You are on page 1of 27

Densification and grain growth during

sintering of nanosized particles


Z. Z. Fang and H. Wang
The sintering of nanosized particles is a scientific and technological topic that affects the
manufacture of bulk nanocrystalline materials and the understanding of the stability of
nanoparticles. Owing to their extremely small size and the high surface to volume ratio,
nanoparticles during sintering exhibit a number of distinctively unique phenomena compared to
the sintering of coarse powders. Particularly, it is generally found that the sintering temperatures
of nanosized particles are dramatically lower than that of their micrometre or submicrometre sized
counterparts. Research has also shown that the grain growth during nanosintering consists of an
initial dynamic grain growth stage that occurs during heating up and the normal grain growth
stage that occurs mostly during isothermal holding. For nanoparticles, the effect of the initial grain
growth cannot be ignored because that it is sufficient to cause the material to lose nanocrystalline
characteristics. This review aims to bring to focus the understanding of the fundamental issues of
nanosinteirng, including the thermodynamic driving force of nanosintering, non-linear diffusion
and the kinetics of nanosintering, and the relationships between agglomeration, densification and
grain growth. This review will also examine the effects of microstructure and processing variables.
Keywords: Nanoparticle sintering, Sintering, Densification, Grain growth, Coarsening, Size effect

Introduction which in turn results in different kinetic rates of


sintering.
Since the emergence of nanoscaled science and technol- In the context of engineering processes, sintering
ogy, the sintering of nanosized particles has been a topic implies the bonding of one solid particle to another. The
of both scientific and technological importance. sintering process can be regarded as consisting of two
Sintering is a phenomenon that occurs in a broad range intertwined processes: densification and grain growth.
of nanomaterials processes, including the synthesis of Another unique issue of nanosintering is that nanopar-
nanoparticles and fabrication of bulk nanocrystalline ticles almost always experience extremely rapid grain
materials. Moreover, sintering is an important factor in growth during sintering, rendering the loss of nanocrys-
determining the stability of nanomaterials, nanoscaled talline characteristics at the fully sintered state. With
coatings and nanodevices. In some cases, sintering of respect to manufacturing bulk nanocrystalline materials
nanosized particle is not desired; in others, sintering is from nanoscaled particles, the objective of nanosintering
required. For example, sintering should be prevented for is to achieve maximum densification while retaining
nanosized catalysts, while nanosized powders must be nanoscaled grain sizes. This goal, however, has been
sintered to manufacture bulk nanocrystalline materials. very difficult to reach. In fact, it is a glaring technolo-
Owing to the importance of the sintering of nanopar- gical challenge for many materials. Fundamentally,
ticles, a large body of literature on nanosintering has there are two main reasons for this technological
been accumulated and published over the past two impasse. One is that the same factors that result in
decades. Although sintering nanoparticles shares the densification also cause grain growth. In other words,
same basic principles as that of sintering coarser both densification and grain growth processes share the
particles, a number of issues and challenges are unique same driving force and mass transport mechanisms.
to nanosintering. For example, the thermodynamic Moreover, in many cases, grain growth is required in
driving force for nanosintering is extremely large, calling order to break the local interfacial energy balance
into question the use of conventional sintering doctrines necessary to sustain continuous elimination of pores.1
based on linear diffusion theories. Non-linear diffusion The literature on nanosintering can be classified into
behaviour leads to different kinetic rates of diffusion, two groups: those focused on the densification and grain
growth behaviour during sintering at the nanoscale; and
those focused on innovative sintering technologies and
Department of Metallurgical Engineering, University of Utah, 135 S. 1460 processes. Fundamental studies included molecular
E. Room 412, Salt Lake City, UT 84112, USA dynamics (MD) simulation at the atomic scale, the
*Corresponding author, email zak.fang@utah.edu kinetic studies of sintering using non-linear diffusion

ß 2008 Institute of Materials, Minerals and Mining and ASM International


Published by Maney for the Institute and ASM International
326 International Materials Reviews 2008 VOL 53 NO 6 DOI 10.1179/174328008X353538
Fang and Wang Densification and grain growth during sintering of nanosized particles

theories, and the modelling of kinetic behaviour based Thermodynamic driving force of nanosintering
on a generalised parabolic model of grain growth. In The thermodynamic driving force for sintering particles
addition, efforts have been directed toward the study of of any size is the reduction of surface energy. Based on
the grain growth of nanosized particles. conventional sintering theories, the driving force of
Innovative sintering technologies include the two step sintering can be given by6
sintering technique and various pressure assisted sinter-  
ing techniques. The two step sintering technique was 1 1
s~ck~c z (1)
designed to decouple the densification and grain growth R1 R2
processes that occur during conventional pressureless where c is the surface energy of the material, k is the
sintering. Pressure assisted sintering techniques were
curvature of a surface, which is defined by k~ R11 z R12 (for
designed to achieve full densification while retaining
nanoscaled grain sizes. Conventional pressure assisted a convex surface, it is taken to be positive; for a concave
processes, including hot pressing (HP) and hot isostatic surface, it is taken to be negative), and where R1 and R2
pressing (HIP), have been used for sintering and are the principal radii of the curvature. The driving force
consolidation of nanosized powders. Other unique for the sintering of nanosized particles is, therefore,
processes such as microwave sintering and spark plasma inversely proportional to the size of the particles. This
sintering (SPS) are widely applied in the research of relationship would lead to a much higher driving force for
sintering of nanoparticles. the sintering of nanosized particles compared to micro-
Since the early 1990s, two comprehensive reviews have metre sized particles. For example, based on equation (1),
been conducted that deal with the topic of nanosintering the driving force for a 10 nm particle is two magnitudes
or processing of nanosized particles.2,3 The issues of higher than that for a 1 mm particle.
nanosintering and the difficulties of manufacturing bulk The large driving force of nanosintering can be even
nanocrystalline materials from nanoscaled powders are higher than the result of equation (1) if the non-linear
discussed in other reviews that cover broader topics of dependency of vacancy concentrations on the particle
nanomaterials.4,5 Collectively, these reviews provide a size is considered. During sintering, mass transport,
usually mediated by vacancies, is driven by the
strong foundation for understanding the status of
difference in vacancy concentration DCv5Cv2Cv0,
technologies and science issues involved in sintering of
where Cv is the vacancy concentration for a surface
nanosized particles. Since that time, however, significant
with curvature of k, and Cv0 is vacancy concentration
progress in the field has been made. This review will focus
for a flat surface. Based on Gibbs–Thomson equation7
on the progress toward understanding the unique
 
behaviours of nanoparticles, particularly the size depen- ckV
dent properties and their effects on nanosintering. The Cv ~Cv0 exp { (2)
kT
authors anticipate that this focus will aid the community’s
efforts to find solutions to controlling and optimising where V is the atomic volume, k is Boltzmann’s constant
densification and grain growth. Specifically, we will and T is the absolute temperature. For micrometre sized
extract information from published studies, both experi- particles, the term
 ckV/kT%1,
 equation (2) becomes
mental and theoretical, and organise the information a ckV
linear, Cv &Cv0 1{ , therefore
way that highlights the differences in the sintering of kT
nanoparticles versus that of coarser particles, as well as the ckV
relationship between grain growth and densification. The DCv &{Cv0 (3)
kT
review is presented in two main sections: densification of
nanoparticles and grain growth during nanosintering. However, when particle size approaches nanoscale, the
Each section will include a discussion of the thermo- linear approximation is no longer valid. The correct
dynamic driving force and kinetic behaviours. expression for the driving force of mass transport should
be given as
 
ckV
Densification during sintering of DCv ~Cv0 exp { {Cv0
kT
nanosized particles    
ckV
In general, the sintering of nanosized or nanocrystalline ~Cv0 exp { {1 (4)
powders follows the same path as larger grain powders. kT
However, compared to conventional micrometre sized Equation (4) shows that the driving force for mass
or submicrometre sized particles, the densification transport during sintering is a non-linear function of
behaviour of nanoparticles during sintering is notably the surface curvature, and it increases exponentially
different with respect to the rate of densification and the when particle size decreases to nanoscale. Figure 1
temperature range at which densification occurs. The DCv
densification of nanoparticles is strongly affected by schematically illustrates the relationship of versus
Cv0
agglomeration of particles, pores and other processing ckV 1
variables. Although many of these factors also impact { ~  , where d* is related to the particle size. This
kT d
the sintering of conventional micrometre sized particles, non-linear relationship of the driving force for nano-
the effects are more dramatic and magnified in the case sintering is expected to have a dramatic effect on the
of nanosized particles. The densification behaviour of kinetics of sintering.
nanosintering can be analysed from the perspectives of The driving force of nanosintering is also affected by
both the thermodynamics and the kinetics of the specific surface energy c. The value of c may change as a
process. function of the particle size. Campbell et al.8 studied the

International Materials Reviews 2008 VOL 53 NO 6 327


Fang and Wang Densification and grain growth during sintering of nanosized particles

3 Oriented attachment of nanosized titania under hydro-


thermal conditions (Reprinted with permission from
Elsevier)

that because nanocrystals have significant surface area,


the problem of anisotropy becomes even more critical.
1 Schematic comparison of vacancy concentration as Although there are numerous studies reporting that
function of particle size between linear approximation nanoparticles have anisotropic faceted morphology, direct
and non-linear equation correlation of their effect on surface energy and sintering
are not available in the literature and, understandably,
effect of size dependent nanoparticle energetics on very difficult to perform. One phenomenon that has been
catalyst sintering. By using microcalorimetric measuring reported regarding nanocrystals with faceted morphology
the heat of adsorption of Pb onto MgO (100), they is that they form oriented attached assembly as shown in
showed that the surface energy increases substantially as Fig. 3.10 The oriented attachment of nanoparticles could
the radius decreases below y3 nm (Fig. 2). affect the coalescence of these particles and the densifica-
Independently, Nanda et al.9 showed that the surface tion and grain growth during sintering.
energy of nanoparticles is significantly higher than that
of the bulk, as demonstrated by studying size dependent
Kinetic behaviour of nanosintering
evaporation of Ag nanoparticles. Sintering temperature
From another aspect, the surface energy is also a Notably, the sintering of nanosized particles occurs at
function of the crystal orientations, which could affect lower temperatures than the sintering of conventional
the driving force of nanosintering. Groza2 pointed out micrometre sized or submicrometre sized powders, as
shown schematically in Fig. 4. Sintering temperature is,
in general, a loose concept, referring to the entire
temperature range of densification. In order to be
specific and quantitative, the starting temperature is
often used for comparison. However, because sintering
and densification are continuous kinetic processes,
rigorously speaking, a single point of demarcation for
the starting temperature of sintering does not exist.
Based on typical experimental behaviour, the starting
temperature can be defined as the temperature at which
the rapid densification stage initiates, as marked on
Fig. 4. In general, the densification versus temperature
plot shifts to the left (lower temperature) when
nanosized powders are used rather than micrometre
sized powders. For example, several studies on the
sintering of nanoyitrium stabilised zirconia (YSZ) have

2 Illustration of discrepancy between measured differential


heat of adsorption of Pb onto MgO(100) and prediction
based on constant c model: surface energy increases
substantially as radius decrease below y3 nm (Reprinted
with permission from American Association for the 4 Schematic diagram illustrating different onset tempera-
Advancement of Science) tures of sintering of nano- and micrometre sized particles

328 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

5 Per cent of densification of WC–Co as function of continuous heating temperature for various initial particle sizes:
After Ref. 17

shown that the sintering temperature of nanocrystalline the melting temperature of very fine particles decreases
ZrO2 initiates at a temperature 200 degrees lower than with the size of particles.25–40 In the case of nanosin-
that of the microcrystalline powders.11–13 An even tering, the decreasing onset temperature of sintering
greater temperature difference of sintering, 400uC, was can be understood, therefore, on the basis of the
reported by Mayo14 for nanosized titania compared to lower melting temperature of nanoparticles. Troitskii
commercial TiO2 powders. Similar results were observed et al.41 studied the initial sintering temperature of
when sintering nanoceria15 and nano-titanium nitride different sized TiN powders and found the relationship
powders.16 between initial sintering temperature Tis and particle
A comprehensive study on the sintering of nanotung- size r is
sten carbide and cobalt (WC–Co) powders was con-
ducted by Muheshwari et al.17 Figure 5 shows the per Tis (r)~T(?)exp½{c(r0 {r)=r (5)
cent of densification as a function of the continuous where T(‘) is the initial sintering temperature of coarse
heating temperature for various initial particle sizes. particles; c is a constant determined by the properties
Clearly, the entire sintering temperature decreases of the material and the energetic state of the surface
steadily as the initial average particle size decreased layer; r9 is arbitrary size.
from 30 mm to 10 nm. It seems, however, that there is Jiang and Shi42 ascribed the size dependent initial
little difference between the onset temperatures of the sintering temperature of nanosized particles to the
sintering of particles greater than 1 mm. decreased melting temperature of nanosized particles
Relatively speaking, fewer systematic studies exist on based on the relationship
the densification behaviour of nanosized metal powders
than on nanoceramic powders.18–20 A direct and Tis (r)~0:3Tm (r)
systematic study of the sintering of nanocrystalline Fe  
2Sm (?) 1
and Cu powder was reported by Dominguez et al. in ~0:3Tm (?) exp { (6)
3k (r=r0 ){1
1998.21 Compared to micrometric powder, the rapid
densification of nano-Fe and Cu powders started at where Tm(‘) is the melting temperature of the bulk
y200uC lower temperatures. A reduction of the sinter- material, Svib(‘) is the bulk melting entropy, r is particle
ing temperatures was also reported for nanotung- radius, r053h (where h is atomic diameter) for nano-
sten.22,23 The reported sintering temperature of particles, k is Boltzman constant. Figure 6 shows the
nanosized tungsten powders produced by high energy predicted size dependent initial sintering temperature of
mechanical milling was dramatically decreased from some metallic powders by using equation (6).42 Note
conventional temperature of 2500 to 1700uC. More that the significant changes of the initial sintering
surprisingly, Oda et al. showed that the nanosized temperature do not occur until the particle size is less
tungsten powders could be sintered even at 1000uC by than y20 nm.
using SPS. The authors’ own recent work on sintering of
nanosized tungsten powder demonstrated that nano- Scaling law: dependence of nanosintering on particle size
tungsten with y20 nm initial sizes can be sintered up to In conventional sintering theories, the dependence of
.98% relative density without pressure at 1100uC in a densification behaviour on the size of particles is
hydrogen atmosphere.24 described by the scaling law. In 1950, Herring44 first
General sintering theories hold that a material’s introduced the scaling law as follows
sintering temperature is often correlated with the
material’s melting point. It has long been known that Dt2 ~ln Dt1 (7)

