You are on page 1of 14

Available online at www.sciencedirect.

com

Acta Materialia 56 (2008) 2119–2132


www.elsevier.com/locate/actamat

Implementation of classical nucleation and growth theories


for precipitation
M. Perez a,*, M. Dumont b, D. Acevedo-Reyes c
a
Université de Lyon, INSA Lyon, MATEIS: UMR CNRS 5510, F69621 Villeurbanne, France
b
Université Paul Cézanne Aix-Marseille III, TECSEN: UMR CNRS 6122, F13397 Marseille, France
c
CREAS, ASCOMETAL, BP 70045, 57301 Hagondange, France

Received 7 November 2007; received in revised form 19 December 2007; accepted 31 December 2007
Available online 10 March 2008

Abstract

Three ways of implementing classical nucleation and growth theories for precipitation are presented and discussed: (i) the ‘‘mean
radius approach” (particle size distribution is restricted to its mean radius and density); (ii) the ‘‘Euler-like multi-class approach” (the
particle size distribution is discretized in several size classes and its time evolution is calculated evaluating the fluxes between neighboring
classes); and (iii) the ‘‘Lagrange-like multi-class approach” (the particle size distribution is again discretized in several size classes, whose
radius time evolution are calculated). In some simple cases, the three approaches lead to similar results, but when more complex heat
treatments are involved, multi-class approaches are required. Although the Euler-like approach involves a more complex class number
management, it is more adapted to the modeling of precipitate chemistry. Some examples of implementation are presented: Cu precip-
itation in ferrite, Al3Sc precipitation in aluminum, VC and NbVC precipitation in austenite.
Ó 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Classical nucleation theory; Precipitation kinetics; Thermodynamics

1. Introduction Among all the existing models, classical nucleation and


growth theories (CNGTs) provide a compelling framework
The precipitation of a second phase is widely used to predicting the nucleation and growth rate of precipitates.
enhance the properties of metallic alloys (strength, tough- When based on accurate thermodynamical functions,
ness, creep resistance, etc.). Age hardening [1] in aluminum CNGTs are capable of describing the quantitative time
alloys and grain size control in low alloyed steels [2] are two evolution of the precipitation state, including parameters
typical examples underlining the benefits of precipitation in (iii)–(vi).
metals. Despite the extensive amount of literature on CNGTs,
Various approaches are being developed to predict the the numerical implementation of nucleation and growth
effect of thermomechanical treatments on the precipitation rate equations are seldom detailed. The aim of this paper
state. The relevant physical parameters describing these pre- is therefore to discuss and compare three classical ways
cipitates are: (i) their crystallography; (ii) their morphology; of implementing CNGTs, namely:
(iii) their chemical composition; (iv) their size distribution;
(v) their volume fraction; and (vi) their number density. (i) the ‘‘mean radius approach”: in this approach, the
particle size distribution (PSD) is restricted to its
*
mean radius and density (number of precipitates per
Corresponding author. Tel.: +33 4 72 43 80 63; fax: +33 4 72 43 85 39.
E-mail addresses: Michel.Perez@insa-lyon.fr (M. Perez), Myriam.
unit volume);
Dumont@univ-cezanne.fr (M. Dumont), D.Acevedo@ascometal.lucchini. (ii) the ‘‘Euler-like multi-class approach”: the PSD is
com (D. Acevedo-Reyes). discretized in several size classes and its time

1359-6454/$34.00 Ó 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2007.12.050
2120 M. Perez et al. / Acta Materialia 56 (2008) 2119–2132

evolution is calculated evaluating the fluxes between


neighboring classes;
(iii) the ‘‘Lagrange-like multi-class approach”:1 the PSD
is again discretized in several size classes, whose mean
radius evolution is now calculated as a function of
time.

First, CNGTs will be briefly reviewed with an emphasis


on the two basic equations describing the nucleation and
growth rates. Then, the three implementation techniques
will be detailed and compared. Finally, the benefits and
limits of each approach will be presented and illustrated
through different examples in steels and aluminum alloys.
Fig. 1. Schematic representation of the Gibbs energy changes associated
2. Classical nucleation and growth theories with precipitate formation as a function of their radius R in the classical
nucleation theory. DG is the nucleation barrier, R is the critical radius for
2.1. Nucleation stable precipitates, RkB T is the radius at which stable precipitates nucleate
and Z is the Zeldovich factor.

Classical nucleation theory (CNT) is based on the


changes in Gibbs energy DG associated with the formation vPat is the mean atomic volume within precipitates. Note
of a precipitate in a supersaturated solid solution. In the that many expressions for the condensation rate b can
case of spherical precipitates2 of radius R, DGðRÞ is given be found in the literature, none of them being supported
by by a clear justification. Russell [6] proposed an expression3
4 involving the diffusion coefficient of solute atoms D, the
DGðRÞ ¼ pR3 Dg þ 4pR2 c ð1Þ matrix mean solute atom fraction X and the lattice param-
3
eter a
where Dg is the driving force for precipitation per unit vol-
ume and c the specific interfacial energy (see Fig. 1). 4pR2 DX
b ¼ ð5Þ
CNGTs deal with (i) the rate at which stable nuclei a4
(R > R ) appear and (ii) the time evolution of the precipi- Kampmann and Wagner [7] introduced incubation time s
tate size distribution. for nucleation, leading to the following expression for the
Nearly 80 years ago, Volmer and Weber [3] proposed nucleation rate:4
the celebrated expression for the nucleation rate:     s
  dN   DG
DG ¼ N Zb exp  exp  ð6Þ
dt 
0
J VW / exp  ð2Þ nucl kT t
kBT
with s ¼ 4=ð2pb Z 2 Þ. In their approach, the nucleus size is
with DG being the energy required to form a critical nu- given by (see Fig. 1)
cleus of radius R (top of the nucleation barrier). Later,
Becker and Döring [4] and Zeldovich [5] provided impor- 1
RkB T ¼ R þ pffiffiffi ð7Þ
tant insight into the kinetic nature of this pre-factor in pZ
Eq. (2), expressing the nucleation rate as
 
