Вы находитесь на странице: 1из 11

Materials Science and Engineering A 373 (2004) 10–20

Quantification of phase transformation kinetics


of 18 wt.% Ni C250 maraging steel
Z. Guo, W. Sha∗ , D. Li
School of Civil Engineering, The Queen’s University of Belfast, Belfast, Northern Ireland BT7 1NN, UK

Received 25 March 2003; received in revised form 5 June 2003

Abstract

Research on the kinetics of precipitate formation and austenite reversion in maraging steels has received great attention due to their
importance to the properties of the alloys. In the present work, the phase transformation kinetics of an 18 wt.% Ni maraging grade C250 was
investigated using differential scanning calorimetry (DSC). Based on the calorimetric data obtained at different heating rates, the kinetics
of precipitate formation and austenite reversion was modeled using the Johnson–Mehl–Avrami (JMA) theory. Good agreement between
the calculated transformed fraction and the experimental data from DSC measurement is demonstrated for both phase transformations. The
derived JMA kinetic parameters were then used to simulate the phase transformation kinetics during isothermal ageing. Comparison between
the calculated kinetics and experimental data from literature was carried out. Mechanisms for precipitate formation and austenite reversion
were discussed based on the value of the kinetic parameters.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Precipitation kinetics; Austenite; Johnson–Mehl–Avrami theory; Differential scanning calorimetry; Steel

1. Introduction steels such as magnetic properties [15] and stress corrosion


cracking resistance [16].
Over the past 40 years, a generic class of ultra-high Research on precipitate formation and austenite reversion
strength maraging steels has been developed mainly for air- has received great attention due to their importance to the
craft, aerospace and tooling applications. Maraging refers to properties of maraging steels. In contrast to extensive exper-
the ageing of martensite, a martensite that is easily obtained imental studies [2–14], there has been limited work on phys-
at normal cooling rates due to the high nickel content [1]. ical modeling of the precipitate formation kinetics [17,18]
The ultra-high strength of maraging steels is due to pre- and austenite reversion kinetics [16] in maraging steels. With
cipitation, usually of intermetallic compounds, during the the development of differential scanning calorimetry (DSC),
ageing process [2–9]. Recent studies revealed that strength- it has been possible to trace the phase transformation ki-
ening of 18 wt.% Ni (18Ni) maraging steels results from netics under a controlled temperature program. Therefore,
the combined presence of Ni3 Ti and Fe2 Mo [4] or Fe7 Mo6 DSC has been widely used to study phase transformation
[5] precipitates. The formation of Ni3 Ti takes place rapidly kinetics in various systems [19–25]. In the present work,
due to the fast diffusion of Ti atoms [4,5]. When good the results of DSC study have been used to model the ki-
toughness is required, maraging steels have to be treated netics of precipitation formation and austenite reversion in
to overaged condition to allow the formation of a certain a commercial 18Ni grade-250 (C250) alloy. The kinetics
amount of austenite that remains stable at room temperature was simulated based on the classical Johnson–Mehl–Avrami
[10–14]. Austenite also affects other properties of maraging (JMA) phase transformation theory. The kinetic parameters
derived from continuous heating experiments were used to
calculate the phase transformation kinetics during isothermal
∗ Corresponding author. Tel.: +44-28-9097-4017; holding. To the authors’ best knowledge, no similar attempt
fax: +44-28-9066-3754. on quantification of the phase transformation kinetics in
E-mail address: w.sha@qub.ac.uk (W. Sha). maraging steels has been made before. Comparison between

0921-5093/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2004.01.040
Z. Guo et al. / Materials Science and Engineering A 373 (2004) 10–20 11

Table 1
Chemical composition of the C250 steel
C Co Mo Ni Ti Al Si Cr Mn Zr P S B Ca Fe

wt.% 0.06 8.29 5.01 18.06 0.47 0.098 0.02 0.01 0.02 0.009 0.003 0.0007 0.003 0.0002 Balance
at.% 0.29 8.10 3.01 17.71 0.56 0.21 0.04 0.01 0.02 0.006 0.006 0.0013 0.016 0.0003 Balance