International Materials Reviews 2008 VOL 53 NO 6 329


Fang and Wang Densification and grain growth during sintering of nanosized particles

6 Theoretical prediction of initial sintering temperature


Tis(r) for selected metals in terms of equation (6): 7 Densification behaviour of nanocrystalline TiO2 with
experimental initial sintering temperature of W, Ni and three different agglomerate sizes: note that the larger
Ag are also plotted in figure for comparison:18,43 After the agglomerate size, the higher the sintering tempera-
Ref. 42 ture (agglomerate size in bold, crystallite size in light).
For the non-agglomerated (N/A) powder, sintering time
is 120 min;52 for the 80 and 340 nm agglomerate pow-
where l5R2/R1, R1 and R2 are particle radius, n depends
ders, sintering time is 30 min:53,54 After Ref. 3
on specific diffusion mechanisms of the densification.
Specifically, n51 for viscous flow, 2 for evaporation and
condensation, 3 for volume diffusion and 4 for surface flux. Furthermore, the scaling law considers different
diffusion or grain boundary diffusion. The scaling law mechanisms separately. In practice, multiple mechan-
states that the time required to sinter powders with a isms of sintering may occur simultaneously, especially
particle radii of R1 and R2 is proportional to the ratio of in nanosintering.
the particle radius. Although the densification behaviour Effect of green density, agglomeration on nanosintering
of nanopowders can be qualitatively understood on the
In the practice of nanosintering, the densification
basis of the scaling law, few direct analyses of
behaviour of nanoparticles are affected not only by the
experimental data exist in the literature. The few studies
intrinsic nature of the nanoscale size of the particles, but
that did apply the scaling law used the following
also by the processing conditions and related difficulties,
expression to analyse the activation energies of the
such as green density and agglomeration.
sintering of nanopowders45,46
First, similar to powder compacts of micrometre sized
   
d1 Q 1 1 powders, the densification of a powder compact depends
nln ~ { (8) significantly on the green density of the compact. Green
d2 R T2 T1
density must be sufficiently high in order to achieve
where d1, d2 are particle sizes, T1 and T2 are adequate densification under similar sintering condi-
corresponding sintering temperatures, R is the gas tions. On the other hand, the finer the particle sizes, the
constant and Q is the activation energy. By using the lower the green density of powder compacts assuming
above equation, some studies obtained activation energy the compaction pressure is the same. Therefore, it is
values that are closer to grain boundary diffusion, while generally observed that the sintering of nanosized
others obtained values closer to volume diffusion, which powders is affected by compaction pressure.2,50,51
is believed to be unlikely at low temperatures. These It has been widely recognised that agglomeration of
discrepancies in the values of activation energies nanoparticles has a critical impact on the sintering of
obtained using this method, therefore, raise questions nanoparticles. Owing to the extremely fine size and the
on the validity of the scaling law for sintering of strong interactive force between particles, nanoparticles
nanoscale particles. tend to form agglomerates. The size and strength of the
The derivation of scaling law is based on an agglomerated particles affect the densification rate. The
assumption that the particle size of two different most direct investigation of the agglomeration of
powder systems used for comparison does not change densification was summarised by Mayo3 whose data
during sintering and microstructural changes remain was based on numerous published experimental results
geometrically similar for the two systems. This as shown in Fig. 7.
assumption points out the key limitations of using In essence, a powder compact can be viewed as
the scaling law: it is difficult to maintain the conditions consisting of a bilevel hierarchical structure: the
in the assumption in real powder systems.47 During compact is made of agglomerates which are made of
nanosintering, particle size changes rapidly and some nanosized particles. There is, therefore, a bimodel pore
characteristics of nanoparticles, such as agglomeration size distribution. The pores existing within agglomerates
and the non-uniformity of green density, increase the are finer than the pores between agglomerates. The
difficulty of maintaining similar microstrutural evolu- densification of an individual agglomerate is relatively
tions during sintering. In addition, the classic scaling easy, while the elimination the interagglomerate pores is
law was derived from linear equations; thus the validity more difficult. By tracking the evolution of pore size
of the scaling law during nanosintering is questionable. distributions, Peterson et al.55 studied the sintering of
The non-linear diffusion behaviour of nanosintering48 fine grain cemented tungsten carbide and cobalt system
and the size dependent diffusion activation energies49 (WC–Co). They showed that during the intermediate
will change the expression for calculating the diffusion stage of sintering, the considerable densification

330 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

8 Derjaguin–Landau–Verwey–Overbeek (DLVO) potential


(green line) is sum of attraction potential and repulsion
potential (Reprinted with permission from Nature
Publishing Group)

a surfaces repel strongly; small colloidal particles remain


obtained is primarily connected to removal of small ‘stable’. b surfaces come into stable equilibrium at sec-
pores rather than shrinkage of larger ones. ondary minimum if it is deep enough; colloids remain
From the perspective of achieving full densification kinetically stable. c surfaces come into secondary mini-
and elimination of pores, it is logically desirable to de- mum; colloids coagulate slowly. d the ‘critical coagula-
agglomerate powders or to avoid the formation of tion concentration’. Surfaces may remain in secondary
agglomerates in the first place. Fundamentally, the minimum or adhere; colloids coagulate rapidly. e sur-
formation of agglomerates is attributed to balance of faces and colloids coalesce rapidly
the interparticle forces, specifically the van der Waals 9 Schematic energy versus distance profile of DLVO
force, which acts to bind particles, and the electrostatic interaction: After Ref. 57
repulsion which opposes agglomeration. Figure 8 illus-
trates potential energy between nanoparticles as a A compact consists of particles and pores, and each
function of the distance between them.56 Agglomerates pore has a volume, shape and a coordination number.
of particles form when either the primary or the The pore coordination number is defined as the number
secondary minimum condition is met. Agglomerates at of touching particles surrounding and defining each void
the secondary minimum are kinetically stable while space. A pore’s surface morphology is determined by the
agglomerates at the primary minimum are thermodyna- dihedral angle and the pore’s coordination number. In
mically stable. Figure 9 shows that lowering the particle general, for a given dihedral angle, a critical coordina-
surface charge by increased ion binding or increasing the tion number nc, exists that defines the transition of the
salt concentration could induce two particles to come pore surface morphology from convex (n.nc) to
into adhesive contact at the primary minimum.57 concave (n,nc). Kingery and Francois61 first recognised
To de-agglomerate, however, opposite measures must that only those pores with n,nc are able to shrink
be taken in order to stabilise a colloidal solution. In during sintering because the concave surface morphol-
other words, the repulsive forces must be boosted to ogy with negative chemical potential is thermodynami-
achieve a balance between the attractive and repulsive cally unstable. As a result, atoms will diffuse to the pore
forces such that the dispersion of particles can be surface and fill the void space. Figure 10 illustrates the
stabilised. Methods for nanoparticle dispersion include stability of pores and their dependence on both the
electrostatic charge stabilisation, steric stabilisation, or a dihedral angles and the coordination number. For any
combination of the two. Details on the principles of
stabilising colloidal solutions can found elsewhere.58
These techniques can be implemented in the processes,
including mixing and milling of the powders that are
necessary prior to sintering of these materials.
To sinter nanoparticles for the fabrication of bulk
engineering components, a colloidal solution must be
dried; agglomerates will inevitably form. Ideally, the
agglomerates are soft and the interagglomerate pores are
small. Lange provided a more extensive discussion on
ceramic powder processing techniques for avoiding
agglomeration and achieving uniform pore distributions
within a powder compact.59

Effect of pores on nanosintering


A common thread for the effect of green density and
agglomeration on sintering is the effect of pores on 10 Relationship between dihedral angle and critical pore
densification.60 coordination number: After Ref. 61

International Materials Reviews 2008 VOL 53 NO 6 331


Fang and Wang Densification and grain growth during sintering of nanosized particles

given dihedral angle, which is dictated by the material, a deform under a constant tensile stress at temperature T1
critical coordination number exists below which the and then rapidly raised or lowered the temperature to a
pores will shrink and above which the pores will grow. new value T2 where the sample was allowed to further
The effects of green density and agglomeration on deform. The deformation rates corresponding to the
densification can be explained by the pore coordination temperatures are then determined for the time in which
number theory. Higher green density and less agglom- the temperature is changed. According to this method,
eration result in fine and uniform pores that shift the the activation energy of the responsible mass transport
pore coordination number distribution from high values phenomenon can be determined using the relationship
to low values, i.e. more pores fall into the category  
below the critical pore coordination number. These RT1 T2 T1 (d e1 =d t)
Q~ ln (11)
pores are easily removed during sintering and, thus, lead T1 {T2 T2 (d e2 =d t)
to denser products. As for the large pores that are where T1 and T2 and are the temperatures of the sample
thermodynamically stable and have coordination num- before and after rapid heating, (de1/dt) and (de2/dt) are
bers higher than the critical value, a process by which the respective shrinkage rates immediately before and
the coordination number can be reduced during sinter- after the temperature rise, R is the universal gas
ing is essential, since the pores will again become constant, and Q is the activation energy. The tempera-
unstable and the densification will then continue. ture difference between T1 and T2 should not exceed
Particle rearrangement and grain growth are the two 50uC.
processes that can play this role, creating a dilemma, of To evaluate activation energies of kinetic processes
course, and difficulty for any attempt to achieve during continuous heating, a number of other methods
maximum densification without grain growth. can be found in literature.65–67
Densification mechanisms during nanosintering Analysis of mechanisms of densification
Calculations of activation energy Using the various methods described above, activa-
In order to understand the mechanisms of sintering, tion energies for sintering various nanosized powders
activation energy is commonly used to indicate the were reported. For example, a very low activa-
internal mechanistic process. In most cases, activation tion energy for densification is observed in initial
energy can be calculated from isothermal experimental sintering: y234 kJ mol21 for nanocrystalline Al2O3 and
data. A derivation by Johnson et al.62 resulted in a 96?2 kJ mol21 for nanocrystalline TiO2,68 268 kJ mol21
densification equation based on two sphere models and for nanocrystalline ZnO,69 66?2 kJ mol21 for nanocrystal-
both volume and grain boundary diffusion mechanisms. line nickel,70 82 kJ mol21 for nanocrystalline a titanium
The initial stage sintering of a powder compact is and 49 kJ mol21 for nanocrystalline b titanium.71
described by Theunissen et al.63 Using these activation energy values to deduce
    sintering mechanisms has inherent shortcomings with
(Dl=l0 ) {mEa
~const exp (9) respect to accuracy, because in almost all cases, multiple
T RT
mechanisms may be operating simultaneously.
where m is a constant characteristic for the sintering Especially during the early stages of sintering, when no
mechanism, i.e. 1/2 for volume diffusion and 1/3 for one clearly dominant mechanism can be identified, the
grain boundary diffusion, Ea is the apparent activation activation energy values calculated using the kinetic data
energy for densification, Dl/l0 is the relative shrinkage gives an effective activation energy value that is the
during continuous heat-up and R and T are the gas combined effect of multiple mechanisms.
constant and absolute temperature respectively. It can be seen from the data, however, that the
Linearisation of the equation results in63 majority of studies point toward lower activation
  energies for early stages of sintering. This is reasonable
(Dl=l0 ) {mEa
ln ~ zconst (10) for the obvious reason of the huge surface areas and the
T RT expected high activity of nanoparticles. Surface diffusion
On plotting (2ln[(Dl/l0)/T]) versus 1/T, a straight line is one of the most cited mechanisms that contribute to
with slope mEa/R is obtained when only one single the sintering of nanosized particles. However, in
mechanism is operative. conventional sintering theories, surface diffusion is
Obviously, isothermal relationships of densification as believed to induce initial neck formation between
a function of time at several temperatures are required particles, but not densification. In principle it is correct
to obtain relatively reliable data of activation energy. An that surface diffusion will cause the bonding of particles
inherent assumption of this approach is that the to each other, but not the dimensional shrinkage (i.e. the
mechanisms of densification, or any other process of densification) of a monosized spherical particle compact.
interest, should not change within the temperature range This seemingly conflicting theory on the effects of
of the experiments. Otherwise, the calculated value does surface diffusion on sintering of nanoparticles can be
not represent a single mechanism; rather, it is an understood from a perspective of the indirect role of
effective activation energy that results from multiple surface diffusion to densification.
mechanisms. First of all, the rapid and active surface diffusion may
As stated earlier, densification and grain growth lead to grain boundary slip and rotation of particles that
during sintering of nanopowders take place rapidly may result in the rearrangement of particles, hence the
during the heating-up process. In order to capture the increased density of the compact. The possibility of
changes during continuous heating, the Dorn method64 grain boundary slip and rotation was mentioned in
can be used to calculate activation energy. To determine numerous sintering studies.2,3,63 Evidence of nanoparti-
the activation energy of creep, Dorn allowed a sample to cle coalescence via surface diffusion was presented in

332 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

words, when the coordination number reaches a critical


value, the pore is at equilibrium. The shrinkage of the
pore (i.e. sintering) can not progress until the equili-
brium condition can be tipped in favour of sintering by
grain growth. With respect to the densification mechan-
isms of nanoparticles, surface diffusion can cause
coarsening of nanoparticles which, in turn, contributes
to the process of densification. Therefore, it can be
stated that by inducing coarsening, surface diffusion will
contribute indirectly to densification. Further discussion
on the relationship between grain growth and densifica-
tion will be given later.
Surface premelting is another mechanism that could
lead to rapid densification at low temperature during
nanosintering. Owing to a large surface to volume ratio
in nanoparticles, surface premelting can happen at low
temperature, and as a result, particle rearrangement is
facilitated by sliding, rotation, or viscous flow. Alymov
et al.18 calculated the dependence of the melting point of
a particle as a function of its size using the following
equation
Tm =T0 ~
 