DG 2.2. Growth
J BD ¼ b N 0 Z exp  ð3Þ
kB T
In binary alloys, the diffusion controlled growth rate of
where b is the condensation rate of solute atoms in a clus- spherical precipitates (molar composition X p , radius R)
ter of critical size R and N 0 is the number of nucleation site embedded in a supersaturated solid solution (mean solute
per unit volume and Z is the Zeldovich factor mole fraction in the matrix X, equilibrium solute mole frac-
rffiffiffiffiffiffiffiffi
vP c tion X i at the precipitate/matrix interface) has been pro-
Z ¼ at  ð4Þ
2pR kBT
3
In multi-component precipitates, the addition of condensation char-
acteristic times for each atomic species i gives
!1
4pR2 X 1
1
The names ‘‘Euler” and ‘‘Lagrange” come from fluid mechanics, where b ¼ :
a4 i
Di X i
lagrangian (one follows a particle) and Eulerian (fluxes are calculated at
boundaries of a fixed box) approaches are used.
2 4
The spherical shape of precipitates is an assumption that is used all The notion of ‘‘incubation time” can be, however, traced back much
along this paper. further than Kampmann’s paper (starting with Turnbull [8]).
M. Perez et al. / Acta Materialia 56 (2008) 2119–2132 2121

posed by Zener [9] under the assumption of small supersat- cipitate formation and growth in supersaturated solid solu-
uration (X 0  X i  aX p  X i ) and local equilibrium tion. As for the LS model, the MLS model does not
account for the explicit form of the size distribution (the
dR D X  X i
¼ ð8Þ ‘‘mean radius approach”).
dt R aX p  X i
More recently, Wagner and Kampmann have developed
where a ¼ vM P
at =vat is the ratio of matrix to precipitates a numerical model (the KWN model [20]) that is capable of
atomic volumes (mean volume per atom). predicting the full evolution of the size distribution. The
Interface curvature plays an important role on equilib- PSD is subdivided into intervals (size classes) and the con-
rium mole fraction X i , this is the so-called Gibbs–Thomson tinuous time evolution of the radius distribution function is
effect [10–14]. Hence, in a stoichiometric binary precipitate split into a sequence of individual decomposition steps.
of composition Ax By , radius R and matrix/precipitate sur- The KWN model has been extended and modified by
face energy c, precipitate/matrix interface equilibrium mole many researchers. Myhr and Grong [21] presented a finite
fractions X iA and X iB are modified [15] difference formulation of the KWN model by evaluating
  the fluxes between neighboring classes, the so-called
x y x y 2cðx þ yÞvPat
X iA ðRÞ X iB ðRÞ ¼ X iA ð1Þ X iB ð1Þ exp ð9Þ ‘‘Euler-like multi-class approach”. Nicolas and Deschamps
Rk B T
studied non-isothermal treatments implementing the same
leading to the growth rate equation5 approach [22].
 The success of MLS or KWN models is largely due to
dR D X  X i ðRÞ
¼ ð10Þ their high versatility, covering a wide range of applications.
dt growth R aX p  X i ðRÞ
The MLS model or ‘‘mean radius approach” has been used
to model transition from metastable to stable phases [23]
2.3. Coarsening and heterogeneous nucleation on dislocations [24]. The
‘‘Lagrange-like multi-class approach” has been imple-
Eq. (9) naturally causes small precipitates to be less sta- mented to predict the chemistry of complex carbonitrides
ble than large ones. At a constant volume fraction, small [25,16].
precipitates will then shrink (until complete dissolution) In the following section, the three ways of implementing
to the benefit of large ones, which will grow. This is the CNGTs (namely ‘‘mean radius approach” and ‘‘Lagrange/
so-called coarsening mechanism that was investigated Euler-like multi-class approach”) will be detailed and
simultaneously by Lifchitz and Slyosov [17] and Wagner discussed.
[18]. They proved that the PSD tends to a self-similar form
(Lifshitz–Slyozov–Wagner (LSW) distribution) and man- 3. Multi-class vs. mean radius approaches
aged to predict the mean radius evolution of a PSD from
the Zener growth equation (Eq. (8)) and a linearized form 3.1. The mean radius approach
of the Gibbs–Thomson equation (Eq. (9))
 In the mean radius approach, we consider the time evo-
dR 4 Xi R0 D
 ¼ p ð11Þ lution of N precipitates of mean radius R. Nucleation (Eq.
dt coars 27 aX  X i R2 (6)) and growth occur simultaneously. Note that the arrival

where R0 ¼ 2cvPat ðk B T Þ. Note that, for the case of a stoi- of dN new nuclei of size RkB T in the existing population of
chiometric binary precipitate of composition Ax By , the term precipitates of radius R modifies the growth equation as
R0 of Eq. (11) should be replaced by R0 ðx þ yÞ=y. follows:

dR D X  X i ðRÞ 1 dN  
2.4. Nucleation, growth and coarsening as concomitant ¼ þ R  R ð12Þ
processes dt growth R aX p  X i ðRÞ N dt kB T

In this approach, the coarsening stage, which is not implic-


Langer and Schwartz described droplet formation and
itly taken into account by nucleation and growth equations
growth in near critical fluids (the LS model [19]). They trea-
(Eqs. (6) and (10)) requires specific expressions for precip-
ted nucleation, growth and coarsening as concomitant pro-
itate number N and mean radius R evolution.
cesses. They used steady-state nucleation theory and a
The LSW theory (Eq. (11)) describes the mean radius
linearized version of the Gibbs–Thomson equation (Eq.
evolution in the asymptotic limit (t ! 1). Within this
(9)), thus giving the time evolution of droplet density and
limit, the solution mole fraction X is considered to be equal
mean radius (‘‘Mean radius approach”).
to the equilibrium solute mole fraction X i . If the volume
Later on, Kampmann and Wagner modified the Langer
fraction of precipitates is small, the mass balance of solute
and Schwartz approach (the MLS model [7]) to treat pre-
atoms (total mole fraction X 0 ) is given by
 
0 4 3 p 4 3
5
In multi-component precipitates, e.g. NbN, the growth process X ¼ a pR NX þ X 1  a pR N ð13Þ
involves one growth equation per atomic species (see Ref. [16]). 3 3
2122 M. Perez et al. / Acta Materialia 56 (2008) 2119–2132

Differentiation of Eq. (13) and use of LSW theory (Eq. and the precipitate volume fraction fT
(11)) gives the precipitate number density evolution during
4 X
the pure coarsening stage: fT ¼ p N i R3i ð21Þ
 3 i
dN  4 Xi
 ¼ p The mass balance is given by
dt coars 27 aX  X i
   
