You are on page 1of 8

Materials Science and Engineering A 527 (2010) 42334240

Contents lists available at ScienceDirect

Materials Science and Engineering A


journal homepage: www.elsevier.com/locate/msea

Strengthening of HSLA steels by cool deformation


A. Fatehi , J. Calvo, A.M. Elwazri, S. Yue
Department of Mining, Metals and Materials Engineering, McGill University, 3610 University Street, Montreal, QC, Canada H3A 2B2

a r t i c l e

i n f o

Article history:
Received 9 June 2009
Received in revised form 9 March 2010
Accepted 11 March 2010

Keywords:
Cool deformation
HSLA
Linepipe steels
Cu precipitation
Dynamic precipitation

a b s t r a c t
In microalloyed steels, the renement of ferrite grains together with a controlled amount of precipitation has key roles in the mechanical properties improvement. Applying small amounts of deformation, at
very low hot working temperatures (i.e. coiling temperature), in the ferrite region (i.e. cool deformation)
has an appreciable strengthening effect via controlling the nal microstructure of the steel. One of the
microstructural effects is thought to be the much ner and more uniformly dispersed precipitates in the
steel matrix. In the present study, the effects of Nb and Cu on mechanical properties and corresponding microstructures in steels with different levels of cool deformation are investigated. The mechanical
properties of the samples were determined using the shear punch test and the microstructure was examined by scanning and transmission electron microscopy. Thermodynamic simulations with FactSage were
done to further analyze the precipitation possibility of different elements. It has been found that these
alloying elements respond very well to cool deformation, with the strength being highest in steels containing both Nb and Cu. However, a cool deformation effect in the non-Nb and Cu bearing steel is also
observed. In all cases, it was conrmed that precipitation plays a key role in the effect of cool deformation,
with much of the precipitation taking place dynamically. Nevertheless, static processes also seem to have
a measurable effect on room temperature properties. Even low amounts of copper (e.g. 0.4 wt%) can
contribute to strengthening of the steel. The Cu addition is found to affect the mechanical properties by
affecting the precipitation and growth of Nb compounds.
Crown Copyright 2010 Published by Elsevier B.V. All rights reserved.

1. Introduction
High-strength low alloy (HSLA) steels demonstrate unique
properties such as high strength, excellent ductility and good
weldability. These steels have found their way into a variety of
applications including pipelines, cars, pressure vessels, ships and
offshore platforms [1]. The excellent properties are essentially due
to small additions of microalloying elements: Ti, Nb, V and B.
Microalloy carbides, nitrides and carbonitrides are important in the
control of austenite recrystallization and grain growth during steel
processing and welding, leading to a desirable grain renement.
Moreover, microalloying, controlled rolling and accelerated cooling lead to the formation of non-polygonal or acicular ferrite as
well as bainite which further contribute to strengthening.
Carbonitride formation can take place in austenite during
rolling, at the austenite/ferrite interface during transformation or
in supersaturated ferrite during nal cooling [2]. Precipitation in
ferrite, however, has a stronger direct effect on strengthening. Precipitation in austenite, although benecial for grain renement,
can consume almost the total amount of microalloy elements in
solution. Since these elements could eventually precipitate in fer-

Corresponding author. Tel.: +1 514 691 4680; fax: +1 514 398 4492.
E-mail addresses: arya.fatehi@gmail.com, arya.fatehi@mail.mcgill.ca (A. Fatehi).

rite, their complete precipitation in austenite will decrease the


room temperature yield strength [36]. Beside conventional thermomechanical processing steps, deformations applied at relatively
low temperatures, i.e. coiling temperatures, are found to enhance
the mechanical properties. It was suggested that cool deformation leads to strain-induced particles that are much ner, more
uniformly dispersed and more numerous than those in the corresponding aged-only specimen [7].
Nowadays, a considerable amount of steels used in industry
comes from scrap leading to a residual copper content. Copper is
very effective in strengthening of steels, having one of the highest
substitutional solid solution strengthening effects out of the common substitutional alloying additions [4]. Apart from that, Cu is
considered to be one of the rare substitutional elements known to
cause age hardening in Fe-based systems, from ne -Cu particles
formed in aging [815].
The precipitation of Cu in iron is usually observed in the temperature range of 400650 C. Therefore any deformation in this
range of temperature, which also coincides with the coiling temperature range, can accelerate such precipitation. Pre-strain has been
reported to be very effective on strengthening by Cu precipitates
(e.g. [16,17]). Although precipitation of Cu has been exhaustively
studied in the past in steels containing >1 wt% Cu, precipitation in
the steels containing relatively low copper amounts, e.g. 0.4 wt%
in the present work, has not been well studied. This work deals with