the calculated results and experimental study was carried and the reference material are subjected to a controlled tem-
out. perature program. The main heat flow from the furnace to
the samples transfers through a disk with good thermal con-
duction (the material of the disk in DSC-404 is platinum).
2. Experimental procedures Both the sample to be tested and the reference material are
put in alumina pans and the two pans are positioned on
The 18Ni grade-250 (C250) alloy studied was provided this platinum disk where the temperature sensors are inte-
by Allvac Ltd., UK. Its composition in both wt.% and at.% grated. The DSC machine was calibrated using standard In,
is given in Table 1. In this paper, without mentioned, alloy Zn, Al, Ag, Au, and Ni samples before commencement of
composition will always be in wt.%. The material was vac- measurements. Samples were cut into small pieces of size
uum induction melted plus vacuum arc remelted (VIM/VAR) 3.5 mm × 3.5 mm × 1.0 mm for DSC analysis. All experi-
to a 500 mm ingot, homogenized at 1250 ◦ C for 12 h, forged ments were performed in a He flow of 30 ml/min after evac-
from 1100 ◦ C to 280 mm diameter. Blocks of size about uating the heating chamber to avoid sample oxidation. An
12 mm × 12 mm × 12 mm were cut from the material in this empty crucible was used as a reference. The specimen was
condition. They were then held at 816 ◦ C for 30 min fol- heated up from 20 to 820 ◦ C at a heating rate 5, 10, 20,
lowed by water quenching. Samples for hardness measure- 30, 40, or 50 K/min, and then cooled to room temperature.
ment were aged at 427, 482, and 538 ◦ C for 1, 3, 10, and Phase transformations taking place in the specimen during
50 h. A layer of 1 mm was removed from the sample sur- heating were recorded.
face to avoid the possible influence of oxygen when speci-
mens were characterized or tested. The hardness measure-
ment was carried out on a Mitutoya HM-124 machine, with 3. Quantification of transformation kinetics
2 kg working load.
A Netzsch DSC-404 apparatus was used to record the Fig. 1 shows the recorded DSC curves of C250 alloy
phase transformations taking place during heating. This ma- at different heating rates. There exist various opinions on
chine is a heat-flux DSC with disk-type measuring system. the interpretation of phase transformations taking place dur-
The so-called heat flux-differential scanning calorimetry is ing continuous heating [26–34], which are demonstrated in
a technique by which the heat input into a sample as com- Table 2, using the DSC curve heating at 50 K/min as an ex-
pared to a reference material is measured while the sample ample. The exothermic process at 310–510 ◦ C may corre-

50 K/min
40 K/min
0.8
30 K/min
20 K/min
Heat flow (mW/mg)

10 K/min
0.6 05 K/min

0.4

0.2

0
300 350 400 450 500 550 600 650 700 750 800
Temperature (°C)

Fig. 1. Results from DSC runs for the C250 alloy at different heating rates.
12 Z. Guo et al. / Materials Science and Engineering A 373 (2004) 10–20

Table 2
Interpretation of DSC results, using the DSC curve obtained at a 50 K/min heating rate as an example
Zone DSC feature Temperature span (◦ C) Possible transformations
I Exothermic 310–510 Recovery of martensite [33]
Formation of carbide precipitates (minor hardening) [33]
Formation of coherent precipitation zones [31]
II Exothermic 535–603 Formation of the main strengthening precipitates [27–29,31–34]
III Endothermic 635–730 Austenite reversion [27–29,33,34]
Formation of retained austenite∗ by diffusion [31]
IV Endothermic 730–800 Martensite to austenite by shear [31]
Resolution of precipitates or recrystallisation [32]
∗ Retained austenite is the austenite not transformed after cooling. This term is used here to denote the part of reverted austenite formed during heating

that retains the austenite structure during the following cooling to room temperature. The other part of reverted austenite will transform back to martensite.

spond to the recovery of martensite [33], the precipitation jor) martensite to austenite reversion reaction. It should be
of carbides [33], or the precipitation of coherent precipi- noted that no Curie point is observed in the DSC curves at
tate zones [31]. These processes contributed only limited different heating rates [32]. The major reactions of austenite
hardening effects [33]. The main precipitation process, i.e. reversion and precipitate formation will be studied in detail
the formation of intermetallic precipitates, is the exother- below.
mic reaction between 535 and 603 ◦ C [27–29,31–34]. The
endothermic peak at 635–730 ◦ C corresponds to the marten- 3.1. Austenite reversion kinetics
site to austenite transformation [27–29,33,34] or the forma-
tion of austenite that can be retained at room temperature The peak parameters as well as the measured transfor-
[31]. The exothermic peak above 730 ◦ C may result from mation enthalpies of the austenite reversion peak at dif-
the martensite to austenite transformation by shear mecha- ferent heating rates are given in Table 3. The usual ac-
nism [31], or the resolution of precipitates [32], or recrys- curacy in determining enthalpy value is within the range
tallisation [32]. Grain growth may also result in exothermic of ±5%, which is slightly affected by the determination
effects [35]. As can be seen that the peak at 535–603 ◦ C is of start and complete temperatures of the transformation.
from the precipitation process has little controversy. How- It is believed that large errors are associated with the en-
ever, there are different opinions with the austenite rever- thalpy values at 5 and 10 K/min, which is due to the dif-
sion peak. Therefore, thermodynamic calculation was car- ficulty in determining the start and end temperatures of
ried out to estimate the reversion completion temperature, the DSC peaks. The enthalpy values at other heating rates
which agrees well with the termination temperature of the are close to the latent heat of the ferrite α → γ trans-
endothermic peak III (Section 3.1.2). The transformation en- formation in pure iron, a confirmation that the peaks cor-
thalpy corresponding to this peak is also in agreement with respond to the austenite reversion transformation. In or-
the value for ferrite α → γ transformation in pure iron, der to demonstrate the influence of heating rate clearly,
i.e. 16 J/g [36] (Section 3.1). Recent study revealed that the the peaks in Fig. 1 have been shifted downwards and base
martensite to austenite reversion during continuous heating line corrected so that all the onset and termination points
could take place through shear, or diffusion, or a combina- are on the same horizontal line, while preserving the shape
tion of these two mechanisms in Fe–Ni alloys [37,38]. The of the curve. The processed austenite reversion peaks in
transformation temperatures were not affected by heating the DSC curves at different heating rates are displayed in
rate in shear mode [28,38]. The increase of transformation Fig. 2.
temperatures with increasing heating rate as observed in the
present work is a strong indication of diffusion mechanism 3.1.1. Determination of transformed fraction
for austenite reversion. Based on these pieces of evidence, Usually for DSC experiments, the degree of transforma-
the transformation at peak III is considered to be the (ma- tion at any given time is taken equal to the fraction of heat