R0 D R0 X i 3 X 0 ¼ afT X p þ X ð1  afT Þ ð22Þ
 3  N  3N ð14Þ
R RðX p  X i Þ 4pR3
To model the evolution of the PSD, two strategies can be
In order to go continuously from the growth stage to the employed
coarsening stage, Deschamps and Bréchet [26] introduced
a coarsening fraction fcoars to weight the pure growth equa-  the exchange of particles between adjacent and predeter-
tion dR=dtjgrowth (Eq. (12)) and the pure coarsening equa- mined size classes is calculated using a modified version
tion dR=dtjcoars (Eq. (11)) of the growth equation (Eq. (10)) (”Euler-like multi-
  class approach”);
dR dR dR  the radius evolution of each class (that contains a fixed
¼ fcoars  þ ð1  fcoars Þ  ð15Þ
dt dt coars dt growth number of precipitates) is calculated using directly the
growth equation (Eq. (10)) (”Lagrange-like multi-class
They proposed an expression for the coarsening fraction approach”);
  
R
fcoars ¼ 1  erf 4   1 ð16Þ In the following, these two approaches will be presented
R in detail and compared with the ‘‘mean-radius approach”.
The last expression is well adapted to precipitation, but
fails to describe precipitate dissolution (R < R ). In this pa- 3.3. The Euler-like approach
per, we choose a more general description (very close to the
one proposed by Maugis [24]) In the Euler-like approach, fluxes of particles at bound-
aries of fixed classes are evaluated. Such an approach is
 2 inspired by the KWN model.
R
fcoars ¼ 1  1000   1 0:99R < R < 1:01R ð17Þ
R
3.3.1. Nucleation and growth
Similarly, the number of precipitates is given by A nucleation stage is performed using the nucleation
(    rate equation of the CNT (Eq. (6)). The number of new
dN
¼ fcoars dN  if dN  > dN 
dt dt coars dt coars dt nucl nuclei DN is given by
   ð18Þ
dN
¼ dN  if dN  < dN 
dt dt nucl dt coars dt nucl dN
DN ¼ Dt ð23Þ
The set of two equations for mean size R and number den- dt
sity N evolution (Eqs. (15) and (18)) can be integrated where Dt is the time step. The size class in which to put
numerically using a Runge–Kutta scheme with adaptive these new nuclei is the one that contains the nucleation ra-
time step. The solute fraction X, given by the mass balance dius RkB T .
(Eq. (13)) is re-evaluated at each time step. The growth stage is computed using the control volume
approach, as proposed by Myhr and Grong [21]. In this
3.2. Multi-class approaches: Euler vs. Lagrange approach, the number density evolution of a size class i is
evaluated from the number densities and growth rates of
To describe the time evolution of the whole PSD, the neighboring size classes (indexed i  1 and i þ 1), whose
continuous size distribution is discretized into a large num- growth rates are ti1 ¼ dR=dtji1 and tiþ1 ¼ dR=dtjiþ1 . This
ber of size classes. A size class is defined by its radius R and case is illustrated in Fig. 2. The calculation of the particle
the number of precipitates, the size of which ranges number density within each class is highly dependent on
between R  DR=2 and R þ DR=2. the sign of the growth rate defined at the limits of the class
From this definition, we can compute the total number i: dR=dtji1 and dR=dtji . A mass balance gives the new pop-
of precipitates N per unit volume ulation of class i
X
N¼ Ni ð19Þ Dt
i N i;tþDt ¼ N i;t þ ti1;t ½SIGNðti1;t ÞN i1;t
DR
the mean radius of the distribution R Dt
P þ SIGNðti1;t ÞN i;t   ti;t ½SIGNðti;t ÞN i;t
Ri N i DR
R¼ P i
ð20Þ þ SIGNðti;t ÞN iþ1;t  ð24Þ
Ni
i where the SIGN(x) function is 1 if x > 0 and 0 otherwise.
M. Perez et al. / Acta Materialia 56 (2008) 2119–2132 2123

volume (often less than 1 particle/m3) will create another


size class (imax þ 2), and so on. Thus the number of classes
would increase without limit, which would be impractical.
To avoid this type of behavior inherent to Euler-like
approach, a new size class can be created exclusively when
the corresponding particle density is higher than a thresh-
old (typically 1 particle/m3). Otherwise the particles are left
in their previous size class.

3.4. The Lagrange-like approach

Fig. 2. Growth of precipitates in the ‘‘Euler-like approach”. At each time The Lagrange-like approach is inspired by the ‘‘Multi-
step, fluxes between each neighboring classes are calculated. Préci” model developed by Maugis and Gouné [25]. In this
approach, the time evolution of each size class radius is cal-
3.3.2. Initial conditions culated, the size class population remaining constant.
Computation is started with a system of size classes, the Hence, at each time step:
size and positions of which are predetermined. These clas-
ses can be either empty (no precipitates, N i ¼ 0) or repre-  a new class is created. Its radius is RkB T and its popula-
sent a given PSD (e.g. coming from an experiment), for tion is

which we want to model the time evolution. dN 
DN ¼ Dt ð26Þ
dt  nucl
3.3.3. Time and boundary conditions
An auto-adapting time increment is calculated at each Nucleation rate is given by Eq. (6) and nucleation radius by
step of the process to avoid any size class moving more Eq. (7). If DN is too low (typically lower than 1010 times
than one size element. Dt is calculated in such a manner the total precipitate number density), no class is created.
that the class presenting the highest velocity dR=dtmax will  the new radius of all existing classes is calculated

move DR=2 dR
Rðt þ DtÞ ¼ RðtÞ þ  Dt ð27Þ
DR 1 dt growth
Dt ¼ ð25Þ
2 MAXjdR=dtji j
The growth rate is given by Eq. (10). Fig. 3 details how
It should be noted that in practice dR=dtjmax is experienced nucleation and growth stages are managed within the La-
by the smallest size class of the distribution, which is grange-like approach.
dissolving. For very small precipitates, infinite velocities
can be calculated due to limitations of the assumptions
involved in the growth rate equation. A cut-off value of 3.4.1. Adaptive time step
the size distribution is then chosen to avoid the interfacial In this approach, the time step drives the precision at
concentration becoming higher than the precipitate which differential equations giving class numbers N i and
composition. radii Ri are solved. Too short time step would lead to
At the opposite end of the distribution, if the largest size time-consuming calculations and unnecessarily large num-
class is in the growth regime, i.e. dR=dtjimax is positive, then ber of classes. Inversely, too long time step could lead to
at each time step a new size class will be created, even if numerical instabilities (e.g. increasing amplitude oscilla-
dR=dtjimax is very small. At the next step this new size class, tions around the solution) or even aberrations (e.g. nega-
imax þ 1, which may contain very few particles per unit tive concentrations).

