001.
Guillon, O., Gonzalez-Julian, J., Dargatz, B., et al., 2014. Field-assisted sintering technology/spark plasma sintering: Mechanisms, materials, and technology developments:
FAST/SPS: Mechanisms, materials, and technology developments. Adv. Eng. Mater. 16, 830–849.
Herrmann, M., Matthey, B., Höhn, S., et al., 2012. Diamond-ceramics composites – New materials for a wide range of challenging applications. J. Eur. Ceram. Soc. 32,
1915–1923.
Hulbert, D.M., Anders, A., Dudina, D.V., et al., 2008a. The absence of plasma in “spark plasma sintering”. J. Appl. Phys. 104, 033305/1–033305/7.
Inoue, K., 1966. Apparatus for Electrically Sintering Discrete Bodies, 3250892.
Maglia, F., Dapiaggi, M., Tredici, I., Maroni, B., Anselmi-Tamburini, U., 2010. Synthesis of fully dense nanostabilized undoped tetragonal zirconia. J. Am. Ceram. Soc. 93,
2092–2097.
Manière, C., Torresani, E., Olevsky, E., 2019. Simultaneous spark plasma sintering of multiple complex shapes. Materials 12, 557.
Masahashi, N., 2007. Fabrication of bulk anatase TiO2 by the spark plasma sintering method. Mater. Sci. Eng. A 452–453, 721–726.
Munir, Z.A., Anselmi-Tamburini, U., Ohyanagi, M., 2006. 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, 763–777.
Munir, Z.A., Quach, D.V., Ohyanagi, M., 2011. Electric current activation of sintering: A review of the pulsed electric current sintering process: Electric current activation of
sintering. J. Am. Ceram. Soc. 94, 1–19.
Muñoz, S., Anselmi-Tamburini, U., 2013. Parametric investigation of temperature distribution in field activated sintering apparatus. Int. J. Adv. Manuf. Technol. 65,
127–140.
Olevsky, E., Froyen, L., 2006. Constitutive modeling of spark-plasma sintering of conductive materials. Scr. Mater. 55, 1175–1178.
Olevsky, E.A., Garcia-Cardona, C., Bradbury, W.L., et al., 2012. Fundamental aspects of spark plasma sintering: II. Finite element analysis of scalability. J. Am. Ceram. Soc. 95,
2414–2422.
Orrù, R., Licheri, R., Locci, A.M., Cincotti, A., Cao, G., 2009. Consolidation/synthesis of materials by electric current activated/assisted sintering. Mater. Sci. Eng. R Rep. 63,
127–287.
Pavia, A., Durand, L., Ajustron, F., et al., 2013. Electro-thermal measurements and finite element method simulations of a spark plasma sintering device. J. Mater. Process.
Technol. 213, 1327–1336.
Saunders, T., Grasso, S., Reece, M.J., 2015. Plasma formation during electric discharge (50 V) through conductive powder compacts. J. Eur. Ceram. Soc. 35,
871–877.
Tokita, M., 1999. Development of large-size ceramic/metal bulk FGM fabricated by spark plasma sintering. Mater. Sci. Forum 308–311, 83–88.
Zhang, Z.-H., Liu, Z.-F., Lu, J.-F., et al., 2014. The sintering mechanism in spark plasma sintering – Proof of the occurrence of spark discharge. Scr. Mater. 81, 56–59.
Zhang, F., Reich, M., Kessler, O., Burkel, E., 2013. The potential of rapid cooling spark plasma sintering for metallic materials. Mater. Today 16, 192–197.

Further Reading

Anselmi-Tamburini, U., Garay, J.E., Munir, Z.A., 2005. Fundamental investigations on the spark plasma sintering/synthesis process. Mater. Sci. Eng. A 407, 24–30.
Anselmi-Tamburini, U., Groza, J.R., 2017. Critical assessment: Electrical field/current application – A revolution in materials processing/sintering? Mater. Sci. Technol. 33,
1855–1862.
Cao, G., Estournes, C., Garay, J., Orrù, R., 2019. Spark Plasma Sintering: Current Status, New Developments and Challenges. Elsevier.
Chen, W., Anselmi-Tamburini, U., Garay, J.E., Groza, J.R., Munir, Z.A., 2005. Fundamental investigations on the spark plasma sintering/synthesis process. Mater. Sci. Eng. A
394, 132–138.
Spark Plasma Sintering 17

Fujii, T., Tohgo, K., Isono, H., Shimamura, Y., 2017. Fabrication of a PSZ-Ti functionally graded material by spark plasma sintering and its fracture toughness. Mater. Sci. Eng.
A 682, 656–663.
Garay, J.E., 2010. Current-activated, pressure-assisted densification of materials. Annu. Rev. Mater. Res. 40, 445–468.
Hulbert, D.M., Jiang, D., Anselmi-Tamburini, U., Unuvar, C., Mukherjee, A.K., 2008b. Experiments and modeling of spark plasma sintered, functionally graded boron carbide-
aluminum composites. Mater. Sci. Eng. Struct. Mater. Prop. Microstruct. Process A488, 333–338.
Manière, C., Pavia, A., Durand, L., et al., 2016. Finite-element modeling of the electro-thermal contacts in the spark plasma sintering process. J. Eur. Ceram. Soc. 36,
741–748.
Munir, Z.A., Quach, D.V., Ohyanagi, M., 2011. Electric current activation of sintering: A review of the pulsed electric current sintering process: Electric current activation of
sintering. J. Am. Ceram. Soc. 94, 1–19.
Olevsky, E.D., Dudina, D.V., 2018. Field-Assisted Sintering: Science and Applications. New York, NY: Springer Science þ Business Media.
Orrù, R., Licheri, R., Locci, A.M., Cincotti, A., Cao, G., 2009. Consolidation/synthesis of materials by electric current activated/assisted sintering. Mater. Sci. Eng. R: Rep. 63,
127–287.
Yushin, D.I., Smirnov, A.V., Pinargote, N.W.S., Peretyagin, P.Y., Millan, R.T.S., 2015. Modeling process of spark plasma sintering of powder materials by finite element
method. Mater. Sci. Forum 834, 41–50.

You might also like