1{2Q{1 r{1
s ssl =(r{d)zslg r{1 (1{rs =rl ) (12)
where T0 is the bulk melting point of the solid, Q is its
latent heat of fusion. ssl and slg are the interfacial
surface tensions between the solid and the liquid and
between the liquid and its vapour respectively, and rs
and rl are the densities of the solid and liquid
respectively, r is the radius of particle and d is the
11 Particle chain sintered with no reorientation: gap thickness of melted layer on a particle surface.
between particles (.1) was filled by surface diffusion Given that the sintering temperature is proportional
that has also roughened the middle particle’s surfaces to the melting point, it is generally understood that as
(.2)73 (Reprinted with permission from Materials the melting point decreases, the sintering temperature
Research Society) decreases. It has been demonstrated that the melting of a
particle with diameter d will result in coagulation with its
Shi’s study on barium titanate72 and Bonevich and neighbours and will become the centre of a new, larger
Marks study of Al2O3,73 using TEM. Figure 11 shows particle. In an independent study of the sintering of
the formation of the neck between two Al2O3 particles nanometric Fe and Cu, Dominguez et al.21 attributed
after sintering at 1000uC for just a fraction of a second. the initial densification to surface melting mechanisms
This study shows that surface diffusion is the predomi- because the activation energies that were obtained from
nant mechanism for sintering, as evidenced by the fact either constant heating or isothermal experiments were
that the faceted interfaces are similar to ledge growth, too small to ascertain lattice diffusion mechanisms. In
and the sintered particles retain their initial adhesion fact, Dominguez et al. claimed that the activation energy
structure with no reorientation occurring during sinter- measured during the course of their experiments is very
ing. The driving force for sintering can be considered a small, compared to the values published for the self-
chemical potential difference between facet surfaces and diffusion of the metal in a liquid state. Therefore, it was
the neck region. reasoned that the presence of a liquid-like layer on the
The indirect role of surface diffusion on densification surface of the nanometric particles during sintering
can also be understood based on theories of the could simultaneously explain such phenomena as high
relationship between coarsening and sintering of parti- diffusivity and enhanced grain growth at a narrow
cles.1,72,74,75 As discussed earlier, effects of pores on temperature range.
nanosintering, according to theories first proposed by Although surface melting is a reasonable mechanism
Kingery and Francois and further elaborated by Lange for sintering nanosized particles, no direct experimental
et al.,60,61 a pore will shrink during sintering only if the evidence supporting the formation of liquid phase has
coordination number of the pore is smaller than a been published to date. In fact, because the formation of
critical value n,nc because only then is the surface of the any liquid at temperatures below equilibrium melting
pore concave. Thermodynamic driving force dictates point will be thermodynamically unstable, there could
that mass will diffuse from convex surfaces to concave be only transient liquid, making experimental verifica-
surfaces. Initial sintering of a compact will develop an tion of the presence of liquid phase even more difficult.
equilibrium configuration at which the driving force for A more generally applicable theory that explains the
further sintering is zero. Grain growth, or coarsening, rapid densification of nanoparticles is based on the
will perturb the equilibrium configuration to reinitiate hypothesis of non-equilibrium high concentration of
neck growth (sintering) and densification. In other vacancies at the interparticle grain boundaries. In 1974,

International Materials Reviews 2008 VOL 53 NO 6 333


Fang and Wang Densification and grain growth during sintering of nanosized particles

Vergnon et al. studied the ‘initial stage for the sintering Csolid, and the atomic spacing a can be given by
of ultrafine particles TiO2 and Al2O3’.68 Using flash  
2D aF
sintering and isothermal experimental techniques, it was J~ sinh (13)
shown that during the first 20 s, a fraction of the total aV 2kT
observed shrinkage, up to 95%, was registered.68 There where D is the diffusion coefficient, V is the atomic
was an initial loss of surface area, before the shrinkage volume, a is the atomic spacing, F is the driving force for
starts during the heating of the compact to the desired diffusion, k and T are the Boltzmann constant and
temperature, a process which requires only a few absolute temperature respectively. Pan pointed out that
seconds. It was reasoned that this almost instantaneous this equation reduces to linear diffusion law, when
loss of the surface area corresponds to the formation of aF(kT, sinh (aF/2kT)<aF/2kT. However, when particle
junction zones between particles of the compact. The sizes are in the range of nanometres, the linear approx-
fast formation of the junctions between particles, before imation is no longer reasonable. Then, the diffusion
the shrinkage onset, involves the creation of a high equation becomes a non-linear equation that can only be
concentration of vacancies inside these junctions. The solved via numerical methods. Applying this approach to
shrinkage of the compact results then from a decrease of sintering two particles, the ratio of the neck to particle
the distance between the centres of particles due to radius as a function of the length of time at a given
annihilation of the trapped vacancies in the junction temperature was calculated and shown by Fig. 12. The
zone. Because the concentration of trapped vacancies differences between prediction by linear solutions and
inside the junction zone largely exceeds the thermo- non-linear solutions are significant during the initial stage
dynamic equilibrium concentration, the diffusion can be of sintering and diminish as sintering time increases. The
considered as independent of time and controlled only distinction between linear and non-linear solutions also
by the probability of jumping of ions, as long as the diminishes as particle size increases.
concentration of vacancies exceed the equilibrium The rapid rate of sintering due to the rapid rate of
content. Any further sintering, after the initial non- diffusion is also supported by recent studies which
equilibrated concentration of vacancies is exhausted, indicate that the coefficient of diffusion D, is size
corresponds with the diffusion of equilibrated vacancies. dependent as shown49
Furthermore, based on the theory that excessive
concentration of vacancies exist (c.1024), Trusov D(r,T)~
et al.20 stipulated that there is a possibility of liquid-   
{E(?) {2Svib (?) 1
like merging (coalescence) of particles into large ones. D0 exp exp (14)
Liquid-like coalescence, as well as slippage, causes the RT 3R r=r0 {1
compact shrinkage of ultrafine particles. where D0 is pre-exponential constant, E(‘) is bulk
In another study focusing on size dependent grain activation energy, Svib(‘) is bulk melting entropy, r is
growth kinetics observed in nanocrystalline Fe, Krill particle radius, r053h (where h is atomic diameter) for
et al.76 also established their model on the basis of nanoparticles, R is ideal gas constant and T is absolute
existence of excess volume at the grain boundaries. temperature.
The ‘excess’ volume is in the form of vacancies, which The dependence of the coefficient of diffusion on
leads to a non-equilibrium vacancy concentration. The particle size is attributed to that, as the size of the
issues of grain growth of nanoparticles during sintering nanocrystals decreases, the activation energy of diffu-
will be further discussed in latter sections of this sion decreases and the corresponding coefficient of
review. diffusion increases based on the Arrhenius relationship
Finally, the rapid densification mechanisms of nano- between them. Together these theories, based on non-
particles is also related to the preferential crystalline linear diffusion law and the increase of the coefficient of
orientations. It has been observed that in loose nanocrys- diffusion with decreasing particle size convincingly
talline powders, the first neck formation occurs, not argue for the rapid formation of necks bonding with
randomly between particles, but by the orderly mating of neighbouring particles.
parallel, crystallographically aligned facets on the particle With the development of modern tools of computa-
surfaces.73,77 Some nanocrystalline powder compacts also tional materials science, considerable effort has been
appear to reflect a kind of ordered structure resulting from made toward numerical simulation of the sintering of
less than random type matings of particles during the nanoparticles. One of the most notable works was
initial stage of sintering.78 published by Zhu and Averback in 1995,79,80 who
simulated the sintering of YSZ using the MD approach.
Kinetic theories, modelling and simulations of Molecular dynamics simulations of sintering have been
nanosintering conducted in numerous other materials as well.81–86 The
Given the unique physics that present when sintering basic approach toward simulating sintering using MD
nanoparticles, considerable work has been reported on method involves tracking the motion of atoms under
the theoretical modelling of thermodynamics and the stress caused by surface or interfacial energy. The
kinetics of sintering of nanosized particles. In one of the kinetics of sintering is given as the rate of decreasing
more significant studies, Pan recognised that the rapid distance between two atoms in the middle of two
kinetic rate of sintering is a direct result of the large particles in contact. It was shown that sintering
driving force for sintering of nanosized particles, and nanoparticles at the atomic level can be accomplished
revised the two sphere sintering model by using non- by dislocation motion and grain boundary rotation, as
linear diffusion law.48 Because the diffusion is the result well as other mechanisms. It was further predicted that
of jumping atoms, the flux of diffusion as a function of the the sintering time of nanoparticles would be in the range
frequency of jumping f, volume atomic concentration of a few hundred picoseconds. Although the predicted

334 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

a R55 nm; b R510 nm; c R5100 nm; d R51000 nm


12 Comparison between linear solution (broken line) and non-linear solutions (solid line) for shrinkage between two
spherical particles as functions of time (Reprinted with permission from Taylor & Francis)

sintering time is far from engineering reality, the results particles, especially during the later stages of sintering,
of the simulation can be used as a basis for under- further details on simulation of sintering is beyond the
standing the initial bonding or formation of the necks scope of this review, which focuses on the size dependent
between nanoparticles. characteristics of sintering of nanoparticles.
Owing to the limitations of studying sintering by using
atomic scale simulation, multiscale simulation is con- Grain growth during nanosintering
sidered a promising approach. Pan87 categorised the
simulations of sintering in three levels: atomic level, Issues of grain growth during nanosintering
particle level and component level. The primary method A primary motivation for studying sintering of nano-
of atomic level simulation is MD, as mentioned earlier. sized particles is rooted in the issue of rapid grain
In particle level simulation, which is also classified as growth during sintering. In many cases, particularly
mesoscale simulation,88 classic sintering models6,62,63,89–101 when the goal is to produce nanocrystalline bulk
based on mass transport between two or multiple materials, the objective of nanosintering is to achieve
particles provide the basis of simulation. The kinetics of full densification as well as the retention of nanoscaled
sintering densification depends on the kinetics of specific grain structure in the sintered material. Research has
diffusion mechanisms that control the rate at this stage generally shown that after sintering, nanosized particles
of the process. In Pan’s review, the typical numerical lose nanoscale characteristics because grain size grows to
methods used in particle level simulations include the greater than 100 nm. Therefore, understanding and
finite difference method, the variational calculus, and controlling grain growth is a critical scientific and
the finite element method. To model the evolution of the technical issue of nanosintering.
microstructure during sintering, however, Monte Carlo In a systematic study of the stability of nanosized
method is often a popular choice.102–106 The finite metal powders, Malow and Koch107–109 reported that
element method is the primary method for simulation at the rate of grain growth of nanocrystalline Fe powders
the component level, which is chiefly concerned with made by ball milling is initially very rapid (,5 min)
macro scale shrinkage, distortion and dimensional when annealed at various temperatures. Grain growth
accuracy control. then stabilises during extended isothermal holding (up to
Overall, most published studies of sintering simula- 142 h). During isothermal holding, grain growth follows
tions were not nanoscale specific, except for the MD a generalised parabolic grain growth law and is similar
simulations described above. Although these methods to that found in bulk materials. It is noted, based on
are arguably applicable to sintering of nanosized Fig. 13, that at the first data point of the isothermal

International Materials Reviews 2008 VOL 53 NO 6 335


Fang and Wang Densification and grain growth during sintering of nanosized particles

13 Evolution of grain size as function of annealing time at three annealing temperatures for nanocrystalline iron: grain
size was determined by Scherrer equation (Reprinted with permission from The Minerals, Metals & Materials Society)

annealing curves at higher annealing temperatures temperatures and that it accelerates dramatically when
(825 K), the grain sizes are already several times (3– the temperature is above an apparent critical tempera-
66) greater than the original as milled grain size ture range. Figure 16 provides another example of the
(y8 nm). In other words, grains grow rapidly during relationship between grain size and temperature during
heat-up, prior to reaching the preselected isothermal heating up of nanocrystalline WC–Co powder at a
holding temperature. heating rate of 10uC min21.111 It shows that the original
In another study of the grain growth of nanocrystal- 20 nm grain size has increased almost 45 fold to 900 nm.
line Fe using in situ synchrotron X-ray diffraction This ‘explosive’ grain growth occurs almost instantly
techniques, Krill et al.76 further demonstrated that grain during heat-up, with no significant holding time. Similar
growth of nano-Fe particles is comprised of three steps: behaviour has also been reported for sintering of other
the ‘initial growth spurt’, a linear growth stage and the nanocrystalline ceramics, as well as for metallic pow-
normal parabolic stage, as shown in Fig. 14. Once again, ders.112–118 It appears that a critical temperature exists,
the normal parabolic stage can be modelled using the above which the grain growth accelerates dramatically
classic grain growth parabolic law, however the ‘initial as a function of temperature.
growth spurt’ of nanocrystalline Fe during annealing The unique issues of grain growth during sintering can
was not captured by isothermal studies. be studied by examining the grain size versus relative
Grain growth during nanosintering is also a strong density relationship. In one of the earliest studies of the
function of temperature. Figure 15 shows the relation- sintering and grain growth of nanosized ceramic
ship between grain size and temperature during the heat powders in 1990s, Owen and Chokshi119 and Averback
treatment of nanocrystalline cobalt powder.110 It is et al.120 showed that oxides densify without significant
obvious that the grain growth is initially slow at very low grain growth until the density reaches y90% of the bulk
density. Then the grain growth becomes very rapid. This
phenomenon is observed in many different materi-
als.2,3,121,122 A typical relationship between grain size

14 Size dependent grain growth kinetics observed in 15 Change of mean grain size (linear intercept) with
nanocrystalline Fe (Reprinted with permission from annealing temperature, measured in pure nanocrystal-
American Physical Society) line Co (Reprinted with permission from Elsevier)

336 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

I: early stage of grain growth; II: late stage of grain


growth
17 Relationship between grain growth and densification
16 Grain size versus temperature during heating up of during nanosintering
nanocrystalline WC–Co powder at heating rate of
10uC min21: After Ref. 111 controlled. The result also shows that at the starting
point for the normal grain boundary controlled grain
and density during nanosintering is schematically shown growth, grain size has grown to 4–6 times the initial
in Fig. 17. This relationship implies that the grain grain size of the nanopowder. This part of grain growth
growth during sintering consists of two stages: the early is attributed to coarsening.
stages of sintering, before the powder compact reaches Another approach for studying the relationship of
90% relative density; and the late stages of sintering, densification versus grain growth was summarised by
when relative density is greater than 90%. It is believed German et al. in a recent study on sintering of nanosized
that the late stages of grain growth can be viewed as tungsten powder.124 The dynamics of densification and
‘normal’ grain growth, similar to that found in bulk grain growth is described by the continuum theory of
materials by boundary migration, but incorporating the sintering98 as follows
effect of pinning by closed pores. In contrast, the early d r 9ch
stage of grain growth during sintering is often referred to ~ (15)
d t 4gG
as ‘coarsening’.
Figure 18 shows another example of grain growth where r is sintered density, h is porosity, c is the surface
versus densification during the sintering of nanocrystal- tension, g is the effective porous body viscosity and G is
line Y2O3 materials.123 This work demonstrated that the grain size. In order to apply the above equation, the
normal grain growth stage during sintering, the mechan- relationship between the grain size and densification is
ism of which will be discussed later in this section, can be described by another relationship