Table 3
Parameters of the DSC peak for austenite reversion at various heating rates
Heating rate (K/min) Start temperature (◦ C) Peak temperature (◦ C) End temperature (◦ C) Enthalpy (J/g)
5 610 636 680 7.4
10 615 649 700 10.2
20 620 662 710 13.5
30 625 671 718 14.6
40 630 677 725 15.5
50 635 683 730 13.8
Z. Guo et al. / Materials Science and Engineering A 373 (2004) 10–20 13

0.3

50 K/min
0.25 40 K/min
30 K/min
20 K/min
Heat flow (mW/mg)

0.2 10 K/min
05 K/min

0.15

0.1

0.05

0
600 620 640 660 680 700 720 740
Temperature (°C)

Fig. 2. Processed DSC peaks for austenite reversion at different heating rates.

released/absorbed [21,22]: constant and K0 and βf are constants. The activation en-
t ergy E is given by the gradient of the straight line obtained
t Hdt by linear regression on the plot of ln(Tf 2 /φ) versus 1/(RTf ).
f(t) =  tSE (1)
tS Hdt Using the peak temperature at different heating rates, the
activation energy of the austenite reversion process is esti-
where H is the heat flow measured, f(t) is the degree of trans- mated as E = 342 kJ/mol, Fig. 5. It should be noted that
formation at any given time (t) and tS and tE are transforma- Kissinger-type methods were originally derived for homoge-
tion start and end times, respectively. The denominator in neous reactions, but it can also be applied to heterogeneous
Eq. (1) is the total enthalpy for a transformation calculated reactions [41].
from the corresponding peak in the DSC curves, Fig. 2. One
should be aware that the total enthalpy is the integration of 3.1.3. Calculation of transformation kinetic parameters
heat flow H over a period of time, instead of temperature. The kinetics of an isothermal transformation is usually
The degree of transformation as a function of time or tem- expressed by the modified Johnson–Mehl–Avrami equation
perature can be calculated from the DSC signal. The calcu- [42–44]:
lated results for different heating rates are plotted in Fig. 3.
The calculated transformed fraction curves trace the course f
= 1 − exp(−ktn ) (3)
of the martensite α → γ reversion of the C250 alloy. feq

3.1.2. Equilibrium austenite fraction and determination of where f is the volume fraction of the product phase, feq the
activation energy equilibrium fraction at the studied temperature, k the reac-
ThermoCalc package was used to estimate the equilibrium tion rate constant, and n is the Avrami index. Within the
austenite phase fraction as a function of temperature, Fig. 4. temperature range of the DSC peaks recorded at various
All the possible phases, bcc, austenite and various types of heating rates (610–730 ◦ C), the equilibrium austenite frac-
precipitates were included in the calculation. The calculated tion increases from about 73–100%. In the following text, f
temperature for austenite reversion completion is 674 ◦ C, in is termed as transformation fraction that varies in the range
good agreement with the value 680 ◦ C measured at heating of 0–feq , and f/feq is the degree of transformation, ranging
rate 5 K/min. from 0 to 1. The reaction rate constant k is temperature de-
The activation energy of the austenite reversion process pendent and can be expressed as:
can be estimated based on the DSC results at different heat-  
−E
ing rates using a modified Kissinger method [39,40]: k(T) = k0 exp (4)
RT
Tf2 E E
ln = + ln + ln βf (2) where k0 is the pre-exponential factor, E the activation en-
φ RTf RK0
ergy, R the gas constant, and T is the temperature. The ac-
where Tf is the characteristic temperature for a given pro- tivation energy of the austenitisation process was estimated
cess, φ the heating rate, E the activation energy, R the gas as 342 kJ/mol in the previous section.
14 Z. Guo et al. / Materials Science and Engineering A 373 (2004) 10–20