Fig. 3. Nucleation-growth sequence within a Lagrange-like approach. At each time step, a new class is created and all existing classes grow.
2124 M. Perez et al. / Acta Materialia 56 (2008) 2119–2132

Thus, the time step needs to be optimized during the cal- classes smaller than the critical size (Ri < R ) shrink, until
culation. A logarithmic increment of the time step is first they completely disappear. To keep an accurate description
performed (Dt 1:1Dt). This new Dt is validated under of the PSD, a minimum number of classes N min should exist
two conditions, calculated at t þ Dt: (typically N min ¼ 500). Therefore, the maximum distance
between two classes is fixed at
 all solute concentrations should range between 0 and 1; Rmax  Rmin
 the critical radius R (which is the most sensitive variable Ri  Ri1 > 2 ¼ DRmax ð28Þ
N min
of the model) may not vary more than 1% between each
time step. If the distance between two neighboring classes i  1 and i
is higher than DRmax , a new class is artificially created.
If any of these conditions is not fulfilled, we return to the Insertion of a new class between classes i  1 and i would
previous time step (this necessitates to keep in memory the inevitably change either the mean radius R or the trans-
state of the system at t  Dt) and make a new attempt with formed fraction fT . Indeed, rigorously keeping these two
a smaller time step (typically Dt Dt=2). This last step parameters would lead to a system of incompatible equa-
could be repeated as many time as necessary. tions. Inserting a new class must lead to preservation of
Fig. 4 illustrates the adaptive time step scheme in the the distribution density Di ¼ N i =ðRiþ1  Ri Þ rather than
case of simultaneous precipitation of stable cementite and the distribution N i .6
metastable e carbides [23]. We observe that instabilities If a new class is added between classes i  1 and i, all
such as dissolution of precipitates lead to a strong decrease classes with indices greater or equal to i are shifted up by
in the time step (see Fig. 4). one unit (see Fig. 5). The new class index is then i and
the old class i becomes i þ 1. In the following, all indices
3.4.2. Class number management are new indices.
In this approach, unlike in the Euler-like approach, the It is then necessary to set new radii R0i1 , R0i , R0iþ1 and
number of classes decreases during the coarsening stage: populations N 0i1 , N 0i , N 0iþ1 of classes i  1, i and i þ 1 (six
unknowns). We then suppose

 radii of old classes are kept unchanged


R0i1 ¼ Ri1 and R0iþ1 ¼ Riþ1 ð29Þ
 the number densities of old classes are unchanged
D0i1 ¼ Di1 and D0iþ1 ¼ Diþ1 ð30Þ
 the new class is put right in the middle of the two old
ones
R0i ¼ ðRi1 þ Riþ1 Þ=2 ð31Þ
 continuity of the number density is provided
Fig. 4. Time-step evolution in the case of simultaneous precipitation of a D0i  D0i1 D0iþ1  D0i1
metastable (e carbide) and a stable (Fe3C) phase in iron [23]. At ¼ 0 ð32Þ
approximately 1500 s, the time step is reduced to accurately model the R0i  R0i1 Riþ1  R0i1
complete dissolution of the metastable phase.

The above conditions lead to the new number of precip-


itates in each class
N iþ1 Riþ1  Ri1 N i1 0
N 0i1 ¼ 0:5N i1 N 0i ¼ þ N iþ1
4 Riþ2  Riþ1 4
i+1
¼ N 0iþ1 ð33Þ
The class insertion procedure cannot keep the precipitate
volume fraction unchanged, which leads to a violation of
the mass balance equation. To overcome this difficulty,
after the transformation ½Ri ; N i  ! ½R0i ; N 0i , a new transfor-
mation ½R0i ; N 0i  ! ½R00i ; N 00i  has to be done in order to get a

6
Note that if the time step had been half of its actual size during the
nucleation stage, we would have had another class right inbetween classes
Fig. 5. Addition of a new class in the Lagrange-like approach. The new i  1 and i (assuming a constant dR=dt), proving that only PSD density Di
class should not modify the distribution density N i =ðRiþ1  Ri Þ. has a physical meaning.
M. Perez et al. / Acta Materialia 56 (2008) 2119–2132 2125

Fig. 6. Comparison between particle size distributions (N i ¼ f ðRÞ) and particle size distribution density (Di ¼ N i =ðRiþ1  Ri Þ) after many insertions of new
classes.

final precipitate volume fraction fT00 , which is rigorously Table 1


equal to the initial precipitate volume fraction fT . Parameters used to compare the three types of approaches (mean radius,
Euler-like and Lagrange-like) in the simple case of a nucleation-growth-
X X coarsening sequence
fT00 ¼ N 00 R300 ¼
i i i i
N i R3i ð34Þ
Xp X0 Xi D (m2/s) c vM;P
at a (m) T
This transformation consists in a rescaling of precipitate (J/m2) (m3) (°C)
numbers N i of all classes without changing their radii 1 0.06 0.01 5  1020 0.13 1:6  1029 4:04  1010 160

R00i ¼ R0i and N 00i ¼ nN 0i ð35Þ

This naturally leads to It is even surprising that the mean radius approach man-
fT ages to describe so accurately the transition from nucle-
n¼ ð36Þ ation-growth regimes to the coarsening regime. However,
fT0
we will see in the next section that multi-class approaches
Fig. 6 shows an example of particle size distribution N i ðRi Þ are required in some more complex cases.
and distribution density Di ðRi Þ. It can be observed that, un-
like particle number density N i , particle size distribution 3.5.2. When the mean radius approach breaks down
density Di ¼ N i =ðRiþ1  Ri Þ remains smooth and un- A multi-class model definitely contains more informa-
changed after many class insertions. tion than a model dealing with mean quantities. However,
the last section showed that the mean radius approach was
3.5. Comparison between three approaches as accurate in a simple case.
In some more complex situations, the mean radius
To compare the three approaches presented above, we approach breaks down. For example, in the case of a rever-
studied (i) simple nucleation-growth-coarsening kinetics sion treatment (i.e. a pre-existing PSD is submitted to high
involving homogeneous nucleation and (ii) a more complex temperature treatment) a dissolution stage has been
situation where we compare two different PSDs with the observed to occur without significant changes in the mean
same mean radius and precipitate transformed fraction. radius in Al–Zn–Mg alloys [22]. The mean radius approach
is then expected to be unable to reproduce such a behavior
3.5.1. A simple nucleation-growth-coarsening kinetics since a decrease in the precipitate transformed fraction
In this example, the driving force for nucleation is given would involve a decrease in the mean radius.
by an ideal solution model. We start with a supersaturated Such kinds of complex treatments have been performed
solid solution containing no precipitate. Parameters used to on a model Fe–0.2 wt.% V–0.5 wt.% C alloy, for which ini-
compare the three types of approaches (mean radius, tial PSDs resulting from long-term treatment at 800 °C are
Euler-like and Lagrange-like) are recalled in Table 1. submitted to thermal aging at 920 °C. In this section, we
Fig. 7 compares evolutions of precipitate radii, precipi- compare multi-class to mean radius approaches during a
tate number density, solute concentration and precipitate reversion treatment, starting from two different initial
volume fraction for a nucleation-growth-coarsening PSDs, which have the same mean radius and precipitate
sequence. The agreement between the three ways of imple- volume fraction. Therefore, the mean radius approach will
menting CNGTs is remarkable. not make any difference between these two PSDs.
2126 M. Perez et al. / Acta Materialia 56 (2008) 2119–2132

Fig. 7. Time evolution of radii, precipitate number density, solute concentration and precipitate volume fraction for multi-class and mean radius
approaches. The agreement between the three ways of implementing CNGTs is remarkable.