0921-5093/$ see front matter. Crown Copyright 2010 Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2010.03.036

4234

A. Fatehi et al. / Materials Science and Engineering A 527 (2010) 42334240

Table 1
Chemical composition of the steels in wt%.
Steel

Mn

Mo

Nb

Cu

Si

Ti

Al

Mo
Hi-Nb
Cu

0.041
0.039
0.049

1.473
1.510
1.589

0.32
0.28
0.28

0.068
0.069

0.38

0.06
0.084
0.066

0.020
0.018
0.020

0.0091
0.0095
0.0090

0.023
0.011
0.040

the effect of cool deformation at 400 C on strength of the microalloyed steels. Moreover the response to the cool deformation with
and without Nb and relatively low amounts of Cu will be studied.
2. Experimental procedures
Three experimental steels were used in this study with the
chemical compositions listed in Table 1. These alloys have very low
carbon contents along with levels of Mn (1.5 wt%) and Mo (0.3 wt%)
that are fairly typical of linepipe steels. Nb is a key microalloying
element, and Cu is an important residual that potentially affects
the properties. The Hi-Nb steel has about 0.07 wt% Nb and the
Cu steel contains 0.38 wt% Cu in addition to a similar amount of
Nb. The Mo steel, which does not contain Nb or Cu, was selected
because it was anticipated that cool deformation would not lead to
precipitation in this alloy, and therefore, by comparison with the
other alloys, the role of any other strengthening mechanisms could
be better identied. These three steels were cast and hot-rolled
to 2 cm thick plates at CANMET Materials Technology Laboratory
(Ottawa, Ontario, Canada), using vacuum melting. Cylindrical compression specimens, 7.6 mm in diameter and 11.4 mm in length,
were machined from the as-hot rolled plate with their longitudinal
axes parallel to the rolling direction.
In order to dissolve the Nb(C,N) and possibly Cu particles,
samples were reheated according to the temperature required to
solutionize all Nb precipitates, which was calculated to be 1151 C
for the Hi-Nb steel and 1177 C for the Cu steel using Irvines formula [18]:

log[Nb] C + 12

N
14

= 2.26

6770
T

(1)

From the FactSage thermodynamics modeling, the corresponding temperatures are 1200 C and 1180 C respectively. The
temperatures are reasonably close, but the trend is different. This
may be because the Al and Ti levels are slightly higher in Cu steel,
leading to less N available for Nb precipitation. This is not accounted
for in Irvines formula. From the FeCu binary phase diagram and
FactSage results, the Cu dissolution temperature is about 640 C for
0.38 wt% Cu amount. The maximum solubility of Cu in austenite is
about 11 wt% at 1200 C in austenite.
The reheating was 1200 C for 20 min in an argon atmosphere
followed by cooling down to 400 C at 1 C/s. To perform cool
deformation, after the temperature was stabilized at 400 C, deformations of 6 or 20% strain at a constant strain rate of 1 s1 were
applied to the samples; following these deformations, aging treatments were carried out, i.e. the specimens were held at 400 C for
10, 30 and 180 min. For aged-only samples, all the stages were the
same but no deformation was applied before aging. Thereafter, the
samples were all air-cooled to ambient temperature. An MTS servohydraulic machine was used for the compression testing. It is
noteworthy that in the compression part of the cool deformation
process, a part of the strain is recoverable elastic strain. Thus, 6%
cool deformation only generated 2% permanent plastic strain, and
a plastic strain of 14% was generated by 20% cool deformation. This
was due to elastic deformation of the specimen and the compliance
of the servohydraulic machine. However, in the rest of this article,
the specimens will be referred to using the nominal cool deformation strains. The strengths were determined at room temperature

by the Shear Punch Test (SPT) method, which was used because the
small size of the test specimen makes it difcult to make tensile test
specimens. Samples of 300350 m thickness were prepared by
grinding, nishing with 1200-grit SiC. A at-tip cylindrical punch
of 1.5 mm diameter was used in order to perform an instrumented
punch out of discs of material. The tensile and yield strength of
the samples were determined from the loaddisplacement curves
obtained during SPT by the following equation [19]:
eff =