1
50 K/min
40 K/min
0.8 30 K/min
20 K/min
Transformed fraction

10 K/min
0.6 05 K/min

0.4

0.2

0
600 620 640 660 680 700 720 740

(a) Temperature (°C)

0.8
Transformed fraction

0.6

0.4

0.2

0
0 100 200 300 400 500 600 700 800 900
(b) Time (s)

Fig. 3. Calculated transformed austenite fraction as a function of (a) temperature and (b) time at different heating rates.

100 Since Eq. (3) is used to depict isothermal process, the


next step is to use finite isothermal steps to represent the
non-isothermal process based on the additivity principle
Austenite fraction (%)

80 from Scheil [45] and Christian [46]. A computer program


was written to optimize the kinetic parameters n and k0 so
that the best fit between the experimental and the calculated
60
transformed fraction versus T/t curves (derived from DSC
data) can be achieved. The Avrami index n was assumed to
40 be temperature independent, which is valid for most trans-
formations in an appreciable temperature range [47]. These
two parameters were determined as n = 0.97 and k0 =
20
1.73 × 1017 . Using the derived n, k0 , and E, the transformed
350 400 450 500 550 600 650 700
fraction as a function of time at various heating rates can
Temperature (°C) be calculated, and results are shown in Fig. 6. Experimental
Fig. 4. Equilibrium volume fraction of austenite as a function of tem-
values from DSC were also plotted in Fig. 6 for comparison.
perature. All the possible phases, bcc, austenite and various types of As can be seen, a good fit between the experimental and the
precipitates were included in the calculation. calculated results is achieved.
Z. Guo et al. / Materials Science and Engineering A 373 (2004) 10–20 15

13 2.2
Precipitate fraction (vol.%)

12 2.1
ln(Tf2/φ)

11
2.0

10
1.9

9
0.124 0.126 0.128 0.13 0.132 0.134 1.8
1000/(RTf) (mol/kJ) 475 500 525 550 575 600 625
Temperature (°C)
Fig. 5. Determination of the activation energy for austenite reversion
from the peak temperatures at different heating rates. Activation energy Fig. 7. Equilibrium volume fraction of Ni3 Ti precipitate as a function of
in kJ/mol is the slope of the straight line. ageing temperature. The phases included in calculation are bcc and Ni3 Ti
only. The equilibrium precipitate fraction is considered as a constant
3.2. Precipitate formation kinetics 2.0 vol.% in the kinetics calculations.

The peak parameters as well as the measured transfor- were close, which indicates that about the same amount of
mation enthalpies of the main precipitate formation peak at precipitate forms regardless of heating rate. Following sim-
different heating rates are given in Table 4. The enthalpy ilar procedures as described in Section 3.1.2, the activation
values of the precipitation process at different heating rates energy of the precipitation process was determined to be

0.9

0.8

0.7

0.6
Transformed fraction

0.5
50 K/min (Exp.)
50 K/min (Cal.)
0.4
40 K/min (Exp.)
40 K/min (Cal.)
30 K/min (Exp.)
0.3
30 K/min (Cal.)
20 K/min (Exp.)
20 K/min (Cal.)
0.2
10 K/min (Exp.)
10 K/min (Cal.)
05 K/min (Exp.)
0.1
05 K/min (Cal.)

0
0 100 200 300 400 500 600 700 800 900 1000
Time (s)

Fig. 6. Comparison between the calculated austenite transformation fraction and DSC experimental results.
16 Z. Guo et al. / Materials Science and Engineering A 373 (2004) 10–20

Table 4
Parameters of the DSC peak for precipitation at various heating rates∗
Heating rate (K/min) Start temperature (◦ C)∗∗ Peak temperature (◦ C)∗∗ End temperature (◦ C) Enthalpy (J/g)
10 505 522 570 −1.8
20 520 536 585 −1.5
30 525 547 590 −1.1
40 530 556 596 −1.7
50 535 562 603 −1.6

The DSC curve at 5 K/min does not allow an accurate assessment of the exothermic peak.
∗∗
These peak temperatures were determined from raw DSC curves shown in Fig. 1, before the base line correction leading to Fig. 8. Thus, the
temperatures listed here differ from what can be read off Fig. 8. However, base line corrected peak temperatures read off Fig. 8 gave the same activation
energy value.