18 a increasing grain size of Y2O3 with density in normal sintering. (Heating schedule shown in inset). Even with fine
starting powders (30 nm), the final grain size of dense ceramics is well over 200 nm regardless of whether dopant
was used. The shaded area indicates the grain size regime commonly defined as nanostructured materials. At lower
densities, the mean grain (particle) size was estimated on the fracture surface. At higher densities, the grain size was
obtained by multiplying by 1?56 the average linear intercept length of at least 500 grains. b grain size of Y2O3 in two
step sintering. (Heating schedule shown in inset.) Note that the grain size remains constant in the second sintering
step, despite density improvement to 100% (Reprinted with permission from Nature Publishing Group)

International Materials Reviews 2008 VOL 53 NO 6 337


Fang and Wang Densification and grain growth during sintering of nanosized particles

 n{1 90%) is significant and sufficient to reach beyond


dG G0 nanoscale. Thus, if grain size is to be maintained at
~xK h{(n{1)=2 (16)
dt G nanoscale, this part of grain growth must be con-
where x is a material constant, K is the reference grain trolled. However, it appears from the search of open
growth rate, G0 is the initial particle size, and n is a literature that although the early stage of grain growth
material constant. By solving equations (15) and (16), process is critical to nanosintering, few published
both the density and grain size can be evaluated. studies to date focus on this aspect of grain growth.
Yet another parametric approach for studying the In the following sections, we will examine the issues of
relationship between the densification and grain growth grain growth during nanosintering in both the ‘normal’
is based on the master sintering curve method, which and ‘initial’ stages.
describes the sintering densification as follows125,126
  Normal grain growth during nanosintering when
dr cV dDb Cb Dv Cv relative density .90%
~ z 3 (17)
3r d t kB T G4 G In general, when the relative density is greater than 90%
where c is surface energy, kB is Boltzmann constant, Db during sintering of micrometre sized particles, bonding
and Dv are grain boundary and volume diffusivity, T is between powder particles by neck growth is well
absolute temperature and V is atomic volume, G is mean developed and the majority of pores within a sintered
grain diameter. Cb and Cv are scaling parameters, which body are closed. Grain growth during the continued
relate the driving force, mean diffusion distance, and sintering densification process is akin to that of bulk
other geometric features of the microstructure on which materials during heat treatment, although the remaining
the sintering rate depends on mean grain diameter G. By pores can still hinder the kinetic rate of grain growth
rearranging the above equation, equation (17) becomes during this stage. Consequently, this stage of grain
ðr ðt   growth is referred to as ‘normal’ grain growth.
½G(r)n cVD0 Q Typically, because the properties of sintered materials
d r~ exp { dt (18)
r0 3rC(r) 0 kT RT depend on their final grain sizes, the final stage of grain
growth during sintering has received, therefore, the most
where r0 is the green density of powder compact, Q is attention.
apparent activation energy, R is the gas constant. For
volume diffusion, D05(Dv)0 and n53; for grain bound- Grain growth law: kinetics of normal grain growth
ary diffusion, D05(dDb)0 and n54. Grain size G is given Classic treatment of grain growth in solid materials was
as a function of the density. Once again, if the reviewed in several studies.7,47,128,129 It is generally
relationship between the grain size and density is known, believed that the mechanism of grain growth in bulk
the model would be complete. Using this approach, Park materials is by curvature driven grain boundary migra-
et al.127 further expressed the relationship of grain size as tion, and the rate of grain growth is inversely propor-
functions of time and temperature as follows tional to grain size130
ðt  
Q dG M
G 3
~G03 z3 K0 exp { dt (19) ~ (20)
RT dt G
0
where M is a constant. By integrating over time, the
where G is grain size, G0 is initial grain size, K0 is pre- classic parabolic grain growth law is
exponential factor, Q is activation energy; R is gas
constant and T is absolute temperature. The advantage G(t)2 {G02 ~kt (21)
of the above relationship is that it is not limited to where k5k0 exp(2Q/RT) is the rate constant of grain
isothermal conditions. It can be applied to continuous growth that is a function of the mobility of grain
heating conditions. However, the above relationships boundary, k0 is a constant, Q is activation energy for
were developed for general situations using the equal grain boundary migration, R is gas constant, and T is
size two sphere model without explicitly considering the absolute temperature.
effects of nanoscale particle sizes. In practice, accounting for the fact that the experi-
In short, from a kinetic perspective of grain growth mental observed grain growth data do not always fit
as a function of time, the experimental observations as equation (21) with the exponent equals to two, the
described above suggest that the grain growth of parabolic grain growth law is usually generalised as
nanosized particles during sintering can be treated as
consisting of two steps: a dynamic grain growth G(t)n {G0n ~kt (22)
process that occurs during heating-up and at the
beginning of isothermal duration; and the static grain Equation (22) is thus termed as generalised parabolic
growth during isothermal holding. From another grain growth model.131 When n is variable, grain growth
perspective of the interrelations of grain growth to kinetics can be better described by equation (22).
densification, the grain growth during sintering con- Considering the effect of factors such as segregation
sists of two stages: first, when the relative density is of impurities on grain boundaries that mitigate grain
lower; and second, when the relative density is greater growth, the dependence of grain size on time is further
than 90%. It should be emphasised that although the modified as132–134
late stage of grain growth, when relative density is  
G0 {G(t) G? {G0
greater than 90%, accounts for the majority of total zln ~kt (23)
G? G? {G(t)
grain growth, early stage grain growth that occurs
during heating (when relative density is still lower than where G‘ denotes the grain size when the grains cease to

338 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

The activation energies calculation, based on the rate


constant deduced from the impediment model, yielded
an activation energy value of 170 kJ mol21, which is in
agreement with other published values for grain
boundary diffusion.136,137 Thus, based on the calculated
activation energies, the authors summarised that the
grain growth model with impediment was a good
candidate for describing the isothermal kinetics of grain
growth of nanocrystalline metals.
As discussed earlier in this article, an approach that
uses activation energy values to gain insight about the
grain growth or densification mechanisms is inherently
limited because the models contain varying parameters
which can result in mathematical flexibility that would
enable the model to fit almost any data. It is very
difficult, however, to assign physical meaning to the
parameters n and k. The fact that the grain growth of
nanoparticles during sintering may be attributed to
multiple mechanisms makes the matter even more
complicated.
In a different approach to applying the generalised
parabolic grain growth models, Liu et al.138 analysed
grain growth data in separate ‘domains’, meaning
different stages of grain growth at different sequential
19 Temperature and time evolution of the volume time periods. For each time period, a fixed value of n52
weighted average crystallite diameters of nano-Fe; the and/or 3 is used in modelling. The overall grain growth
lines represent fit with different kinetic grain growth equation is expressed as follows
models. Broken lines represent fit with the general-
G(t)2 {G02 ~
ised parabolic grain growth model; solid lines repre- 8 9
sent fit with the grain growth model with impediment; >
> k1 t t¡t1 >
>
>
> >
>
broken dotted lines represent fit with size dependent >
> k1 tzk2 (t{t1 ) t2 wtwt1 >
>
>
> >
>
impediment (Reprinted with permission from American >
> >
>
>
< >
=
Chemical Society) 
(25)
>
>  tn{1 wtwtn{2 >
>
>
> >
>
grow at tR‘. This is also termed as grain growth model >
> >
>
>
> k1 tzk2 (t{t1 )z    >
>
with impediment.131 >
> >
>
>
: >
;
In the process of grain growth, the volume fraction of zkn (t{tn{1 ) tn wtwtn{1
grain boundaries decreases; therefore, the concentration
According to this model, the rate constant k and the
of solute and impurity atoms segregated to the grain
activation energies change as grain growth progresses.
boundary is expected to increase, resulting in a grain size
Specifically, the activation energies increase as grain size
dependent retarding force on grain boundary migration.
increases. This is attributed to the increasing segregation
On this basis, Michels et al.135 proposed a grain growth
of impurities on grain boundaries which reduces grain
model with size dependent impediment.131
boundaries, hence the driving force for grain growth.
 2 2
1=2 Within each growth time period, or ‘domain’, the
G(t)~ G? { G? {G02 expð{ktÞ (24)
activation energies and rate constants k hold constant.
In a detailed study of the grain growth of nanocrystal- Compared to conventional parabolic models, this model
line Fe prepared by pulsed electrodeposition, Natter131 takes into account the effect of the changes in
applied all three grain growth models to the data of thermodynamic properties on the kinetics of grain
grain size versus time, as shown by Fig. 19. It was noted growth. The assumption of the model is that the grain
that the generalised parabolic growth model appropri- growth is fully accomplished by grain boundary diffu-
ately fits the data at low temperature, with unrealisti- sions. Direct evidence of the mechanisms of grain
cally large grain growth exponents n.10. The growth growth of nanoparticles during sintering is, however,
model with impediment yields the best fit among the still sorely needed.
three models considered.
Using the adjustable parameters contained in these Effect of pores on grain growth in late stages of sintering
models, the authors calculated the rate constants and, As in sintering of coarser particles, in the final stage of
hence, the activation energies for grain growth of sintering nanoparticles, grain growth will be affected by
nanocrystalline Fe. It was concluded that, although the remaining pores that pin the grain boundaries and
generalised parabolic model yielded a good fit with reduce the kinetic rate of grain growth. The effect of
variable n values, the calculated activation energy based pores on grain growth during the final stage of
on that fit was unreasonably high (220 kJ mol21) for conventional sintering are reviewed by Rahaman and
what is assumed to be grain boundary diffusion in Kang.47,129
nanocrystalline Fe. Therefore, it was declared by the In the final stage of sintering, isolated spherical pores
authors that the generalised parabolic model failed. are situated at the grain boundaries, and in particular, at

International Materials Reviews 2008 VOL 53 NO 6 339


Fang and Wang Densification and grain growth during sintering of nanosized particles

triple junctions. Therefore it is expected that a large grain growth. This phenomenon was further studied in
number of pores exist in a system with nanosized grains. multiple publications of Kim et al.140,141
Pores are considered to be a second phase with an It is noted, once again, that in this successful work to
inhibiting force against boundary movement. Owing to decouple grain growth from densification by exploiting
the presence of large number of pores in the system at differences in grain boundary mechanisms, the authors
the beginning of the final stage of nanosintering, the explicitly showed that at the beginning of the second
boundary is at first dragged by the pore and the rate of sintering step, the grain size increases 4–6 times larger
grain growth is slow. But as sintering proceeds, both the than the original size of the powder, which is attributed
number and the size of the pore decrease as a result of to coarsening during the first sintering step (Fig. 18).
densification. When the density reaches the specific value The following section will focus on coarsening, or in
at which the boundary is able to break away from other words, the initial grain growth of nanoparticles.
inhibition of the pore (i.e. pore/boundary separation
occurs), grain growth will accelerate dramatically. Initial grain growth – coarsening – of
Ostwald ripening during liquid phase sintering nanoparticles during early stages of sintering
The generalised grain growth law describes another (relative density ,90%)
category of grain growth which is based on Ostwald The above discussion provides evidence of an initial
ripening. The theory of Ostwald ripening was originally stage of grain growth. This part of grain growth occurs
developed for the coarsening of precipitates in two phase in the beginning of the sintering, often during heating up
materials. The mechanism of the coarsening of second when the relative density is less than 90%. In conven-
phase particles is solution and reprecipitation. During tional sintering of micrometre sized powders, the
liquid phase sintering, the solution reprecipitation contribution of the initial grain growth to the final
mechanism is responsible for increased average grain grain size is relatively minor, compared to that of the
size due to the growth of larger particles at the expense normal grain growth during late stages of sintering. For
of smaller particles. The term ‘coarsening’ is also used in nanosintering, however, the amount of the initial grain
a more general sense to describe the increase of grain growth is significant and sufficient in many cases to
sizes, as well as particle sizes, during sintering. At cause the material to lose its nanocrystalline character-
isothermal conditions, the kinetics of grain growth by istics. Further, the initial grain growth can be described
solution reprecipitation is also given by the polynomial by the generalised classic parabolic grain growth law
law as follows only if very large values of growth exponent is used,
which represent no physical processes. This implies that
G(t)n {G0n ~kt (26) the mechanism of grain growth in the initial stage may
where n is the grain growth exponent. Usually n53 for be different from that of the normal grain growth stage.
diffusion controlled processes and n52 for interface Neck formation and coarsening of contacting nanoparticles
controlled processes. Published reports on grain growth
during liquid phase sintering of nanosized tungsten To understand initial grain growth, the key issue is the
carbide particles generally have found that the equa- interaction between ultrafine particles at the start of
tion (26) is applicable with varying n values to nanosized sintering. According to classical sintering theories by
particles during isothermal sintering.139 Efforts to apply Kuczynski,100 Kingery,96 Coble142 and Johnson,143
this equation to the initial grain growth stage, however, necks will form and grow between adjacent particles,
failed to yield meaningful results.139 which are assumed to have equal diameter. Densification
is modelled as the approach of the centres of the two
Two step sintering: decoupling of grain growth from particles. In this situation, no grain growth occurs at the
densification beginning of sintering. In practical fine or ultrafine
As an example of understanding and controlling normal powder material systems, however, there are always
grain growth during nanosintering, Chen and Wang wide particle size distributions. The densification and
developed a clever approach to decouple grain growth grain growth behaviour will be markedly different.
from densificaton of nanosized particles,123 using a Figure 20 illustrates that when very fine particles are in
pressureless sintering process to fully densify nanocrys- contact, where the particle sizes are not uniform,
talline Y2O3. In a simple two step process, the compact interparticle diffusion will lead to coarsening of parti-
is briefly heated to 1310uC; the temperature is then cles, in addition to formation of the neck. Large
lowered to 1150uC and held at that temperature for an particles will grow at the expense of small particles.
extended period of time. As a result, the material can be The coarsening of particles can be understood using the
sintered to full density with minimum grain growth. If criteria shown by equation (27), which was first
the lower temperature is applied at the onset, complete expressed by Lange1 based on Kingery’s initial concept
densification would not be possible. It is reasoned, then, of pore stability.61
that suppression of the final stage grain growth is 1
achieved by exploiting the difference in kinetics between Rc ~{ (27)
cos we
the grain boundary diffusion and the grain boundary
migration. Grain growth requires grain boundary where Rc is called critical particle size ratio for boundary
migration which requires higher activation energy than migration and we is the dihedral angle relating surface
grain boundary diffusion. At a temperature that is high energy and grain boundary energy. Lange explained that
enough to overcome the energy hurdles for grain when the size ratio between two particles is larger than
boundary diffusion, but low enough to deactivate grain the critical size ratio Rc, grain boundary migration will
boundary migration, the densification will proceed via occur, resulting in grain growth. When actual size ratio
grain boundary diffusion without triggering significant is less than Rc, boundary migration will yield an increase