Fig. 8a shows two initial distributions of VC precipitates sion; the equilibrium carbon fraction at the precipitate/
in austenite resulting from a precipitation treatment of 10 matrix interface X iC is then equal to the carbon solute frac-
days at 800 °C.7 The first distribution is a classical LSW tion X C . The solubility product of vanadium carbides is
distribution resulting from a simulation involving nucle- given by: K VC ðRÞ ¼ K VC expðR0 =RÞ ¼ X iV X iC expðR0 =RÞ
ation-growth-coarsening stages (see parameters used for and the growth rate of the precipitates is
this model alloy in Table 2; more details can be found in
dR DV X V  X iV
Ref. [29]). The second distribution is an experimental one ¼ ð37Þ
dt R a=2  X iV
resulting from a fine scanning electron microscope charac-
terization in the transmission mode (see Ref. [27]). Note Fig. 8b–d shows that the two initial PSDs give quite differ-
that the experimental distribution has a log-normal shape ent results in terms of the mean radius (b), precipitate vol-
in contradiction with the simulated one, which has an ume fraction (c) and precipitate number density (d)
LSW shape.8 In spite of their different shapes, these two evolutions, especially for intermediate times. These differ-
distributions exhibit exactly the same mean radius and pre- ences can be perfectly explained by the different shapes of
cipitate transformed fraction. the PSDs and cannot be reproduced with the mean radius
From these two distributions, an aging treatment at model. Indeed, the PSD that contains more small precipi-
920 °C is modeled with the three approaches: Lagrange- tates will experience more precipitate dissolution during
like, Euler-like and mean radius (see parameters in Table the beginning of the coarsening stage, thus exhibiting a lar-
2). The growth of precipitates is limited by vanadium diffu- ger mean radius. From this example, it appears clearly that
the mean radius approach is unable to predict a dissolution
at constant mean radius as observed experimentally during
7
Although all observed vanadium carbide precipitates have been reversion treatments.
positively identified as monoclinic ordered V6C5 carbides [28], they will It can be seen in Fig. 8b–d that both multi-class
be considered, for the sake of simplicity, as stoichiometric carbides VC in
this paper. approaches lead to identical results, whereas the mean radius
8
This shape difference, reported many times in the literature, has, to our approach is in accordance with the two other approaches
knowledge, never been explained. only for the LSW-type initial precipitate distribution.
M. Perez et al. / Acta Materialia 56 (2008) 2119–2132 2127

Fig. 8. (a) Precipitate size distribution resulting from a 10-day precipitation treatment at 800 °C: LSW distribution and experimental distribution [27]
fitted by a log-normal function. These two distributions have the same mean radius AND precipitate volume fraction. Mean radius (b), transformed
fraction (c) and number density (d) evolutions during an aging treatment at 920 °C for each the two distributions presented in (a): comparison between the
three approaches. Both multi-class approaches are superimposed and fit remarkably well with experiments [27], whereas mean radius approach is
unappropriate to model non-LSW precipitate size distribution.

Table 2
in perfect agreement with both multi-class approaches pro-
Parameters used to model the VC precipitate size distribution evolution of
an FeVC model alloy viding that the initial PSD is an experimental log-normal
distribution.
log10 K VC [30] DV (m2/s)[29] c (J/m2)[29] vPat (m3)
5
Such a reversion treatment has also been performed at
9500
þ 3:42 10 expð 240;000
RT Þ 0.4 9:9  1030
T 950 °C. The initial LSW distribution leads to a complete
dissolution of all precipitates, whereas the biggest precipi-
Indeed, equations that were used to build the LSW theory tates remain for the case of an initial log-normal
for coarsening (Eqs. (11) and (14)), on which the mean radius distribution.
approach is based, suppose an LSW-type PSD. Therefore,
the mean radius approach is unappropriate to model non- 4. Applications
LSW precipitate size distribution evolution.
From Fig. 8b–d, finally, it can be noticed that experi- 4.1. The mean radius approach: copper precipitation in iron
mental measurements (radii measured by scanning trans-
mission electron microscopy (STEM) or scanning Precipitation of copper in iron has been extensively stud-
electron microscopy (SEM), and the transformed fraction ied in the last 50 years because copper induces precipitation
with electrolytic extraction – for details see Ref. [31]) are hardening after rapid cooling and tempering.
2128 M. Perez et al. / Acta Materialia 56 (2008) 2119–2132

Despite the relative complexity of the precipitation Table 3


sequence (body-centered cubic (bcc) ? 9R ? face-centered Parameters used to model copper precipitation in iron
cubic (fcc)), the binary iron–copper system is quite often Parameter Value Reference
used as a model alloy to validate precipitation models, lX p
1
due to the spherical nature of precipitates and the fact that X0 0.0125
6 4