205.1 kJ/mol, with the coefficient of linear regression of the trace the precipitation kinetics [41,48]:
plot of ln(Tf 2 /φ) versus 1/(RTf ) being 0.993. f
ThermoCalc package was used to estimate the equilib- = 1 − exp(−θ n ) (5)
feq
rium precipitate fraction as a function of temperature. The
type of precipitate was assumed to be Ni3 Ti only. There- The difference between Eqs. (3) and (5) is the introduction
fore, the phases included in calculation are bcc and Ni3 Ti. of a new state variable θ [20,21,39]:
Although austenite is a stable phase within the temperature  t
range for ageing treatments, it does not form until the alloy θ= k(T)dt (6)
is seriously overaged. Therefore, when the fraction of pre- 0
cipitate was calculated, the austenite phase was not included where k(T) has been given by Eq. (4). The above model
in the calculation. As can be seen from Fig. 7, the equilib- has been successfully applied to describe the kinetics of
rium precipitate fraction decreases with increasing temper- phase transitions or transformations during continuous heat-
ature. However, such temperature dependence is very weak. ing such as crystallization, recrystallization, and precipita-
In the following calculations on precipitation kinetics, pre- tion process [20,21,39]. A computer program was written
cipitation process is assumed to be complete, reaching the to optimize the kinetic parameters n and k0 so that the best
equilibrium precipitate fraction, which is assumed to be tem- fit between the experimental and the calculated transformed
perature independent, 2.0 vol.%. The processed DSC curves fraction versus T/t curves (derived from DSC data) can
of the precipitation process at different heating rates are dis- be achieved. The JMA parameters for precipitation process
played in Fig. 8. were then derived as n = 1.46 and k0 = 1.60 × 1011 . Using
Since the equilibrium precipitate fraction can be consid- the derived parameters, the calculated degree of precipita-
ered as a constant as the function of temperature, a simpler tion as a function of time at various heating rates is shown
approach than applying the additivity rule can be used to in Fig. 9, where the experimental values from DSC were

0.01
Heat flow (mW/mg)

0.02

50 K/min
40 K/min
0.03 30 K/min
20 K/min
10 K/min

0.04
500 510 520 530 540 550 560 570 580 590 600 610
Temperature (°C)

Fig. 8. Processed DSC peaks for the main precipitation process at different heating rates.
Z. Guo et al. / Materials Science and Engineering A 373 (2004) 10–20 17

0.9

0.8

0.7
Degree of precipitate fraction

0.6

0.5

0.4 50 K/min (Exp.)


50 K/min (Cal.)
40 K/min (Exp.)
0.3 40 K/min (Cal.)
30 K/min (Exp.)
0.2
30 K/min (Cal.)
20 K/min (Exp.)
20 K/min (Cal.)
0.1 10 K/min (Exp.)
10 K/min (Cal.)

0
0 50 100 150 200 250 300 350 400
Time (s)

Fig. 9. Comparison between the calculated degree of precipitation fraction and DSC experimental results.

plotted for comparison. A good agreement was achieved be- than 10 h, the present model demonstrated good agreement
tween the calculated and experimental results. with experimental observations. When ageing time is longer
than 10 h at 538 ◦ C, the experimentally observed reverted
austenite fraction is lower than the predicted amount. It
4. Quantification of isothermal ageing kinetics should be noted that the kinetic parameters used are de-
rived from continuous heating experiments. In essence,
4.1. Quantification of austenite reversion during ageing calculation of isothermal kinetics using such parameters is
an extrapolation procedure and errors may be expected. A
The good agreement between experimental and calculated close look at Fig. 6 reveals that the discrepancy between
degree of transformation at different heating rates confirms experimental and calculated transformed fractions becomes
the applicability of the JMA theory to the transformation large when the heating rate is slow. The long-term isother-
and the accuracy of the derived JMA kinetic parameters n, mal transformation kinetics may deviate from the experi-
k0 , and E. These parameters can be used to trace the course mental values. The alloy used in the earlier work by Peters
of austenite reversion in the C250 alloy in real practice for is Fe–18.6Ni–7.65Co–4.97Mo–0.45Ti–0.1Al–0.02C (wt.%)
any heating path. In the present work, the austenite reversion [15], in comparison with Fe–18.06Ni–8.29Co–5.01Mo–
at a given ageing temperature can be simulated by employ- 0.47Ti–0.098Al–0.06C, the alloy studied. Such composition
ing Eq. (3) for an isothermal process. Fig. 10 shows the difference, though not large, may contribute to some extent
degree of austenite reversion as a function of time at three to the deviation illustrated in Fig. 11.
ageing temperatures, 427, 482 and 538 ◦ C. Peters studied Detectable amount of austenite forms after 50 h at 482 ◦ C
the austenite reversion during ageing of a C250 grade at [4], which can be reflected from our calculation, Fig. 10.
538 ◦ C [15] and his results are plotted in Fig. 11, where the One may have noticed that at the early stage of the trans-
calculated austenite fraction and the equilibrium fraction at formation, Fig. 6, the discrepancy between the experimen-
this temperature are also shown for comparison. As can be tal and the calculated curves is large. However, no efforts
seen at the range of practical use, i.e. ageing time no longer were made to improve the fitting by allowing n to change
18 Z. Guo et al. / Materials Science and Engineering A 373 (2004) 10–20