LF
= C
2rt

(2)

where L is the load (N), F is the frictional load (N), r is the punch
radius (mm), t is the sample thickness (mm) and  is the corresponding uniaxial stress (MPa). The coefcient C is an empirical
factor equal to 0.64 for tensile strength and 0.54 for yield strength
[19]. At least three shear punch tests per data point were carried
out.
Microstructures of the cool deformed and aged-only specimens
were studied under the optical and scanning electron microscopes.
Additionally, in order to investigate the microstructural evolution,
specimens were quenched after 0 min and 5 min aging (holding) times before applying the cool deformation strain, and their
microstructures were compared. Cross-sectional samples were cut
and ground using silicon carbide papers to 800-grit. Then, further
polishing was done with 3 and 1 m water based diamond suspensions on short nap cloths with an alcohol-based lubricant. Samples
were etched using 2% Nital. A Hitachi s-4700 eld emission gun
SEM (FEG-SEM) was used for microstructure observation. The SEM
was equipped with an Oxford Instruments energy dispersive spectroscopy (EDS) detector. To further analyze the precipitates in the
structure a Philips CM-200 transmission electron microscope was
utilized with a Philips EDS detector. The operating voltage was
200 keV. Samples were rst mechanically thinned to below 100 m
using 1200 grit grinding paper. A Struers Tenupol-3 jet polisher
was used for further thinning using 95% glacial acetic acid and 5%
perchloric acid at 70 V and 15 C. Thermodynamic modeling was
performed using the FactSage software.

3. Results
3.1. Mechanical properties
In Fig. 1a the yield strengths of the three alloys in the aged-only,
6% and 20% cool deformation conditions are compared at a constant aging time of 30 min. The yield strength (Ys ) of the Mo steel
increases from 470 MPa to 614 MPa after 20% cool deformation. The
trend is similar for the other two steels (638704 MPa for the Hi-Nb
and 672746 MPa for the Cu steel) and, despite some minor uctuations for the Hi-Nb steel, there is a general increase in strength
of all steels after cool deformation. The ultimate true stress (Su )
behavior, as shown in Fig. 1b, is similar to that of the Ys ; however,
cool deformation is found to be more effective on the Ys than the Su .
The effect of aging time on yield and tensile strengths in the
aged-only conditions is provided in Fig. 2. Aging changes the yield
strength in the Hi-Nb and Cu steels, peaking at 30 min, but has no
effect on the Mo steel. The peak yield strength is 638 and 672 MPa
in the Hi-Nb and Cu steels after 30 min aging. For all the steels, Su
in the aged-only condition is not affected by aging time. In the case
of 20% cool deformation, the maximum Ys is found after 30 min of
aging in the Hi-Nb and Cu steels whereas the strength is higher
in the Mo steel after 180 min of aging (Fig. 3). Su has an identical behavior although showing a less signicant response to aging
time. It can be noted that the Mo steel shows an aging behavior
after cool deformation, in contrast to the aged-only condition.

A. Fatehi et al. / Materials Science and Engineering A 527 (2010) 42334240

4235

Fig. 3. Effect of aging time after 20% cool deformation on yield strength.
Fig. 1. Effect of cool deformation after 30 min aging on (a) yield strength (Ys ) and
(b) tensile strength (Su ).

3.2. Microstructure
Fig. 4 shows the optical microscopy images of the three alloys in
the aged-only and 20% cool deformed conditions, both after 30 min
aging. Since no austenite conditioning was done on the steels the

grains are relatively coarse. The microstructure of the Mo steel


generally consists of a matrix with polygonal (equiaxed) ferrite
grains and in some regions, islands of second phases which can be
degenerate pearlite, bainite, martensite/retained austenite (M/A)
or acicular ferrite (AF). In the Hi-Nb and Cu steels, Fig. 4cf, the
structures mainly consist of a matrix of ferrite that is no longer
polygonal and has irregular grains (called quasi-polygonal ferrite
[20]). Higher magnications (SEM images) reveal islands of second
phases distributed between ferrite grains. This two-phase structure
of quasi polygonal ferrite and the second phase islands is named
granular bainite in the literature [21,22].
In the specimens quenched from 400 C, just before aging or
applying the cool deformation strain, the M/A phase has no resolvable substructure (Fig. 5a), i.e. the M/A phase is featureless. After
aging, the featureless M/A phase is transformed to M/A with substructure (Fig. 5b). Grain renement is observed comparing the
microstructures of the Mo steel in the aged-only and cool deformed
conditions (Fig. 4a and b). Moreover, in the cool deformed steel
(Fig. 4b) the islands of second phase are ner and more numerous.
In other words, the second phases are more homogenously distributed in the structure, partly as a result of the general renement
of the structure. The average grain size (using mean linear intercept
(Heyn) method [4]) decreases from 21 to 17 m in the case of the
Mo steel. Determining the grain size is more problematic in the
other two steels, particularly in the Cu steel because of the high
volume fraction of acicular ferrite. Nonetheless, the granular part
of these structures appears to be ner after the cool deformation in
these steels as well.
3.3. Precipitates

Fig. 2. Effect of aging time on yield strength in aged-only condition.