340 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

a just in touch without formation of interface; b when


r1/r2,Rc; c r1/r25Rc; d r1/r2.Rc
20 Linear array of two spheres of initial radii of r1 and r2
(r1.r2): After Ref. 74
a configuration when r1/r2,Rc (boundary cannot move);
in the grain boundary area and is energetically b configuration resulted from mass transport between
unfavourable. In this situation, interparticle mass particles before boundary motion; c transient configura-
tion after boundary motion where r1/r2 becomes .Rc; d
transport will happen first, in order to increase the size
final configuration either directly by mass transport or
ratio between adjacent particles. This coarsening process by combined mass transport and boundary motion
will not stop until the size ratio R5r1/r2 reaches Rc. 21 Particle configuration change after formation of di-
Then grain boundary migration will take over because hedral angle shown in Fig. 20: After Ref. 74
the condition for grain boundary migration is now
energetically satisfied. Considering the coarsening mechanisms described
The studies by Lange and Kingery aimed to explain above, the initial grain growth can, therefore, be
the stability of pores in the intermediate stage of described by a two step qualitative growth model.144
sintering. Shi further applied the critical size ratio When particles of different sizes are in contact, the first
criteria to the initial sintering of ultrafine particles.74,75 step in grain growth is coarsening due to interparticle
It was shown that the driving force for neck growth and mass transport via the growth of larger particles into
interparticle diffusion are given respectively is as follows smaller particles, which result in the increase of the
  material’s average grain size regardless of whether the
1 1
Dmn ~cs V { (28) size ratio r1/r2 is larger or smaller than Rc. During
X r
the coarsening and sintering progress, the size ratio
  between particles can increase. When the condition of
1 1
Dmc ~2cs V { (29) size ratio r1/r2.Rc is reached, grain boundary migration
r1 r2 will occur, leading to the second step of grain growth by
where Dmn and Dmc are chemical potential for neck the grain boundary migration. Figures 20 and 21
formation and mass transport between two particles, cs schematically illustrate the two step process.
is surface energy, V is atomic volume, X is radius of the The differences in grain growth kinetics at different
neck and r is radius of particles (r1 and r2 are radii of stages of sintering suggest differences in grain growth
two particles with different sizes). mechanisms, which would also be a function of specific
Equation (29) indicates that if a difference in the material systems. The mechanism for normal grain
radius of curvature exists, mass transport would take growth is widely believed to be grain boundary
place from the area of larger curvature to the area of migration. However, as indicated in the above discus-
smaller curvature. This process is related to the particle sion, the mechanism of coarsening, or initial grain
coarsening. growth, could be different depending on the specific
Considering equations (28) and (29) together, both situation with regard to relative particle size ratios and
the neck growth and coarsening, driven by the surface dihedral angles. Possible mechanisms of initial grain
tension between the particles, can take place concur- growth are discussed in the following sections.
rently. However, the magnitude of the driving force for Initial grain growth mechanisms
the two processes is different. Assuming the interface Based on conventional grain growth theories, several
energy is not considered, then |Dmn|.|Dmc|, which implies possible grain growth mechanisms for grain growth
that neck formation takes place before coarsening. during nanosintering exist, including: grain boundary
On the other hand, if the interface energy between migration, surface diffusion, solution reprecipitation
particles is considered in the analysis of the driving and coalescence. Understandably, surface diffusion is
forces as an energy barrier to neck growth, Shi74 showed expected to be a highly probable mechanism, due to the
that equation (28) becomes extremely fine particle sizes.
  It has been demonstrated that surface diffusion leads
’ 1 1
Dmn ~cs V { {cb f ðr,w,we Þ (30) to grain growth of nanosized particles. In a study of the
X r
sintering BaTiO3, Shi et al. observed that the contacting
where cb is boundary energy, w is the contact angle and particles become one particle via surface diffusion as
we is the equilibrium dihedral angle. From a thermo- shown in Fig. 22.72 Surface diffusion transported the
dynamic point of view, when w5we, the driving force for atoms from the dissolving small particle to be redepos-
the neck growth is zero. Intuitively, it is possible under ited on the surface of the larger particle. This is direct
certain conditions when w,we, driving force for coarsen- evidence of the role of surface diffusion in the
ing may equal that for neck growth. Hence, coarsening coarsening of nanoparticles at the beginning of sintering.
by interparticle mass transport may take place signifi- Although surface diffusion induced grain growth is not a
cantly prior to the achievement of the equilibrium widely recognised grain growth mechanism, it could play
dihedral angle and the beginning of grain boundary an important role in the initial grain growth. It is noted
migrations. that surface diffusion causes coarsening of larger

International Materials Reviews 2008 VOL 53 NO 6 341


Fang and Wang Densification and grain growth during sintering of nanosized particles

22 Observations of grain growth in BaTiO3 powder at different temperatures from a 940uC, b,c 950uC to d–o 960uC:
grains grow by through reduction of smaller grains and enlargement of larger ones; distance between particle centres
decreases simultaneously (Reprinted with permission from Springer Science and Business Media)

particles by consuming small particles, i.e. grain growth dramatically enhanced during the relaxation pro-
without requiring either grain boundary migration, cess,147–150 which may contribute to dynamic grain
rotation, or grain boundary diffusion. growth at the beginning of sintering. Dynamic grain
Considering that the initial grain growth during growth usually dominates during the heat-up stage and
sintering is the result of the coarsening of nanoparticles for the first few minutes after reaching a preset
due to interparticle diffusions, there are other inter- isothermal holding temperature. Therefore, rapid
particle diffusion mechanisms, other than surface diffu- dynamic grain growth accounts for the experimental
sion, that could also contribute to coarsening of observation that the first data point during isothermal
particles. In particular, there is a relaxation period for holding is several times of that of the initial grain size.
migration, redistribution and annihilation of the defects The relaxation time depends on materials, nanoparticle
due to the fact that nanoparticles are usually not at production methods and temperature.
equilibrium states and are likely to contain excess The role of grain boundary migration should also be
amounts of various defects that are created during the considered in discussing the initial grain growth during
production of nanoparticles.145,146 Owing to the non- nanosintering. As discussed earlier, for single phase
equilibrium structure of nanoparticles, diffusivity is materials at late stages of sintering when relative density

342 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

a one particle’s ‘grain’ has grown, outlined, and has distinct grain boundary; b grain boundary migrated into small particles
23 Alumina particles cluster sintered at 1200uC (Reprinted with permission from Materials Research Society)

is great than 90%, grain boundary migration is the most evidence of coalescence is, however, very difficult to
logical mechanism of grain growth as that found in bulk identify. Fang et al.111 studied the grain growth of nano-
single phase materials. Grain boundary migration has WC during sintering and found the growth of nanosized
also been observed during early stages of sintering tungsten carbide grains within aggregates via coalescence,
nanosized Al2O3.73 Figure 23 shows that when the as shown in Fig. 24. It has been speculated that when
nanosized Al2O3 particles were subjected to high nanoscale is approached, atomic mechanisms become
temperatures in a flash sintering set-up, instant grain more obvious. For example, the rotation and alignment of
growth was observed and grain boundary migration was nanosized grains may be easier than coarse grains.2
believed to be part of the process. This confirms the Kumar and Fang’s analysis of the sintering of WC–Co
analysis of the coarsening of nanoparticles that when composites suggests that the lattice shift along low energy
r1/r2 is greater than Rc, grain boundary migration will CSL grain boundaries is a viable mechanism for materials
take place. with high degree of crystallographic anisotropy.155
Coalescence is another grain growth mechanism that
is often cited to qualitatively explain rapid grain growth. Effects of agglomerates on initial grain growth
Coalescence is a term that is often loosely used to Another important factor in grain growth mechanisms
describe various phenomena. For example, coalescence during nanosintering is the role of agglomerates in grain
is sometimes used interchangeably with the term growth. Agglomerates are defined as loosely packed
‘sintering’ to describe the growth of particles during
particle synthesis and growth process.151–154 For clarity
in this article, coalescence is used strictly to describe the
increase of grain size due to the merging of two grains by
eliminating the common grain boundaries between
them. Differing from other grain growth processes,
which may also be described as the merging of two
grains, the two original grains should not demonstrate
significant change from their morphology prior to
coalescence.
The term coalescence, as defined above, describes a
unique way of grain growth, which can be accomplished
only through various diffusion mechanisms. Possible
mechanisms for coalescence include grain boundary
diffusion, dislocation climb along grain boundaries, or 24 Coalescence of two platelet shaped grains of nano-
even grain rotations. In liquid phase sintering systems, it crystalline WC–Co compact heated up to 1200uC at
is believed that the solution reprecipitation mechanism heating rate of 10uC min21 and held for 1 min: After
may also help facilitate the coalescence of grains. Direct Ref. 111

International Materials Reviews 2008 VOL 53 NO 6 343


Fang and Wang Densification and grain growth during sintering of nanosized particles

25 Schematic diagrams of a agglomerated and b aggre- 27 Microstructure of same sample as Fig. 26 at 1200uC:
gated powder (Reprinted with permission from Maney agglomerates were transformed into individual grains
Publishing)

agglomerates was characterised as ‘nucleation’ sites. A


particles forming fractals, while aggregates are particles key point of these observations, with respect to the issue
packed together in a more defined equi-axial shape. of grain growth during nanosintering, is the fact that
Figure 25 is a schematic illustration of the differences grains grow within agglomerates, probably with a
between agglomerates and aggregates. Mayo3 pointed relatively lower energy barrier and thus rapidly during
out that grain size is often related to the size of the initial stages of sintering.
agglomerates at the beginning of sintering. As Mayo
To explain the effect of agglomerates, Lange60
summarised, the larger the agglomerate size, the higher
classified the structure of a powder compact as a
the sintering temperature required to eliminate large
hierarchical structure of agglomerates, domains, and
interagglomerate pores. By contrast, the crystallite size
primary particles, as shown by Fig. 28. Defining the
has little effect on the temperature required to reach full
coordination number as the number of particles
density. The same temperatures, however, promote
surrounding the pore, Lange explained that pores within
grain growth to such an extent that the grain size can
domains have the lowest coordination number, pores
easily balloon to the agglomerate size.
between domains have higher, and pores between
Fang et al. observed a similar phenomenon. Figure 26
agglomerates have the highest coordination number.
shows an agglomerate of WC–10Co when heated to
800uC within a powder compact, while Fig. 27 shows the Figure 29 shows schematically the volume distribution
structure when the same compact is heated to 1200uC. It of the three classes of pores as a function of coordina-
can be seen that the original agglomerates, within which tion number. When N,Nc, a pore is unstable.
the WC grains are visible at 800uC, no longer exist at Otherwise, grain growth, or coarsening of particles
1200uC. Instead, the individual grains with sizes similar within agglomerates, will be necessary for elimination of
to those of the agglomerates at lower temperatures the pore and continuation of the sintering. This explains
constitute the microstructure. It is thus deduced that the the correlation between grain growth and the size of
densification and grain growth processes during nano- agglomerates.
sintering, progressed via consolidation and grain growth
within individual agglomerates, and then proceeded to
the consolidation and elimination of porosities between
agglomerates. This mechanistic process of sintering was
also observed and discussed by Petersson and Agren.55
The process that first takes place within individual

28 Schematic diagram of hierarchical structure of


26 Densification and grain growth within individual agglomerates (large circle), domains (small circle) and
aggregated particles before bulk densification primary particles (dots within small circles)

344 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

29 Schematic of pore coordination number distribution of agglomerated powder indicating three classes of pores, i.e.
those within domains, those between domains and those between agglomerates (R stands for coordination number):
After Ref. 60

Techniques for controlling grain growth while not been used. Maheshwari et al.17 found that the effect
achieving full densification of VC on grain growth during sintering of WC–Co
The difficulty of controlling grain growth while achiev- significantly inhibits the rapid grain growth during the
ing maximum density is rooted in the fact that both solid state as well as the liquid phase sintering stage.
grain growth and densification rely on thermally However, the finest grain size that has been reportedly
activated diffusion mechanisms. As discussed through- achieved using grain growth inhibitors in pressureless
out this article, densification relies on grain growth to liquid phase sintering processes is approximately 100 to
break local balance of surface and grain boundary 200 nm or larger, which is significantly larger than the
energies that mitigate the elimination of pores. Further, goal of being smaller than 100 nm. One explanation for
to achieve nanoscaled grain sizes in sintered materials, it the limited effect of grain growth inhibitors has to do
is necessary to control both the normal grain growth due with the size of agglomerates. If the mixing and
to grain boundary migration and the initial grain distribution scale of grain growth inhibitors is larger
growth, which is a crucial part of the overall grain than the original nanoscaled grain sizes and closer to the
growth process. size of the agglomerate, then the grain growth inhibitors
are effective only in that same dimensional scale.
Two step sintering technique However, more fundamental studies are needed to
As mentioned earlier, two step sintering is a proven extend their effectiveness into true nanoscale.
effective strategy for controlling normal grain growth.
Pressure assisted sintering
As shown by Fig. 18, in the two step sintering process,
the compact is briefly heated to a high normal sintering The use of pressure assisted processes is another
temperature, and then cooled to and held at a lower straightforward approach for minimising grain growth
temperature. It is necessary for the first step to be carried while achieving maximum densification. A variety of
out at a higher temperature, in order to quickly achieve pressure assisted sintering processes have been used in
a high relative density. The second isothermal step at the sintering nanosized powders, including HP, HIP, SPS
lower temperature is selected so that the densification and sinter forging. It is generally believed that the
can continue, whereas grain growth is limited because applied pressure is able not only to enhance densification
grain boundary migration is suppressed. This significant by increasing sintering driving force, assisting particle
discovery proves that it is possible to decouple grain rearrangement and promoting diffusional creep,2,165–167
growth from densification and, hence, to achieve full but also to inhibit grain growth by decreasing the
densification while retaining nanoscaled grain sizes. diffusivity and thus the grain boundary mobility.165 The
total sintering driving force when an external pressure is
Use of grain growth inhibitors applied includes both the intrinsic curvature driven
The use of grain growth inhibitors is a common method sintering stress and the applied external stress. The
for controlling grain growth during sintering. For significance of the applied pressure on sintering depends
example, with the addition of SiC to Al2O3,156 ZrO2 to on the relative magnitudes of the two components. The
b0-Al2O3,157 and Al2O3 to cubic ZrO2,158 large grain applied pressure is independent of the particle size, while
growth can be effectively prevented. The use of grain the intrinsic sintering pressure increases when the
growth inhibitor can also be found during sintering of particle size is reduced, and it can reach a very high
other materials.119,159–161 value as particle size is in nanoscale. In order to make
Grain growth inhibitors are widely used in manufac- the effect of applied pressure on densification become
turing fine and ultrafine grained cemented tungsten dominant, the applied pressure has to be larger than the
carbide (WC–Co) materials.162–164 When VC, or VC intrinsic curvature driven pressure.2,167 Therefore, a
combined with Cr2C3, is used in liquid phase sintering of threshold pressure, which is dependent on the particle
nanosized WC–Co powder, grain size after sintering is size, must be exceeded so that the applied pressure can
dramatically finer than if grain growth inhibitors had have a significant influence on sintering. The existence of