0.6

538 °C
482 °C
427 °C
Austenite fraction

0.4

0.2

0
3 4 5 6 7
0.01 0.1 1 10 100 1 10 1 10 1 10 1 10 1 10
Time (h)

Fig. 10. Calculated degree of austenite reversion at different ageing temperatures.

with temperature and/or heating rate. The model obtained plied a precipitation of Ni3 Ti or Ti-enriched particles. There
that way would be more complicated and less useful than is not yet effective experimental method to characterize the
the present model where n is taken as a constant. precipitate fraction. The fact that the composition of the
hardening precipitates changes with time makes the estima-
4.2. Quantification of precipitation kinetics during ageing tion of precipitate fraction even more difficult [4,9]. Nev-
ertheless, the quantification method in this work provided
Using the kinetic parameters n, k0 , and E derived in a way to simulate the precipitation kinetics in precipitation
Section 3.2, the degree of precipitation process at a given hardening steels.
temperature can be simulated. Fig. 12 shows the ageing ki-
netics at three temperatures, 427, 482 and 538 ◦ C. It is worth 4.3. Age hardening kinetics of C250
recalling that a 100% precipitation corresponds to an actual
precipitate fraction 2.0 vol.%. The precipitation process in The age hardening kinetics of the C250 alloy was inves-
the C250 alloy was considered as one stage, without dis- tigated by hardness measurement, and results are shown in
tinguishing the types of precipitates and their precipitation Fig. 13. The ageing time to reach peak hardness for 538,
periods, though Vasudevan et al. [4] reported that Fe2 Mo 482, and 427 ◦ C is about 3, 10, and >50 h respectively. Com-
appears after ageing over 3 h at 482 ◦ C. The contraction in paring the kinetics curves in Figs. 10 and 13, there is no
a dilatometric curve due to precipitate formation increases austenite formed at peak hardness position for temperature
with the amount of titanium in the matrix [26], which im- 427 and 482 ◦ C, but about 4.4% for 538 ◦ C. The austenite
formed at 538 ◦ C may partly contribute to the lower peak
hardness at this temperature. The models indicate that pre-
0.6
cipitation growth completes (10, 1, and 0.1 h at 427, 482 and
538 ◦ C) ahead of the peak hardness (>50, 10, and 3 h at 427,
0.5
482 and 538 ◦ C). This is of particular interest. It indicates
Austenite fraction

0.4
that the size increment of precipitates during age hardening
should mainly follow coarsening rather than growth kinet-
0.3
ics, in agreement with previous work on a PH13-8 Mo alloy
[9].
0.2

Equ. (0.55)
0.1 Cal. 5. Discussion on kinetics parameters
Exp. (Ref.15)
0
The activation energy of austenite reversion can be con-
0.1 1 10 100 1000
sidered close to the activation energy for lattice diffusion
Time (h)
of substitutional atoms in ␣-iron, Ni 245.8 kJ/mol [49],
Fig. 11. Calculated fraction of retained austenite in comparison with Ti 272 kJ/mol [50], and Mo 238 kJ/mol [50]. According
experimental data at 538 ◦ C. to Christian’s kinetics theory [47], the n value for grain
Z. Guo et al. / Materials Science and Engineering A 373 (2004) 10–20 19

538 °C
Degree of precipitate fraction

0.8 482 °C
427 °C

0.6

0.4

0.2

0
4 3
1 10 1 10 0.01 0.1 1 10 100
Time (h)

Fig. 12. Calculated degree of precipitation at different ageing temperatures. A 100% precipitation process corresponds to an actual precipitate fraction
of 2.0 vol.%.

boundary nucleation with site saturation is 1, to which the experiments, Vasudevan et al. [4] estimated the activation
n value for austenite reversion, 0.97, is very close. During energy for a C250-grade as from 101 to 181 kJ/mol with in-
reversion austenite forms on the existing precipitates, which creasing time, whereas n value is about 0.2–0.4 for different
makes site saturation a rational assumption. An estimation temperatures. It should be noted that, however, although the
of n value for C250 was carried out using the data from Ref. value of kinetic parameters may give some indication to the
[15] (Fig. 11) by plotting ln(ln(1/(1 − f/feq ))) versus ln(t). mechanism of the phase transformations taking place, there
The value of n varies considerably from 1.27 for t ≤ 9 h, is yet lack of fundamental understandings on this topic.
to 0.15 for longer time, which may be due to the change of Strictly, the theoretical analysis in the present work is
transformation mechanisms. The value of n for the austen- flawed in the sense that the interaction between precipitation
ite formation during ageing was calculated as 0.5 for a and austenite formation was not considered. It is believed
PH17-4 precipitation hardening stainless steel, with acti- that there is an overlap between these two transformations in
vation energy as 138.8 kJ/mol [16]. The activation energy the DSC curve. Since each of them seems to be a complete
for precipitate formation can be considered as being close peak on its own, it is difficult to separate them. However,
to the lattice diffusion activation energy of Ni, Ti, and Mo this should not undermine the general conclusions of the
in ␣-Fe, in agreement with the study by Peters and Cupp paper.
[33] on a Fe–18Ni–8Co–5Mo alloy from resistivity data.
The n value, 1.46, seems to suggest that the precipitation
process follows the diffusion-controlled growth of small
dimensions at zero nucleation rate [47]. Based on resistivity 6. Conclusions