copper precipitates are reported to be pure copper even in Xi 5:7710


T2
 1:5810
T þ 9:89
[33]
D (m2/s) 6  108 exp  166;400 [34]
the earlier stages of formation. cbcc (J/m2) 0.25
RT
[20]
The LS model [32] and the N model [20] have been cfcc (J/m2) 0.5 [20]
recently used on the binary Fe–Cu system to treat nucle- vP;M
at (m3) 1:2  1029
ation, growth and coarsening as concomitant processes.
In this section, we will apply the mean radius approach distribution of precipitates as obtained in age-hardenable
to model the precipitation of copper in iron. alloys has generally formed on homogeneously distributed
To take into account the loss of coherency on the precip- defects such as vacancy-rich clusters or dislocations loops,
itation kinetics, a size-dependent interfacial energy is intro- so nucleation is heterogeneous in nature. Dilute Al–Sc bin-
duced. As the bcc to fcc transition is usually observed for ary alloys have proven to be a model system where pure
particles larger than RT ¼ 3 nm [20,33], the following form homogeneous nucleation takes place [37]. Therefore the
for c has been chosen (see the inset of Fig. 9) precipitation kinetics has been extensively studied in this
  system. Detailed experimental characterization [36,37] pro-
R
c ¼ min cbcc þ ðcfcc  cbcc Þ; cfcc ð38Þ vided essential data for modeling which gave satisfactory
RT
results on the basis of classical nucleation and growth the-
All parameters used in this simulation are recalled in Table ory [37–39].
3. The object of the following example using the Euler-like
Fig. 9 shows the time evolution of precipitate mean approach applied to the nucleation of Al3Sc precipitates in
radius and number density. The mean radius approach is an Al–0.12 at.% Sc is to give an illustration on how a class
compared with SANS [20,35] and SAXS experiments [33] model can help in understanding phenomenological pro-
performed on binary Fe–Cu alloys at 500 and 600 °C. cesses. Novotny and Ardell [36] observed (see Fig. 10) that
The agreement is quite good in view of the simplicity of this Al3Sc precipitates were highly resistant to coarsening in
approach: homogeneous nucleation, mean field approach, this alloy. Even after long aging times (107 s) at 350 °C,
no impingement of precipitate diffusion fields (all precipi- precipitates do not undergo coarsening. Moreover, they
tates lying within a solid solution, which solute fraction is observed that the particle size distribution (PSD) was in
updated at each time step). good agreement with the LSW prediction at early stages
of aging, but the PSD became narrower than that predicted
4.2. The Euler-like approach: homogeneous nucleation of by the LSW theory for longer aging times. They conclude
Al3Sc in a dilute Al–Sc binary alloy that this behavior was difficult to reconcile with the classi-
cal theory for nucleation, growth and coarsening.
Pure homogeneous nucleation is rarely observed during The Euler-like approach was applied to this system
aging of commercial alloys. Even an homogeneous using classical nucleation and growth theories. Parameters

Fig. 9. Kinetic evolution of the precipitate radius (left) and number density (right) as predicted by the mean radius model. Solid points refer to
experimental data from nuclear and magnetic SANS experiments on Fe–1.38 at.% Cu (from Wagner et al. [20] and Mathon et al. [35]), and SAXS
experiments on Fe–1.25 at.% Cu (from Perez et al. [33]). Inset: to take into account the transition from the bcc coherent to the fcc incoherent phase, the
precipitates interfacial energy varies from cbcc to cfcc depending on precipitate radius.
M. Perez et al. / Acta Materialia 56 (2008) 2119–2132 2129

Fig. 10. Nucleation and growth of Al3Sc precipitates in an Al–0.12 at.% Sc alloy during an isothermal treatment at 350 °C. Evolution of the average
particle radius, as compared with Novotny and Ardell data [36], of the total number of precipitates and of the particle size distribution as predicted by the
Euler-like multi-class model. Parameters used in the simulation: c ¼ 0:127 J m2, D ¼ 5:31  104 expð173; 000=RT Þ m2 s1, X Sc ð1Þ ¼ 3:884  105 at,
V Pm ¼ 4:166  105 m3 mol1.

used in the model were extracted from the literature and nucleation growth to a pure coarsening process is charac-
are listed in the caption of Fig. 10. Only the interfacial terized by a constant mean radius stage. This plateau is
energy was slightly adapted to fit at best the experimental generally predicted by modeling [26] but rarely observed
data for two compositions (Al–0.12 and 0.18 at.% Sc) experimentally since nucleation in commercial alloys is
[36]. The retained value of 0.127 J m2 is reliable with the generally finished after a few minutes of aging. This stage
values available in the literature [40,36,37,39] varying from proceeds until the decrease in the total number of precipi-
0.04 to 0.2 J m2. The results obtained by modeling are in tates starts, i.e. the set-up of the so-called coarsening stage.
very good agreement with experimental data as shown on From this example it appears that the criterion generally
Fig. 10. In particular, modeling predicts a retardation of used to determine the beginning of the coarsening stage
the coarsening stage in the dilute Al–0.12 at.% Sc alloy that based on the R=R ratio is not satisfactory in this case. In
starts for very long aging times: 107 s at 350 °C (last exper- fact, R closely approaches R early in the intermediate
imental data point). In this alloy, the nucleation stage char- stage, i.e. 105 s, whereas the coarsening stage only begins
acterized by the increase in the total particle number takes after 106 s. The criterion R  R is a necessary condition
place over a large time scale, up to 104 s, because of the that ensures the dissolution of the smallest particles that
small supersaturation in this system. The transition from will supply solute atoms to the largest ones but appears
2130 M. Perez et al. / Acta Materialia 56 (2008) 2119–2132

not to be sufficient. A second condition to satisfy is that the vVC þ ð1  vÞNbC $ Vv Nb1v C ð39Þ
smallest particles have to be entirely dissolved to make the
Growth and Gibbs–Thomson equations presented in Eq.
total number of particles decrease. In a global view of the
(10) are written for (V, Nb)C precipitates as follows:
PSD (see Fig. 10), it means that the PSD, relatively sym- 8
X V X iV
metrical at the transition from nucleation to growth stage >
> dR
¼ DRV av=2X
>
> dt i
(105 s), has to enlarge and generate small particle size >
>
>
V

> dR DNb X Nb X iNb


classes before coarsening can begin. This takes quite a long >
> ¼
>
> dt R að1vÞ=2X iNb
time in the dilute Al–0.12 at.% Sc system because at the end <
X C X iC
of the nucleation-growth stage even the smallest precipi-
dR
dt
¼ DRC a=2X i ð40Þ
>
> C

tates are relatively large (10 nm), and also involves a >
> 2
ð1vÞ XC
>Xi Xi
> ¼ vK VC exp VNb
exp RR0
small dissolution rate. These two factors make them long >
> V C kB T
>
> 2 C
to dissolve completely. >
>
: X i X i ¼ ð1  vÞK NbC exp v XVNb exp R0
Nb C kB T R