International Materials Reviews 2008 VOL 53 NO 6 345


Fang and Wang Densification and grain growth during sintering of nanosized particles

the threshold pressure had been confirmed experimen- processes during pressure induced transformation. For
tally by Skandan et al. on sintering of nanosized zirconia consolidation of nanocrystalline TiO2, pressures as high
powder.13,168 as 8 GPa was used.172 A new transformation from rutile
The logic for this approach is based on the belief that to srilankite phase starts at y5 GPa. The theory
densification can be achieved at lower than normal regarding the effect of high pressure on the sintering
sintering temperatures with the aid of pressure. The behaviour of these materials has two key points. One is
lower temperature would, of course, slow the kinetic rate that the pressure reduces the nucleation barrier for a
of grain growth. For example, Haji-Mahmood and phase transformation that is accompanied by volume
Chumbley169 investigated the consolidation of molybde- reduction, as in the case of anatase to rutile of TiO2, and
num disilicide (MoSi2) using HIP. Nanosized powders c to a phase for Al2O3. The second point is that the high
produced by mechanical attrition were hipped under hydrostatic pressure reduces the diffusion rate and, thus,
300 MPa pressure at temperatures ranging from 800 to grain growth rate.
1500uC. The grain sizes before hipping ranged from 20 The consolidation of nanosized powders by utilising
to 30 nm. After hipping the range of grain size was from the pressure induced phase transformation was also
30 to 40 nm, which is the minimum grain growth characterised as ‘transformation assisted consolidation’
considering that the relative density of the samples were (TAC).173 Transformation assisted consolidation has
from 80 to 90%. In another example, Hayashi and also been used to densify several materials, including
Etoch170 studied sintering behaviour of nanosized Fe, Al2O3 and Si3N4, as shown in Table 1.174 The key
Co, Ni and Cu metal powders under pressures ranging criteria for the suitability of this technique is that the
from 400 to 500 MPa. The nanosized powders were starting material must be a metastable phase, which
fabricated by evaporation and condensation method in transforms into the desired stable phase in a controlled
an inert gas. The starting average particle size of Fe, Co way during pressing and sintering. The combination of
and Ni were y20 nm; that of Cu was y50 nm. The increased nucleation and controlled grain growth
results showed that sintering temperatures were effec- produces fully consolidated material with nanosized
tively decreased by increasing pressure. Under 400 MPa, grain structure.
the sintering temperature for Fe, Co, Ni and Cu Sinter forging is another unique technique that has
powders were about 590, 640, 450 and 450 K respec- been used to produce fully dense materials with
tively, which were 380–620 K lower than those without nanosized grains from nanosized ceramic or metal
pressure. The minimum average grain sizes of Fe, Co, Ni powders. First of all, sinter forging, also termed powder
and Cu in the fully densified compacts after sintering forging, is a routine process that has been used to
were about 80, 210, 120 and 400 nm, which were much manufacture ferrous powder metallurgy automobile
smaller compared to those obtained by pressureless parts. In this case, green compacts are first sintered via
sintering. a standard atmospheric sintering process, and then the
Pressure induced phase transformation is a phenom- sintered compacts are placed in a forging die and forged
enon that was observed in several ceramic materials at an elevated temperature. The forging process results
during the consolidation of nanosized powders under in fully dense metal parts.
high pressure. It was found that several materials, In recent years, the term ‘sinter-forging’ has been used
including nanosized TiO2, undergo phase transforma- to describe the processing of nanoscaled powder
tion from anatase to rutile when pressure is higher than materials. However, sinter forging of nanosized powder
1 GPa during a pressure assisted consolidation pro- is somewhat different from conventional sintering
cess.165,171 The grain size of the product phase (rutile) is forging, although it shares the same traits of ‘forging’
smaller than that of the parent phase (anatase) after and distinguishes the process from conventional HP. In
consolidation. Under such high pressure, sintering this process, powders are placed in a die without lateral
temperature is as low as one-third of Tm. Further, the constraint. The die and powder are heated via techni-
transformed grain size of the rutile phase decreases as ques similar to the HP process and a uniaxial load is
temperature increases, which was explained as the applied. During sintering forging, the powder compact is
competition between nucleation and grain growth allowed to ‘upset’, i.e. bulge laterally, and is densified

Table 1 Transformation assisted consolidation processing of five material systems: After Ref. 174

Material system Starting powder Consolidated and sintered parts

TAC sintered structure Percentage theoretical density Grain size, nm


Anatase n-Rutile, n-Srilankite 95–99 10–30
TiO2
nanopowder
c-alumina n-a-alumina 95–98.2 49–73
Al2O3
nanopowder
Amorphous n-a or mixed a/b 85–96 25–40
Si3N4
nanopowder
Al2O3/13w/oTiO2 Plasma sprayed 2–4 phase: Al2O3 (a); Cannot be determined 10–30
Al2O3/40w/oTiO2 powder (mm scale) TiO2 (anatase, rutile or (theoretical density
srilankite); Al2O3 is not known)
(TiO2 (x, or b)
2–5 phase: Al2O3 (a);
TiO2 (anatase, rutile or
srilankite); Al2O3
(TiO2 (x, or b)

346 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

30 Prediction of grain size as a function of density for sinter forging at 1100uC. Higher strain rates are able to eliminate
large pores before the smaller pores enter the final stage of densification and allow grain growth. Therefore, density
is increased while grain growth is minimised (Reprinted with permission from Elsevier)

under load at specific temperature. The process is often forging. In another study on consolidation of nano-
conducted under a constant load or a constant loading TiO2, sinter forging was conducted using 20 MPa
rate. In most of the published research regarding pressure at 1573 K for 1?0 h.119 The results demonstrate
sintering of nanosized powders using sinter forging that the relative density of the samples increased from 94
process, a typical hot press is adapted for experiments. to 98% due to the use the compressive stress during
Load capacity and loading rate are thus comparable to a sinter forging.
conventional HP process. In a study on the consolida- In a study of the consolidation of nanosized metal
tion of nano-3Y-TZP (3 mol.-% yttria stabilised tetra- powders using a sinter forging process, Ma et al.176
gonal zirconia polycrystals) using sinter forging process, applied a ‘constrained sinter forging’ or ‘upset hot
the loading rate varied from 0?005 to 0?1 kN21, sintering forging’ process scheme to consolidate nanocrystalline
temperature was around 1000–1100uC. Maximum load Fe and Fe–Cu alloyed powders. A schematic of their
was not allowed to exceed 30 kN.175 For sinter forging, two step sinter forging technique is illustrated in Fig. 31.
two densification mechanisms were assumed to be Powders were prepared using mechanical milling tech-
operating: stress assisted diffusion and plastic flow niques. The initial grain sizes of mechanical milled
induced elimination of pores. Figure 30 shows the powder were 5–20 nm. In the first step, a preform was
relationship of grain size versus density at different compacted at room temperature or elevated temperature
strain rates, illustrating that finer sintered grain size can using a tungsten carbide die at a pressure of y1 GPa
be achieved by using a higher strain rate during sinter for a few to 24 h. In the second step, a compact with

a green compact is made using first die set at room temperature; b,c full density is reached in second die set using
elevated temperature and pressure
31 Schematic showing two step consolidation process, using procedure for Fe consolidation as example: After Ref. 176

International Materials Reviews 2008 VOL 53 NO 6 347


Fang and Wang Densification and grain growth during sintering of nanosized particles

32 Relative density and average grain size of aluminium oxide in compacts quenched from sintering temperatures ran-
ging from 1000 to 1400uC when a heating rate of 100uC min21 and a pressure of 50 MPa are applied. Within tempera-
ture regime I there is no densification or grain growth, whereas in regime II densification occurs, accompanied by
very limited grain growth; within regime III fast grain growth occurs in a fully dense body (Reprinted with permission
from Elsevier)

70–80% of theoretical density after the first step was plasma that may be generated between particles as a
consolidated in a second die with a larger diameter. It result of pulsed electric current. A comprehensive and
was then pressed using high pressure (0?5–1 GPa) at the critical review of the SPS process is beyond the scope of
desired temperature for 30 min to a few hours in a this paper. Readers can refer elsewhere.167,177 However,
laboratory hot press system. The final consolidation was due to the extensive body of research on the consolida-
carried out at temperatures below 600uC. Full density tion of nanosized powders that have been conducted
was reached for the sample consolidated at 500uC with using the SPS process, an examination of the effects of
an average sintered grain size of 138 nm. Consolidated SPS process characteristics on densification and grain
samples with 80 nm grain size were achieved with 99% growth of nanosized powders is necessary.
relative density. It was noted that, compared with grain Groza et al. conducted a series of studies on the
sizes in as milled powders (5–20 nm), grain growth was sintering of different materials using field assisted
present in all consolidated samples approaching full sintering techniques, or SPS.178–182 In the sintering of
density. Nonetheless, consolidated nanocrystalline Fe nanocrystalline TiN using field assisted technique179, a
with grain sizes within the typical range of nanocrystal- constant uniaxial pressure of 66 MPa was used. Relative
line (,100 nm) was achieved. The successful consolida- densities ranging from 91?5 to 94?3% were achieved for
tion with minimum grain growth during sinter forging is monolithic TiN powder. Initial average grain size of the
attributed to two aspects of applied pressure: the plastic powder was 70 nm. The average grain size after sintering
strain controlled pore closure and the stress assisted was determined to be 150–200 nm, a moderate increase
diffusion induced densification. by a factor of y2. By comparison, conventional vacuum
Spark plasma sintering is a new pressure assisted sintering conducted at 1500uC showed that the sample
sintering process that quickly gained popularity with had a final grain size of 0?5–1?0 mm with 5–7% porosity.
researchers looking for ways to consolidate materials Shen et al. studied the sintering of nanocrystalline
with nanoscale or simply very fine grain sizes, or with aluminium oxide, barium/strontium titanites, and silicon
other non-equilibrium microstructures. Spark plasma nitride using the SPS process.112,183,184 Figure 32 shows
sintering, a commercial version of several related the densification and grain growth of Al2O3 as functions
consolidation techniques, was developed primarily dur- of temperature during SPS. It was found that grain
ing 1980s and 1990s. The other variations or terms of the growth starts at a temperature that is slightly higher
technique include plasma activated sintering, field than in the densification process. Above the critical
assisted sintering, and electroconsolidation, all of which temperature, the grain growth rate, and hence the grain
share the same basic process elements as conventional size, increases rapidly, with results similar to vacuum
HP. The main difference between the SPS process and a sintering of WC–Co (Fig. 33). It demonstrates again
hot press is that rather than using an external heating that the initial grain growth during sintering of
source, a current (either or AC or pulsed) is allowed to nanosized powders is primarily temperature dependent.
pass through the electrically conducting die, and the In order to avoid excessive grain growth, the critical
sample itself if it is conductive. Further, the technique of temperature, which should vary depending on the
heating using the graphite die has been used in material, must not be exceeded.
conventional HP. Therefore, the real uniqueness of the Overall, the positive effects of SPS process for sinter-
SPS process may lie in the claimed electric spark and ing of nanosized powder reflect its unique characteristics

348 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

33 Grain size and densification of nanocrystalline WC–Co as function of sintering temperature at heating rate
10uC min21

which, as summarised by Nygren and Shen,185 include reported studies.59,186–191 In particular, Jae-Pyoung
rapid heat transfer; the application of a mechanical Ahn et al.192 showed that the grain size of sintered
pressure exceeding that used in normal HP processes; nano-SnO2 particles that have a dense green structure
the use of fast heating and cooling rates; and the use of a are dramatically smaller than that of a compact with
pulsed direct current, implying that the samples are also loose structure.
exposed to an electric field. In the authors’ opinion,
however, adequate data is not available in the literature
to support separating the effects of pressure and electric
Summary
field. Unfortunately, no data have been published that Sintering of nanosized particles is a uniquely important
compare the SPS process to the ‘hot press’ using the topic that is both scientifically and technologically
same heating rate and pressure without the pulsed challenging. From a scientific perspective, the markedly
current. different sintering behaviour of nanosized powders
compared to micrometre sized powders, raises funda-
De-agglomeration mental issues that cannot be fully comprehended based
De-agglomerating the agglomerated particles prior to on conventional theories of sintering. It is notable that
sintering is another critical strategy for minimising the the driving force for sintering nanoparticles is signifi-
initial dynamic grain growth. Based on the under- cantly higher than for micrometre sized particles. The
standing that a pore’s stability is dependent on its linear approximation used in conventional theories of
coordination numbers, a powder compact that has sintering, with regard to modelling the driving force and
uniformly distributed fine pores would have the most the diffusion equations is no longer valid. In agreement
efficient densification without relying on coarsening. with the experimental observations, the rate of sintering
An ideal scenario for minimising grain growth while, at predicted by the non-linear diffusion model is much
the same time, achieving full densification is the higher than that predicted by the conventional linear
utilisation of green powder compacts composed of diffusion model.
monosized, spherical nanoparticles without agglomer- With regard to the complex mechanisms of nanosin-
ates. The pores within such a compact would be evenly tering rapid densification of nanoparticles at relatively
distributed and would have uniform size. To achieve low temperatures cannot be explained by the activation
this type of green compact structure, Lange proposed a energy values obtained from experiments. Several
methodology in which the powders would be treated possible mechanisms discussed in this article appear to
prior to sintering according to colloidal processing contribute to the initial densification including the rapid
principles.58 First, in order to de-agglomerate the diffusion rate due to non-equilibrium defect concentra-
particles, dry powders are dispersed in fluid containing tions in nanosized powders, the indirect role of mass
a surfactant that produces interparticle repulsive transport by surface diffusion, and the possible surface
forces. After removing large particles that cannot be melting of nanoparticles, all of which contribute to the
de-agglomerated, the powder slurry is flocked by densification as well as coarsening of the nanoparticles.
changing the interparticle forces from repulsion to From a technology perspective, proof that the
attraction. As the result, the powder becomes a weak, sintering temperature drastically decreases as particle
continuous network of touching primary particles. size decreases to nanoscale represents an actionable
Colloidally treated slurries could be used directly for knowledge that can be exploited in the production of
consolidation. The effects of de-agglomeration on the engineering materials from nanosized powders. As
density of green compacts and sintering of nanosized reported in literature, the onset temperature of sintering
particles are experimentally demonstrated in many for most materials can be reduced by more than 200uC