Differential scanning calorimetry was used to trace the


700 phase transformation kinetics of an 18 wt.% Ni maraging
538˚C
steel C250 during heating. Based on the calorimetric data,
482˚C
600 the kinetics of precipitate formation and austenite reversion
427˚C
was modeled using the Johnson–Mehl–Avrami theory. The
Vickers hardness

500 derived JMA kinetic parameters were used to calculate the


isothermal ageing kinetics. Results lead to the following
400
conclusions.

300
(i) Reactions corresponding to precipitate formation
(exothermic) and austenite reversion (endothermic)
were revealed in the DSC curves during heating.
200 //
(ii) Good agreement between the calculated transformed
00 1 10 100
Ageing time (h) fraction and the experimental data from DSC measure-
ment was achieved for both precipitate formation and
Fig. 13. Age hardening curves at different ageing temperatures. austenite reversion.
20 Z. Guo et al. / Materials Science and Engineering A 373 (2004) 10–20

(iii) Both the precipitation process and austenite reversion [21] A. Borrego, G. Gonzalez-Doncel, Mater. Sci. Eng. A 245 (1998) 10.
during ageing seem to be controlled by the lattice dif- [22] R. Benedictus, A. Bottger, E.J. Mittermeijer, Z. Metallkd. 89 (1998)
168.
fusion of substitutional atoms. [23] J.A. Augus, J.E. Bennett, J. Therm. Anal. 13 (1978) 283.
(iv) Precipitation process may reach equilibrium before the [24] S. Malinov, Z. Guo, W. Sha, A. Wilson, Metall. Mater. Trans. A 32
peak hardness is achieved during ageing, which indi- (2001) 879.
cates that the hardening effects may be due to precipi- [25] S. Malinov, Z. Guo, W. Sha, A.F. Wilson, Z.X. Guo, in: Proceedings
tate coarsening. of the IOM Communications Conference on Titanium Alloys at El-
evated Temperature: Structural Development and Service Behaviour,
Birmingham, UK, 11–12 September 2000, Institute of Materials,
London, 2001, pp. 69–88.
Acknowledgements [26] H. Kessler, W. Pitsch, Acta Metall. 15 (1967) 401.
[27] F. Habiby, A. ul Haq, F.H. Hashmi, A.Q. Khan, in: Proceed-
Ms. W. Beck at Allvac UK Ltd. is thanked for providing ings of the International Conference on Martensitic Transformations
the C250 alloy. This work was carried out within the project (ICOMAT-86), Nara, Japan, 26–30 August 1986, The Japan Institute
of Metals, 1986, pp. 560–565.
of ‘Computer Modeling of the Evolution of Microstruc-
[28] A. Goldberg, D.G. O’Connor, Nature 213 (1967) 170.
ture during Processing of Maraging Steels’ sponsored by [29] A. Goldberg, in: R.F. Decker (Ed.), Source Book on Maraging Steels,
the Engineering and Physical Sciences Research Council, ASM, Metals Park, OH, 1979, pp. 41–51.
UK, under Grant No. GR/N08971. Dr. Sha’s contribution [30] K. Detert, Trans. Q. ASM 59 (1966) 262.
to this work is sponsored under The Royal Academy of [31] N. Bui, F. Dabosi, Cobalt 57 (1972) 192.
[32] G. Saul, J.A. Roberson, A.M. Adair, in: R.F. Decker (Ed.), Source
Engineering’s Global Research Award Scheme.
Book on Maraging Steels, ASM, Metals Park, OH, 1979, pp. 52–56.
[33] D.T. Peters, C.R. Cupp, in: R.F. Decker (Ed.), Source Book on
Maraging Steels, ASM, Metals Park, OH, 1979, pp. 317–325.
References [34] U.K. Viswanathan, G.K. Dey, M.K. Asundi, Metall. Trans. A A24
(1993) 2429.