4.3. The Lagrange-like approach: (V, Nb)C dissolution in The first three equations of system (40) drive the diffusion
austenite kinetics of V, Nb and C solute atoms. The last two equa-
tions define the solubility products of both VC and NbC
Microalloying elements such as Nb, V and Ti are com- carbides forming the complex Vv Nb1v C precipitate. The
monly used in special steels because of the strong effect they expðR0 =RÞ stands for the Gibbs–Thomson effect and XCVNb
can have on mechanical properties. A small addition of is the interaction parameter of the regular solution model.
these elements leads to precipitation of carbides, nitrides For the sake of simplicity, the v dependence of a and R0
and carbonitrides that can have a hardening effect or can have been neglected. System (40) is solved numerically
control grain size growth. Enhancement of mechanical using a Newton–Raphson scheme at each time step for
properties depends on the stability and evolution of the all precipitate classes.
precipitation state during heat treatments. Parameters used for the simulation are presented in
In the specific case of vanadium and niobium carbides, Tables 2 and 4. Initial PSD has been characterized combin-
the perfect miscibility of both carbides and the existence ing TEM and SEM techniques [29].
of a miscibility gap have been confirmed experimentally Fig. 11 shows the evolution of Vv Nb1v C precipitate
[41,42]. It has also been shown that the chemical composi- mean radius and chemical composition during isothermal
tion of the mixed carbides formed depends on the size of holding at 950 and 1200 °C. Lagrange-like gives a good
the particles [43–45]. Hence, to model the evolution of description of both mean radius and chemical composition.
the precipitation state, only a PSD-evolution model can
be used.
A Lagrange-like approach has been used to model the Table 4
evolution of the precipitation state during isothermal heat Parameters used to model the (Nb, V)C precipitate size distribution
evolution of a FeNbVC model alloy
treatments in austenite of a FeCVNb model alloy. More
details of the experimental characterization of the precipi- log10 K NbC DNb DC c XCVNb
(m2/s) (m2/s) (J/ (J/
tation state will be published elsewhere. (V, Nb)C carbides m2) mol)
are supposed to be stoichiometric for simplicity, and are
7900
þ 0:13 8:3  105 exp 266;500 105 exp 137;500 0.5 30,090
modeled as a regular solution of VC and NbC: T RT RT
χ

Fig. 11. Kinetic evolution of the precipitate radius (left) and chemical composition (right) as predicted by the Lagrange-like approach. Solid points refer to
experimental data from SEM, TEM (precipitate radius) and electrolytic extraction (precipitate chemestry).
M. Perez et al. / Acta Materialia 56 (2008) 2119–2132 2131

Just as for the case of Nb(C, N) precipitation in ferrite (5) To accurately describe the coarsening regime with a
[16], the Lagrange-like approach is particularly well adapted Lagrange-like approach, new classes should be added
to model the chemistry of precipitates. The reason is that all in the PSD. It is then necessary to keep the distribu-
precipitates that appear at the same time shall have the same tion density Di ¼ N i =ðRiþ1  Ri Þ unchanged and to
chemistry throughout all the precipitation treatment. rescale all class populations to keep the precipitate
volume fraction unchanged.
5. Conclusions (6) The Euler-like approach has the advantage of not
requiring a class number management as classes are
(1) On a simple nucleation-growth-coarsening sequence, fixed and delimited, but the optimal class size and
mean radius and multi-class approaches are equiva- position actually depends on the thermal treatment.
lent and give same results. (7) The Lagrange-like approach offers better potentiali-
(2) In many relatively simple cases (precipitation ties to model the chemistry of precipitates (e.g. pre-
sequence and thermal treatments), the mean radius cipitation of VxNb1xC or NbCxNy).
approach is faster and as accurate as more complex
multi-class approaches.
(3) The mean radius approach is unappropriate to model Acknowledgement
non-LSW precipitate size distribution evolution.
(4) A multi-class model is essential to take into account This work was partially supported by Ascometal
more complex treatments or shapes of the initial CREAS. Thanks are due to P. Maugis, F. Perrard and
PSD: two different initial PSDs with the same (i) A. Deschamps for fruitful discussions.
mean radius and (ii) precipitate volume fraction will
actually lead to different mean radius evolution dur- References
ing a simulated aging treatment.
[1] Porter DA, Easterling KE. Phase transformation in metals and
alloys. London: Chapman and Hall; 1992. 514p.
[2] Gladman T. The physical metallurgy of microalloyed steels. Lon-
don: The Institute of Materials; 2002.
Table A.1 [3] Volmer M, Weber A. Keimbildung in übersättigen gebilden. Z Phys
Table of symbols used in this paper Chem 1926;119:277–301.
Symbol Signification [4] Becker R, Döring W. Kinetische behandlung der keimbildung in
übersättigten dämpfen. Annalen der Physik 1935;416:719–52.
a Ratio of matrix to precipitates mean atomic volume a ¼ vM P
at =vat [5] Zeldovich YB. On the theory of new phase formation: cavitation.
b Condensation rate of solute atoms in a cluster of critical size R
Acta Physicochim USSR 1943;18:1–22.
Dg Driving force for nucleation per unit volume
[6] Russell KC. Phase transformations. American Society for Metals;
c Precipitate/matrix interfacial energy
1968. p. 219–68 (Chapter Nucleation in solids).
DG Gibbs energy change associated with the formation of a
[7] Kampmann R, Wagner R. Decomposition of alloys: the early
precipitate
stages. Pergamon Press; 1984. p. 91–103 (Chapter Kinetics of
DG Nucleation barrier
precipitation in metastable binary alloys – theory and application to
Dt Time step
Cu–1.9 at% Ti and Ni–14 at% Al).
s Incubation time s ¼ 4=ð2pb Z 2 Þ
[8] Turnbull D. Transient nucleation. Trans AIME 1948;175:774–83.
ti Growth rate of class i
[9] Zener C. Theory of growth of spherical precipitates from solid
a Matrix lattice parameter
solution. J App Phys 1949;20:950–3.
D Diffusion coefficient of solute atoms
[10] Gibbs JW. Collected works. Green and Co.; 1928 (Chapter On the
Di Precipitate size distribution density Di ¼ N i =ðRiþ1  Ri Þ
equilibrium of heterogeneous substances).
fcoars Coarsening fraction
[11] Thomson J. Theoretical considerations on the effect of pressure in
fT Precipitate transformed fraction
lowering the freezing point of water. Trans R Soc Edin
kB Boltzmann constant
1849;16:575–80.
N Precipitate number density
[12] Thomson J. On crystallization and liquefaction, as influenced by
Ni Precipitate number density of class i
stresses tending to changes of form of crystals. Proc R Soc
N0 Number of nucleation site per unit volume
1862;11:473–81.
R Precipitate radius
[13] Thomson W. On the equilibrium of vapour at a curved surface of
R Precipitate mean radius
liquid. Phil Mag 1871;42:448–52.
R0 Capillary radius R0 ¼ 2cvPat =ðk B T Þ
[14] Thomson JJ. Application of dynamics to physics and chemistry. Lon-
Ri Radius of class i
don: Macmillan; 1888.
R Critical radius for stable precipitates
[15] Perez M. Gibbs–Thomson effect in phase transformations. Scripta
RkB T Nucleation radius
Mater 2005;52:709–12.
T Temperature
[16] Perez M, Courtois E, Acevedo D, Epicier T, Maugis P. Precipitation
vM;P
at Mean atomic volume within the matrix or precipitates
of niobium carbonitrides in ferrite: chemical composition measure-
X Solute mole fraction
ments and thermodynamical modelling. Phil Mag Lett
X0 Alloying element composition in mole fraction
2007;87:645–56.
Xi Equilibrium solute mole fraction at the precipitate/matrix
[17] Lifchitz IM, Slyosov VV. The kinetics of precipitation from super-
interface
saturated solid solution. J Phys Chem Solids 1961;19(1/2):35–50.
Xp Molar composition of precipitates
[18] Wagner C. Theorie der alterung von niedershlSgen durch umlsen
Z Zeldovich factor
(ostwald-reifung). Z Electrochem 1961;65:581.
2132 M. Perez et al. / Acta Materialia 56 (2008) 2119–2132