International Materials Reviews 2008 VOL 53 NO 6 349


Fang and Wang Densification and grain growth during sintering of nanosized particles

depending on the particle size. For refractory metals 6. R. M. German: ‘Sintering theory and practice’; 1996, New York,
Wiley-Interscience.
such as tungsten, the reduction of sintering temperature
7. D. A. Porter and K. E. Easterling: ‘Phase transformation in metal
is enabling. In other words, it is usually impossible to and alloys’; 1992, London, CRC.
sinter coarse tungsten powders at temperatures below 8. C. T. Campbell, S. C. Parker and D. E. Starr: Science, 2002, 298,
1800uC. With the reduction of the particle sizes to 811–814.
20 nm, however, it is now feasible to sinter tungsten at 9. K. K. Nanda, A. Maisels, F. E. Kruis, H. Fissan and S. Stappert:
Phys. Rev. Lett., 2003, 91, 106102.
temperatures below 1200uC. 10. R. L. Penn and J. F. Banfield: Geochim. Cosmochim. Acta, 1999,
The greatest challenge for sintering nanosized pow- 63, 1549–1557.
ders is the ability to retain nanoscale grain sizes while 11. H.-Y. Lee, W. Riehemann and B. L. Mordike: J. Eur. Ceram.
achieving full densification. The available literature Soc., 1992, 10, 245–253.
12. H. Hahn: Nanostruct. Mater., 1993, 2, 251–265.
clearly demonstrates that the grain growth of nanosized
13. G. Skandan: Nanostruct. Mater., 1995, 5, 111–126.
powders is characterised by two parts of grain growth: 14. M. J. Mayo: Mater. Design, 1993, 14, 323–329.
the initial dynamic growth and the normal grain growth 15. Y. C. Zhou and M. N. Rahaman: J. Mater. Res., 1993, 8, 1680–
which is reminiscent of that in bulk materials. The initial 1696.
grain growth is the result of the coarsening of particles 16. T. Rabe and R. Waesche: Nanostruct. Mater., 1995, 6, 357–360.
17. P. Maheshwari, Z. Z. Fang and H. Y. Sohn: Int. J. Powder
via the interparticle mass transport. For nanosized
Metall., 2007, 43, 41–47.
powders, the initial grain growth causes the material to 18. M. I. Alymov, E. I. Maltina and Y. N. Stepanov: Nanostruct.
lose nanocrystalline characteristics. Therefore, if at least Mater., 1994, 4, 737–742.
part of the goals of sintering is the retention of 19. J. S. Lee and T. H. Kim: Nanostruct. Mater., 1995, 6, 691–694.
nanoscaled grain sizes, the initial grain growth must be 20. L. I. Trusov, V. N. Lapovok and V. I. Novikov: ‘Problems of
sintering metallic ultrafine powders’, in ‘Science of sintering’, (ed.
controlled and minimised. On the other hand, in the D. P. Uskokovic et al.), 185–192; 1989, New York, Plenum Press.
absence of the need to retain nanoscaled grain size, 21. O. Dominguez, Y. Champion and J. Bigot: Metall. Mater. Trans.
the initial grain growth or coarsening is one of the A, 1998, 29A, 2941–2949.
mechanisms that can be exploited to aid densification. 22. R. Malewar, K. S. Kumar, B. S. Murty, B. Sarma and S. K. Pabi:
J. Mater. Res., 2007, 22, 1200–1206.
The methods for retaining grain growth include the
23. E. Oda, K. Ameyama and S. Yamaguchi: Mater. Sci. Forum,
use of grain growth inhibitors, various high pressure hot 2006, 503–504, 573–578.
consolidation processes, and decoupling grain growth 24. H. Wang and Z. Z. Fang: ‘Study of size-dependant sintering
from densification by manipulating different diffusion behavior of tungsten powders’, Proc. 2008 Int. Conf. on
mechanisms at different temperatures. The popular use ‘Tungsten, refractory and hardmaterials VII’, Washington, DC,
USA, June 2008, Metal Powder Industries Federation, 2008,
of SPS for consolidation of nanosized powders combines 05.72–05.77.
the advantages of rapid heating rate and pressure. 25. M. Attarian Shandiz, A. Safaei, S. Sanjabi and Z. H. Barber:
However, regardless of the sintering technique used, J. Phys. Chem. Solids, 2007, 68, 1396–1399.
powder processing and green compact fabrication 26. P. Buffat and J. P. Borel: Phys. Rev. A, 1976, 13A, 2287–2298.
techniques are crucial for controlling grain growth and 27. T. Castro, R. Reifenberger, E. Choi and R. P. Andres: Phys. Rev.
B, 1990, 42B, 8548–8556.
densification. In general, it is desirable to have minimum 28. P. R. Couchman: Philos. Mag. A, 1979, 40A, 637–643.
agglomeration of particles, minimum pore size and 29. P. R. Couchman and W. A. Jesser: Nature, 1977, 269, 481–483.
uniformly distributed pores. Even though the density of 30. P. R. Couchman and C. L. Ryan: Philos. Mag. A, 1978, 37A, 369–
a green compact usually decreases as the particle size 373.
31. Q. Jiang, S. Zhang and M. Zhao: Mater. Chem. Phys., 2003, 82,
decreases, one must strive to increase green compact
225–227.
density as high as possible to achieve maximum 32. E. A. Olson, M. Y. Efremov, M. Zhang, Z. Zhang and L. H.
densification with minimum grain growth. Allen: J. Appl. Phys., 2005, 97, 034304–034301.
33. W. H. Qi and M. P. Wang: Mater. Chem. Phys., 2004, 88, 280–
Acknowledgements 284.
34. J. Ross and R. P. Andres: Surf. Sci., 1981, 106, 11–17.
The authors acknowledge the support by US 35. F. G. Shi: J. Mater. Res., 1994, 9, 1307–1313.
36. C. Solliard: Solid State Commun., 1984, 51, 947–949.
Department of Energy under Award no. DE-FC36- 37. M. Wautelet: Solid State Commun., 1990, 74, 1237–1239.
04GO14041 with cost sharing by Kennametal Inc. and 38. M. Wautelet: Eur. J. Phys., 1995, 16, 283–284.
Smith International Inc., and technical collaboration 39. M. Wautelet: Phys. Lett. A, 1998, 246, 341–342.
with Idaho National Laboratory. The authors also 40. P. Yang, Q. Jiang and X. Liu: Mater. Chem. Phys., 2007, 103, 1–4.
received support from US Army Research Laboratory 41. V. N. Troitskii, A. Z. Rakhmatullina, V. I. Berestenko and S. V.
Gurov: Poroshk. Metall., 1983, 22, 12–14.
through a subcontract by Kennametal Inc. Several 42. Q. Jiang and F. G. Shi: J. Mater. Sci. Technol., 1998, 14, 171–172.
graduate students’ thesis works are acknowledged 43. H. Ezaki, T. Nambu, M. Morinaga, M. Udaka and K. Kawasaki:
including Mr Praveen Maheshwari, Mr Xu Wang, and Int. J. Hydr. Energy, 1996, 21, 877–881.
Mr Vineet Kumar. 44. C. Herring: J. Appl. Phys., 1950, 21, 301–303.
45. G. L. Messing and M. Kumagai: Am. Ceram. Soc. Bull., 1994, 73,
88–91.
References 46. M. F. Yan and W. W. Rhodes: Mater. Sci. Eng., 1983, 61, 59–66.
47. M. N. Rahaman: ‘Ceramic processing and sintering’, 2nd edn;
1. F. F. Lange and B. J. Kellett: J. Am. Ceram. Soc., 1989, 72, 735– 2003, New York, CRC.
741. 48. J. Pan: Philos. Mag. Lett., 2004, 84, 303–310.
2. J. R. Groza: ‘Nanocrystalline powder consolidation methods’, in 49. Q. Jiang, S. H. Zhang and J. C. Li: Solid State Commun., 2004,
‘Nanostructured materials – processing, properties and potential 130, 581–584.
applications’, (ed. C. C. Koch), 1st edn, 115–178; 2002, Norwich, 50. E. Y. Gutmanas: Prog. Mater. Sci., 1990, 34, 261–366.
NY, William Andrew Publishing/Noyes. 51. E. Y. Gutmanas, L. I. Trusov and I. Gotman: Nanostruct. Mater.,
3. M. J. Mayo: Int. Mater. Rev., 1996, 41, 85–115. 1994, 4, 893–901.
4. V. Viswanathan, T. Laha, K. Balani, A. Agarwal and S. Seal: 52. M. F. Yan and W. W. Rhodes: Mater. Sci. Eng., 1983, 61, 59–66.
Mater. Sci. Eng. R, 2006, R54, 121–285. 53. E. A. Barringer, R. Brook and H. K. Bowen: ‘The sintering of
5. C. C. Koch: J. Mater. Sci., 2007, 42, 1403–1414. monodisperse TiO2’, in ‘Sintering and heterogeneous catalysis’,

350 International Materials Reviews 2008 VOL 53 NO 6


Fang and Wang Densification and grain growth during sintering of nanosized particles