In the Mo-steel the only kind of precipitates detected were


Ti(C,N), which can vary in size from several microns to several
nanometers. The large Ti(C,N) precipitates were usually found
together with MnS and Al concentrations (nitrogen was not

4236

A. Fatehi et al. / Materials Science and Engineering A 527 (2010) 42334240

Fig. 4. Optical micrographs of (a) Mo steel: aged only for 30 min at 400 C. (b) Mo steel: 20% cool deformed and aged at 400 C. (c) Hi-Nb steel: aged only for 30 min at 400 C.
(d) Hi-Nb steel: 20% cool deformed and aged at 400 C. (e) Cu steel: aged only for 30 min at 400 C. (f) Cu steel: 20% cool deformed and aged at 400 C.

detected). In the Hi-Nb steel in addition to Ti(C,N) precipitates,


numerous Nb(C,N) precipitates were found in the structure (Fig. 6).
The composition of the very ne precipitates could not be veried, even using TEM, but the relatively large ones were identied
as (Nb,Ti)(C,N) (Fig. 6a). It should be noted that the copper peak
in the EDS spectrum in Fig. 6c originates from the TEM sample
holder and not from the sample. In the Cu-steel the main precipitates are also Nb bearing. Fig. 7a and b shows the effect of cool
deformation on precipitation level of the Cu steel. No ne precipitates were observed in the micrograph of the aged-only steel
(Fig. 7a). Numerous precipitates, below 10 nm in diameter, appear
in the microstructure after applying cool deformation (Fig. 7b). Cool
deformation stimulates the precipitation of a considerable amount
of alloying elements that otherwise would remain in the solid solution. Fig. 7c and d presents bright eld and dark eld images of the
Cu steel, cool deformed and aged. The ne precipitates and parts
of large cuboidal Ti,Nb(C,N) precipitates are all illuminated in the
dark eld image taken using the [2 0 0]Nb diffraction in selected area
diffraction pattern. It can be inferred that the ne precipitates must
also be mainly Nb(C,N).

4. Discussion
4.1. Effect of cool deformation
The cool deformation performed at 400 C is found to possess
an appreciable strengthening effect. Similar effects have also been
observed in previous work [7,23,24]. The strengthening mechanisms which could be operative as a result of cool deformation
are strain hardening, phase transformation, grain renement and
precipitation strengthening. Separating the contributions of each
of these mechanisms has not been done rigorously. However, it
is clear that the strengthening effect cannot be only attributed to
strain hardening [24].
Applying cool deformation leads to strain-induced precipitation
of precipitate forming alloying elements in the structure, as shown
in Fig. 7. The strain induced precipitation can be dynamic or static.
The former precipitates form during the deformation whereas the
latter form after the deformation and during the aging stage.
The effect of cool deformation on the static precipitation can be
attributed to the effect of strain on dislocation density. An incre-

A. Fatehi et al. / Materials Science and Engineering A 527 (2010) 42334240

4237

Fig. 5. M/A phase observed in the aged-only Cu steel: (a) 0 aging time and (b) 5-min
aging time.

ment in dislocation density provides more nucleation sites for


precipitates during the deformation (dynamic precipitation) and
afterwards during aging (static precipitation). The coarsening rate
of these precipitates may also be increased by the presence of dislocations, which increase diffusivity. Numerous ne precipitates can
also be observed in the aged-only steels. However, longer aging
times are required to get optimum precipitate characteristics for
strengthening.
Another observation is the absence of featureless M/A phase
after aging, which suggests that, at 400 C, the austenite transformation is incomplete just prior to cool deformation and that cool
deformation leads to more transformation. There are no carbides
in granular bainite [21,22] and carbon partitioned from ferritic bainite formation retains metastable austenite. It is possible that this
retained austenite, i.e. the featureless M/A phase, can transform
to martensite during cool deformation (strain-induced transformation). The resulting martensite could be tempered during aging,
leaving small tempered martensite regions in the structure and a
higher strength.
4.2. Effect of aging time
Considering variations of properties with regard to aging time,
Figs. 2 and 3, reveal that all the steels demonstrate an agehardening behavior. The YS and Su increase at 30-min aging and,
with the exception of the Mo steel cool deformed to 20%, decrease
after 180 min. There is an optimum precipitate size and number in
order to get the best age-hardening, which is attained after a certain
aging time. Further increasing this time leads to overaging conditions resulting from excessive growth of precipitates which affects