[19] Langer JS, Schwartz AJ. Kinetics of nuleation in near-critical fluids. [33] Perez M, Perrard F, Massardier V, Kleber X, Deschamps A, De
Phys Rev A 1980;21:948–58. Monestrol H, et al. Small angle X-ray scattering and tomographic
[20] Wagner R, Kampmann R. Materials science and technology: a atom probe. Phil Mag 2005;85(20):2197–210.
comprehensive treatment. John Wiley & Sons Inc.; 1991. p. 213–302 [34] Deschamps A, Nicolas M, Perrard F, Bley F. Non-isothermal
(Chapter Homogeneous second phase precipitation). precipitation kinetics in metallic alloys. In: Solid to solid phase
[21] Myhr OR, Grong O. Modelling of non-isothermal transformations in transformations in inorganic materials. TMS; 2005.
alloys containing a particle distribution. Acta Mater 2000;48: [35] Mathon MH, Barbu A, Dunstetter F, Maury F, Lorenzelli N, de
1605–15. Novion CH. Experimental study and modelling of copper precipita-
[22] Nicolas M, Deschamps A. Characterisation and modelling of tion under electron irradiation in dilute FeCu binary alloys. J Nucl
precipitate evolution in an Al–Zn–Mg alloy during non-isothermal Mater 1997;245:224–37.
heat treatments. Acta Mater 2003:6077–94. [36] Novotny GM, Ardell AJ. Precipitation of Al3Sc in binary AlSc alloys.
[23] Perez M, Deschamps A. Microscopic modelling of simultaneous two- Mater Sci Eng A 2001;A318:144–54.
phase precipitation: application to carbide precipitation in low- [37] Hyland RW. Homogeneous nucleation kinetics of Al3Sc in a dilute
carbon steels. Mater Sci Eng A 2003;A360:214–9. Al–Sc alloy. Metall Trans A 1992;23A:1947–55.
[24] Perrard F, Deschamps A, Maugis P. Modelling the precipitation of [38] Robson JD. Modelling the overlap of nucleation, growth and
NbC on dislocations in a-Fe. Acta Mater 2007;55:1255–66. coarsening during precipitation. Acta Mater 2004;52:4669–76.
[25] Maugis P, Gouné M. Kinetics of vanadium carbonitride precipitation [39] Clouet E, Nastar M, Sigli C. Nucleation of Al3Zr and Al3Sc in
in steel: a computer model. Acta Mater 2005;53:3359–67. aluminum alloys: from kinetic Monte-Carlo simulations for classical
[26] Deschamps A, Bréchet Y. Influence of predeformation on an ageing theory. Phys Rev B 2004;69:064109.
in an Al–Zn–Mg alloy – II Modeling of precipitation kinetics and [40] Jo H-H, Fujikawa S-I. Kinetics of precipitation in Al–Sc alloys and
yield stress. Acta Mater 1999;47(1):293–305. low temperature solid solubility of scandium in aluminium studied by
[27] Acevedo-Reyes D, Perez M, Epicier T, Bogner A, Epicier T, Verdu C. electrical resistivity measurements. Mater Sci Eng A
Characterisation of precipitates size distribution: validation of low 1993;A171:151–61.
voltage STEM. J Micros 2008. [41] Inoue K, Ishikawa N, Ohnuma I, Ohtani H, Ishida K. Calculation for
[28] Epicier T, Acevedo D, Perez M. Crystallographic structure of phase equilibria between austenite and (Nb, Ti, V)(C, N) in microal-
vanadium carbide precipitates in a model Fe–C–V steel. Phil Mag loyed steels. ISIJ Int 2001;41:175–82.
2008;88:31–45. [42] Kieffer R, Nowotny H, Neckel A, Ettmayer P, Usner L. Zur
[29] Acevedo D. Evolution de l’état de précipitation au cours de entmischung von kubischen mehrstoffcarbiden. Monatshefte fnr
l’austénitisation d’aciers microalliés au vanadium et au niobium. Chemie 1986;99:1020–7.
PhD thesis, INSA Lyon; 2007. [43] Speer JG, Michael JR, Hansen SS. Carbonitride precipitation in
[30] Narita K. Physical chemistry of the groups IVa (Ti, Zr), Va niobium/vanadium microalloyed steels. Metall Trans A
(V, Nb, Ta) and the rare earth elements in steel. Trans ISIJ 1987;18A:211–22.
1975;15:145–52. [44] Thompson SW, Krauss G. Precipitation and fine structure in
[31] Acevedo D, Perez M, Epicier T, Kozeschnik E, Perrard F, Sourmail medium-carbon vanadium and vanadium–niobium microalloyed
T. Kinetics of precipitation and dissolution in model FeCV and steels. Metall Trans A 1989;20A:2279–88.
FeCNbV alloy, and a microalloyed ultra high strength spring steel. [45] Touryan LA, Thompson SW, Matlock DK, Krauss G. Microstruc-
In: New Dev. Metall. Appl. High Strength Steels. Bueno Aires: TMS; ture and properties of medium-carbon steels microalloyed with
2008. vanadium and niobium. In: Finn M, editor. Proceedings of interna-
[32] Zhang C, Enomoto M, Yamashita T, Sano N. Cu precipitation in a tional symposium on microalloyed bar and forging steels. Montreal,
prestrained Fe1.5 wt pct Cu alloy during isothermal aging. Metall Quebec: Canadian Institute of Mining and Metallurgy; 1990. p.
Trans A 2004;35A:1262–72. 63–83.

You might also like