[1] K. Rohrbach, M. Schmidt, Metals Handbook, vol. 1, 10th ed., ASM, [35] W. Sha, Z. Guo, J. Alloys Compd. 290 (1999) L3.
Materials Park, OH, 1990, pp. 793–800. [36] ASM International Handbook Committee, Metals Handbook, vol. 2,
[2] S. Floreen, R.F. Decker, in: R.F. Decker (Ed.), Source Book on 10th ed., ASM, Materials Park, OH, 1990, p. 1119.
Maraging Steels, ASM, Metals Park, OH, 1979, pp. 20–32. [37] V.V. Sagaradze, V.E. Danilchenko, P. L’Heritier, V.A. Shabashov,
[3] R.F. Decker, S. Floreen, in: R.K. Wilson (Ed.), Maraging Steels: Mater. Sci. Eng. A A337 (2002) 146.
Recent Developments and Applications, TMS-AIME, Warrendale, [38] Y.K. Lee, O. Kwon, J. Korean Inst. Met. Mater. 30 (1992) 1317–
PA, 1988, pp. 1–38. 1325.
[4] V.K. Vasudevan, S.J. Kim, C.M. Wayman, Metall. Trans. A 21 (1990) [39] E.J. Mittemeijer, L. Cheng, P.J. Van der Schaaf, Metall. Trans. A
2655. A17 (1986) 1441.
[5] W. Sha, A. Cerezo, G.D.W. Smith, Metall. Trans. A 24 (1993) 1221. [40] L. Cheng, C.M. Brakman, B.M. Korevaar, E.J. Mittemeijer, Metall.
[6] W. Sha, A. Cerezo, G.D.W. Smith, Metall. Trans. A 24 (1993) 1233. Trans. A A19 (1988) 2415.
[7] W. Sha, A. Cerezo, G.D.W. Smith, Metall. Trans. A 24 (1993) 1241. [41] E.J. Mittemeijer, J. Mater. Sci. 27 (1992) 3977.
[8] W. Sha, A. Cerezo, G.D.W. Smith, Metall. Trans. A 24 (1993) 1251. [42] W.A. Johnson, R.F. Mehl, Trans. Am. Inst. Min. Metall. Eng. 135
[9] Z. Guo, W. Sha, D. Vaumousse, Acta Mater. 51 (2003) 101. (1939) 416.
[10] V. Seetharaman, M. Sundararaman, R. Krishnan, Mater. Sci. Eng. [43] M. Avrami, J. Chem. Phys. 9 (1941) 177.
47 (1981) 1. [44] J. Burke, The Kinetics of Phase Transformation in Metals, Pergamon
[11] P.W. Hochanadel, C.V. Robino, G.R. Edwards, M.J. Cieslak, Metall. Press, Oxford, 1965, p. 46.
Mater. Trans. A 25 (1994) 789. [45] E. Scheil, Arch. Eisenhüttenwes. 12 (1935) 565.
[12] R. Taillard, A. Pineau, B.J. Thomas, Mater. Sci. Eng. 54 (1982) 209. [46] J.W. Christian, The Theory of Transformations in Metals and Alloys,
[13] R. Taillard, A. Pineau, Mater. Sci. Eng. 56 (1982) 219. Part I: Equilibrium and General Kinetic Theory, second ed., Pergamon
[14] A. Ali, M. Ahmed, F.H. Hashmi, A.Q. Khan, Mater. Sci. Technol. Press, Oxford, 1975, pp. 418–548.
10 (1994) 97. [47] J.W. Christian, The Theory of Transformations in Metals and Alloys,
[15] D.T. Peters, in: R.F. Decker (Ed.), Source Book on Maraging Steels, Part I: Equilibrium and General Kinetic Theory, second ed., Pergamon
ASM, Metals Park, OH, 1979, pp. 304–316. Press, Oxford, 1975, p. 542.
[16] S. Isobe, M. Okabe, Denki Seiko 54 (1983) 253. [48] S.J. Jones, H.K.D.H. Bhadeshia, Metall. Mater. Trans. A A28 (1997)
[17] Z. Guo, W. Sha, E.A. Wilson, Mater. Sci. Technol. 18 (2002) 377. 2005.
[18] Z. Guo, W. Sha, Mater. Sci. Technol. 18 (2002) 529. [49] E.A. Brandes, G.B. Brook, Smithells Metals Reference Book, seventh
[19] M.P. Jackson, M.J. Starink, R.C. Reed, Mater. Sci. Eng. A 264 ed., Butterworth-Heinemann, Oxford, 1992, pp. 13–20.
(1999) 26. [50] M.K. Krishtal, Diffusion Processes in Iron Alloys, Translated from
[20] J. Vazquez, P. Villares, R. Jimenez-Garay, J. Alloys Compd. 257 Russian by the Israel Programme for Scientific Translations Ltd.,
(1997) 259. Jerusalem, Israel, 1970, pp. 175–203.

Вам также может понравиться