(ed. G. C. Kuczynske et al. ), 1–21; 1984, New York, Plenum 97. J. K. Mackenzie and R. Shuttleworth: Proc. Phys. Soc. B, 1949,
Press. 12B, 833–852.
54. D. C. Hague: ‘Chemical precipitation, densification, and grain 98. E. A. Olevsky: Mater. Sci. Eng. R, 1998, R23, 41–100.
growth in nanocrystalline titania systems’, MSc thesis, The 99. G. W. Scherer: J. Amer. Ceram. Soc., 1977, 60, 236–239.
Pennsylvania State University, Pittsburgh, PA, 1992. 100. G. C. Kuczynski: J. Met., 1949, 1, 169–178.
55. A. Petersson and J. Agren: Acta Mater., 2005, 53, 1673–1683. 101. H. E. Exner: Rev. Powder Metall. Phys. Ceram., 1979, 1, 11–251.
56. G. Malescio: Nature Mater., 2003, 2, 501–503. 102. H. Matsubara: J. Ceram. Soc. Jpn, 2005, 113, 263–268.
57. J. N. Israelachvili: ‘Intermolecular and surface forces’; 1992, 103. L. R. Madhavrao and R. Rajagopalan: J. Mater. Res., 1989, 4,
London, Academic Press. 1251–1256.
58. W. B. Russel, D. A. Saville and W. R. Schowalter: ‘Colloidal 104. M. Braginsky, V. Tikare and E. Olevsky: Int. J. Solids Struct.,
dispersions’; 1992, Cambridge, Cambridge University Press. 2005, 42, 621–636.
59. F. F. Lange: J. Am. Ceram. Soc., 1989, 72, 3–15. 105. V. Tikare, M. Braginsky and E. A. Olevsky: J. Am. Ceram. Soc.,
60. F. F. Lange: J. Am. Ceram. Soc., 1984, 67, 83–89. 2003, 86, 49–53.
61. W. D. Kingery and B. Francois: ‘Sintering of crystalline oxide, I. 106. A. Luque, J. Aldazabal, A. Martin-Meizoso, J. M. Martinez-
Interactions between grain boundaries and pores’, in ‘Sintering Esnaola, J. G. Sevillano and R. Farr: Model. Simul. Mater. Sci.
and related phenomena’, (ed. G. C. Kuczynske et al.), 471–498; Eng., 2005, 13, 1057–1070.
1967, New York, Gordon and Breach Science Publishers. 107. T. R. Malow and C. C. Koch: ‘Grain growth of nanocrystalline
62. D. L. Johnson: J. Appl. Phys., 1969, 40, 192–200. materials – a review’, Proc. TMS Ann. Meet. on ‘Synthesis and
63. G. S. A. M. Theunissen, A. J. A. Winnubst and A. J. Burggraaf: processing of nanocrystalline powder’, Anaheim, CA, USA,
J. Eur. Ceram. Soc., 1993, 11, 315–324. February 1996, TMS, 33–44.
64. J. J. Bacmann and G. Cizeron: J. Am. Ceram. Soc., 1968, 51, 209– 108. T. R. Malow and C. C. Koch: Mater. Sci. Forum, 1996, 225–227,
212. 595–604.
65. C.-R. Li and T. B. Tang: J. Mater. Sci., 1999, 34, 3467–3470. 109. T. R. Malow and C. C. Koch: Acta Mater., 1997, 45, 2177–2186.
66. T. Ozawa: J. Therm. Anal., 1970, 2, 301–324. 110. X. Song, J. Zhang, L. Li, K. Yang and G. Liu: Acta Mater., 2006,
67. J. H. Flynn and l. A. Wall: Polym. Lett. 1966, 4, 191. 54, 5541–5550.
68. P. Vergnon, M. Astier and S. J. Teichner: Polymer preprints, 111. Z. Fang, P. Maheshwari, X. Wang, H. Y. Sohn, A. Griffo and
Division of Polymer Chemistry, American Chemical Society, R. Riley: Int. J. Refract. Met. Hard Mater., 2005, 23, 249–257.
1974, 299–307. 112. Z. J. Shen, H. Peng, J. Liu and M. Nygren: J. Eur. Ceram. Soc.,
69. K. G. Ewsuk, D. T. Ellerby and C. B. DiAntonio: J. Am. Ceram. 2004, 24, 3447–3452.
Soc., 2006, 89, 2003–2009. 113. S. Okuda, M. Kobiyama, T. Inami and S. Takamura: Scr. Mater.,
2001, 44, 2009–2012.
70. B. B. Panigrahi: Mater. Sci. Eng. A, 2007, A460–461, 7–13.
114. D. J. Chen and M. J. Mayo: Nanostruct. Mater., 1993, 2, 469.
71. V. V. Dabhade, T. R. Rama Mohan and P. Ramakrishnan:
115. W. Dickenscheid, R. Birringer, H. Gleiter, O. Kanert, B. Michel
Mater. Sci. Eng. A, 2007, A452–453, 386–394.
and B. Guenther: Solid State Commun., 1991, 79, 683–686.
72. J.-L. Shi, Y. Deguchi and Y. Sakabe: J. Mater. Sci., 2005, 40,
116. G. Hibbard, K. T. Aust, G. Palumbo and U. Erb: Scr. Mater.,
5711–5719.
2001, 44, 513–518.
73. J. E. Bonevich and L. D. Marks: Mater. Res. Soc. Symp. Proc.,
117. R. Klemm, E. Thiele, C. Holste, J. Eckert and N. Schell: Scr.
1993, 286, 3–8.
Mater., 2002, 46, 685–690.
74. J. L. Shi: J. Mater. Res., 1999, 14, 1378–1388.
118. F. Zhou, J. Lee and E. J. Lavernia: Scr. Mater., 2001, 44, 2013–
75. J. L. Shi: J. Mater. Res., 1999, 14, 1389–1397.
2017.
76. C. E. Krill, III, L. Helfen, D. Michels, H. Natter, A. Fitch,
119. D. M. Owen and A. H. Chokshi: Nanostruct. Mater., 1993, 2,
O. Masson and R. Birringer: Phys. Rev. Lett., 2001, 86, 842–845.
181–187.
77. A. H. Carim: ‘Microstructure in nanocrystalline zirconia powders
120. R. S. Averback: Z. Phys. D, 1993, 26D, 84–88.
and sintered compacts’, in ‘Nanophase materials: synthesis,
121. J. G. Li and Y. P. Ye: J. Am. Ceram. Soc., 2006, 89, 139–143.
properties, applications’, (ed. G. C. Hadjipanayis and R. W.
122. R. Vassen: CFI/Ber. DKG, 1999, 76, 19–22.
Siegel), 283–286; 1994, Dordrecht, Kluwer Academic Publishers.
123. I. W. Chen and X. H. Wang: Nature, 2000, 404, 168–171.
78. J. Y. Ying, L. F. Chi, H. Fuchs and H. Gleiter: Nanostruct. 124. R. M. German, J. Ma, X. Wang and E. Olevsky: Powder Metall.,
Mater., 1993, 3, 273. 2006, 49, 19–27.
79. H. Zhu and R. S. Averback: J. Eng. Appl. Sci. A, 1995, 204A, 96– 125. J. D. Hansen, R. P. Rusin, T. Mao-Hua and D. L. Johnson:
100. J. Am. Ceram. Soc., 1992, 75, 1129–1135.
80. H. Zhu and R. S. Averback: Mater. Manuf. Process., 1996, 11, 126. H. Su and D. L. Johnson: J. Am. Ceram. Soc., 1996, 79, 3211–
905–923. 3217.
81. Z. Xing, W. Shaoqing and Z. Caibei: J. Mater. Scie. Technol., 127. S. J. Park, J. M. Martin, J. F. Guo, J. L. Johnson and R. M.
2006, 22, 123–126. German: Metall. Mater. Trans. A, 2006, 37A, 3337–3346.
82. V. N. Koparde and P. T. Cummings: J. Phys. Chem. B, 2005, 128. H. V. Atkinson: Acta Metall., 1988, 36, 469–491.
109B, 24280–24287. 129. S.-J. L. Kang: ‘Sintering: densification, grain growth and
83. K. Tsuruta, A. Omeltchenko, R. K. Kalia and P. Vashishta: microstructure’; 2005, Burlington, MA, Butterworth-Heinemann.
Europhys. Lett., 1996, 33, 441–446. 130. J. E. Burke and D. Turnbull: Prog. Met. Phys., 1952, 3, 220–292.
84. J. S. Raut, R. B. Bhagat and K. A. Fichthorn: Nanostruct. Mater., 131. H. Natter, M. Schmelzer, M. S. Loeffler, C. E. Krill, A. Fitch and
1998, 10, 837–851. R. Hempelmann: J. Phys. Chem. B, 2000, 104B, 2467–2476.
85. C. Win-Jin, F. Te-Hua and C. Jun-Wei: Microelectr. J., 2006, 37, 132. E. A. Grey and G. T. Higgins: Acta Metall., 1973, 21, 309–321.
722–727. 133. J. E. Burke: Trans. Metall. Soc. AIME, 1949, 180, 73–91.
86. D. Hai, M. Kyoung-Sik and C. P. Wong: J. Electr. Mater., 2004, 134. J. E. Burke: Trans. Metall. Soc. AIME, 1950, 188, 1324–1328.
33, 1326–1330. 135. A. Michels, C. E. Krill, H. Ehrhardt, R. Birringer and D. T. Wu:
87. J. Pan: Int. Mater. Rev., 2003, 48, 69–85. Acta Mater., 1999, 47, 2143–2152.
88. E. A. Olevsky, V. Tikare and T. Garino: J. Am. Ceram. Soc., 136. H. Mehrer: ‘Diffusion in solid metals and alloys’; 1990, Berlin,
2006, 89, 1914–1922. New York, Springer-Verlag.
89. M. F. Ashby: Acta Metall., 1974, 22, 275–289. 137. P. Guiraldenq and P. Lacombe: Acta Metall., 1965, 13, 51–53.
90. W. Beere: Acta Metall., 1975, 23, 139–145. 138. F. Liu, G. C. Yang, H. F. Wang, Z. Chen and Y. H. Zhou:
91. R. L. Coble: J. Appl. Phys., 1961, 32, 787–792. Thermochim. Acta, 2006, 443, 212–216.
92. H. E. Exner: Powder Metall., 1980, 23, 203–209. 139. X. Wang, Z. Z. Fang and H. Y. Sohn: Int. J. Refract. Met. Hard
93. H. E. Exner and E. Arzt: ‘Sintering processes’, in ‘Physical Mater., 2008, 26, 232–241.
metallurgy’, (ed. R. W. Cahn and P. Haasen), 4th edn, 2628–2662; 140. H.-D. Kim, Y.-J. Park, B.-D. Han, M.-W. Park, W.-T. Bae,
1996, Amsterdam, Elsevier Science. Y.-W. Kim, H.-T. Lin and P. F. Becher: Scr. Mater., 2006, 54,
94. H. E. Exner and T. Kraft: ‘Review on computer simulations of 615–619.
sintering processes’, Proc. 1998 World Cong. on ‘Powder 141. Y.-I. Lee, Y.-W. Kim, M. Mitomo and D.-Y. Kim: J. Am. Ceram.
metallurgy’, Shrewsbury, UK, October 1998, European Powder Soc., 2003, 86, 1803–1805.
Metallurgy Association, 278–283. 142. R. L. Coble: J. Am. Ceram. Soc., 1958, 41, 55–62.
95. J. Frenkel: J. Phys., 1945, 9, 385–391. 143. D. L. Johnson and I. B. Cutler: J. Am. Ceram. Soc., 1963, 46, 541–
96. W. D. Kingery and M. Berg: J. Appl. Phys., 1955, 26, 1205–1212. 545.

International Materials Reviews 2008 VOL 53 NO 6 351


Fang and Wang Densification and grain growth during sintering of nanosized particles

144. C. Greskovich and K. W. Lay: J. Am. Ceram. Soc., 1972, 55, 142– 167. Z. A. Munir, U. Anselmi-Tamburini and M. Ohyanagi: J. Mater.
146. Sci., 2006, 41, 763–777.
145. R. A. Andrievski: J. Mater. Sci., 2003, 38, 1367–1375. 168. G. Skandan, H. Hahn, B. H. Kear, M. Roddy and W. R. Cannon:
146. R. Wuschum, S. Herth and U. Brossmann: ‘Diffusion in Mater. Lett., 1994, 20, 305–309.
nanocrystalline metals and alloys – a status report’, in 169. M. S. Haji-Mahmood and L. S. Chumbley: Nanostruct. Mater.,
‘Nanomaterials by severe plastic deformation’, (ed. M. 1996, 7, 95–112.
Zehetbauer and R. Z. Valiev), 755–766; 2002, Weinheim, Wiley- 170. K. Hayashi and H. Etoh: Mater. Trans. JIM, 1989, 30, 925–931.
VCH. 171. L. Shih-Chieh, K. D. Pae and W. E. Mayo: Nanostruct. Mater.,
147. L. G. Kornelyuk, A. Y. Lozovoi and I. M. Razumovskii: Diffus. 1997, 8, 645–656.
Defect Data A, 1997, 143A, (1), 1481. 172. S. C. Liao, Y. J. Chen, W. E. Mayo and B. H. Kear: Nanostruct.
148. R. Wuerschum, K. Reimann and P. Farber: Diffus. Defect Data Mater., 1999, 11, 553–557.
A, 1997, 143A, (1), 1463. 173. B. H. Kear, J. Colaizzi, W. E. Mayo and S. C. Liao: Scr. Mater.,
149. V. N. Perevezentsev: ‘Theoretical investigation of nonequilibrium 2001, 44, 2065–2068.
grain boundary diffusion properties’, in ‘Nanomaterials by severe 174. J. Colaizzi, W. E. Mayo, B. H. Kear and S.-C. Liao: Int. J. Powder
plastic deformation’, (ed. M. Zehetbauer and R. Z. Valiev), 773– Metall., 2001, 37, 45–54.
779; 2002, Weinheim, Wiley-VCH. 175. D. C. Hague and M. J. Mayo: Mater. Sci. Eng. A, 1995, A204, 83–
150. A. Nazarov: Phys. Solid State, 2003, 45, 1166–1169. 89.
151. K. Nakaso, M. Shimada, K. Okuyama and K. Deppert: J. Aerosol 176. E. Ma, D. Jia and K. T. Ramesh: ‘Nanophase and ultrafine-
Sc., 2002, 33, 1061–1074. grained powders prepared by mechanical milling: full density
152. K. E. J. Lehtinen and M. R. Zachariah: J. Aerosol Sci., 2002, 33, processing and unusual mechanical behavior’, Proc. Symp. on
357–368. ‘Powder materials: current research and industrial practices’,
153. W. Koch and S. K. Friedlander: J. Aerosol Sci., 1989, 20, 891–894. Indianapolis, IN, USA, November 2001, TMS, 257–266.
154. D. Mukherjee, C. G. Sonwane and M. R. Zachariah: J. Chem. 177. J. R. Groza and A. Zavaliangos: Rev. Adv. Mater. Sci., 2003, 5,
Phys., 2003, 119, 3391–3404. 24–33.
155. V. Kumar, Z. Fang, S. Wright and M. Nowell: Metall. Mater.
178. G. Aldica, P. Badica and J. R. Groza: J. Optoelectr. Adv. Mater.,
Trans. A, 2006, 37A, 599–607.
2007, 9, 1742–1745.
156. F. F. Lange and M. Claussen: ‘Some processing requirements for
179. J. R. Groza, J. D. Curtis and M. Kramer: J. Am. Ceram. Soc.,
transformation-toughened ceramics’, in ‘Ultrastructure processing
2000, 83, 1281–1283.
of ceramics, glasses, and composites’, (ed. L. L. Hench and D. R.
180. V. Y. Kodash, J. R. Groza, K. C. Cho, B. R. Klotz and R. J.
Ulrich), 493; 1984, New York, Wiley.
Dowding: Mater. Sci. Eng. A, 2004, A385, 367–371.
157. D. J. Green: J. Mater. Sci., 1985, 20, 2639–2646.
181. V. Sandu, G. Aldica, P. Badica, J. R. Groza and P. Nita:
158. F. J. Esper, K. H. Friese and H. Geier: ‘Mechanical, thermal, and
Superconduct. Sci. Technol., 2007, 20, 836–842.
electrical properties in the system of stabilized ZrO2(Y2O3)/alpha -
Al2O3’, in ‘Science and technology of zirconia II’, 528–536; 1984, 182. L. A. Stanciu, V. Y. Kodash and J. R. Groza: Metall. Mater.
Stuttgart, CRC. Trans. A, 2001, 32A, 2633–2638.
159. T. K. Gupta: J. Am. Ceram. Soc., 1972, 55, 276–277. 183. J. Liu, Z. Shen, M. Nygren, B. Su and T. W. Button: J. Am.
160. T.-G. Nieh and J. Wadsworth: J. Am. Ceram. Soc., 1989, 72, Ceram. Soc., 2006, 89, 2689–2694.
1469–1472. 184. Z. J. Shen, Z. Zhao, H. Peng and M. Nygren: Nature, 2002, 417,
161. R. S. Averback, H. Hahn, H. J. Hofler, J. L. Logas and T. C. 266–269.
Shen: ‘Kinetic and thermodynamic properties of nanocrystalline 185. M. Nygren and Z. Shen: Solid State Sci., 2003, 5, 125–131.
materials’, Proc. Symp. on ‘Interfaces between polymers, metals 186. O. Vasylkiv and Y. Sakka: Scr. Mater., 2001, 44, 2219–2223.
and ceramics’, San Diego, CA, USA, April 1989, Materials 187. O. Vasylkiv and Y. Sakka: J. Am. Ceram. Soc., 2001, 84, 2489–
Research Society, 3–12. 2494.
162. C. W. Morton, D. J. Wills and K. Stjernberg: Int. J. Refract. Met. 188. F. F. Lange: Curr. Opin. Solid State Mater. Sci., 1998, 3, 496–500.
Hard Mater., 2005, 23, 287–293. 189. L. Bergstrom, K. Shinozaki, H. Tomiyama and N. Mizutani:
163. R. K. Sadangi, L. E. McCandlish, B. H. Kear and P. Seegopaul: J. Am. Ceram. Soc., 1997, 80, 291–300.
Int. J. Powder Metall., 1999, 35, 27–33. 190. C. Duran, Y. Jia, Y. Hotta, K. Sato and K. Watari: J. Mater.
164. O. Seo, S. Kang and E. J. Lavernia: Mater. Trans., 2003, 44, Res., 2005, 20, 1348–1355.
2339–2345. 191. P. Bowen, C. Carry, D. Luxembourg and H. Hofmann: Powder
165. S. C. Liao, W. E. Mayo and K. D. Pae: Acta Mater., 1997, 45, Technol., 2005, 157, 100–107.
4027–4040. 192. J.-P. Ahn, M.-Y. Huh and J.-K. Park: Nanostruct. Mater., 1997,
166. B. J. Kellett and F. F. Lange: J. Am. Ceram. Soc., 1988, 71, 7–12. 8, 637–643.

352 International Materials Reviews 2008 VOL 53 NO 6

You might also like