Fig. 6. Precipitates in Nb steel containing both Ti and Nb: (a) SEM micrograph
and corresponding EDS elemental map and (b) TEM micrograph (ne precipitates
shown by black arrows) and (c) EDS spectrum from the large (Nb,Ti)(C,N) precipitate
indicated by the white arrow in (b).

the strength. This further indicates that precipitation strengthening


is a key strengthening mechanism.
In all conditions, the Cu steel has the highest strength while
the Mo steel has the least. For the Mo steel the peak strength
occurs after 180 min aging, whereas in the other steels, 30 min
aging generates the peak strength. The Mo steel starts showing an
aging response after applying the cool deformation and in fact the
response is the highest compared to other two steels, as will be discussed below. The effect of cool deformation on aging behavior of
steels indicates strain-hardening, although being important, is not
the only strengthening mechanism.
4.3. Effect of chemical composition
In order to further analyze the response of each steel to cool
deformation, the percentage increase in the strength compared to
the aged-only condition were determined as,

Ys increment (%) =

Ys,CD Ys,aged only


Ys,aged only

(3)

where Ys,CD and Ys,aged only represent the yield strength of the samples with and without cool deformation, respectively. As can be
seen in Fig. 8, all the steels are stronger after cool deformation. In the
case of the Mo and Hi-Nb steels, the increments in strength increase

4238

A. Fatehi et al. / Materials Science and Engineering A 527 (2010) 42334240

Fig. 7. TEM micrographs of the Cu-steel in (a) age-only condition (30 min) and (b) cool deformed + 30 min aged; (c) bright eld and (d) corresponding dark eld TEM images
of the cool deformed + 30 min aged Cu-steel.

with the aging time, probably due to increase in precipitation. However, in the case of the Cu steel, the increments decrease with aging
time. There are two possible interpretations of this behavior (i) the
steel is continuously softening or (ii) there is a peak strength which
occurs between 0 and 30 min aging time, i.e. the Cu addition leads
to peak strength at shorter aging times.
In terms of response to cool deformation, the effect of composition depends on the aging time. After 10 min, the Mo and Cu
steels show the best response. After 30 min, the Mo steel is the
best, whereas the Cu and Hi-Nb steels show the same response.
After 180 min, the Mo steel is still the best, but the Cu steel is now
the least responsive, although it still exhibits the best strength. In
general, then, the Mo steel responds best to 20% cool deformation
at all aging times.
4.4. Precipitation characteristics
Precipitation strengthening in the Mo steel can be attributed to
titanium carbonitride precipitates in the structure. However, Tibearing precipitates in microalloyed steels are usually expected
to form on solidication and prevent austenite grains coarsening
during reheating and roughing. Ti may form precipitates at lower

Fig. 9. Thermodynamic simulation of precipitation in the Cu steel.

temperatures, but these are not expected to contribute signicantly


to strength. Precipitation of molybdenum carbides in the structure is also possible although not detected. Five types of these
carbides reportedly form in the steel, i.e. MoC, Mo2 C, M23 C6 , M6 C,
and Fe2 MoC [1,25] and the rst two of them are detected in high
strength low carbon steels [1].
Table 2
Ti content inside the precipitates as well as solid solution in the three steels at low
temperatures.

Fig. 8. Percent increments in the yield strength of the steels due to 20% cool deformation.

Steel

Total Ti, wt%

In precipitates,
wt% (FactSage)

In solid
solution, wt%
(FactSage)

Mo
Hi-Nb
Cu

0.04
0.04
0.04

0.019
0.020
0.020

0.021
0.020
0.020

A. Fatehi et al. / Materials Science and Engineering A 527 (2010) 42334240

4239

Table 3
Volume fraction of precipitates calculated from FactSage results.
Precipitates

Steel

Precipitates (wt%)

Precipitates composition

Ti-bearing

Mo
Hi-Nb
Cu

0.025
0.025
0.01

47
45
49

Nb-bearing

Mo
Hi-Nb
Cu

0.085
0.095

3
15

at% Ti

The precipitation type and amount was predicted in simulations by the FactSage software. The typical result for the Cu steel
is demonstrated in Fig. 9. The large precipitates can be assumed to
form during solidication. The amount of these precipitates can be
calculated from the modeling results at 1200 C (Table 2). Due to
deviation from the equilibrium conditions, these FactSage results
cannot be employed for precipitates forming during the continuous
cooling. These precipitates are usually ner and can contribute to
strengthening more effectively. Table 3 shows the amount of the
Ti inside the precipitates and the matrix based on the thermodynamic modeling results. The Ti-bearing precipitate has a weight
fraction of 0.03% which contains 40 at% Ti. Therefore, the precipitation amount of Ti is calculated to be 0.019 wt%. Table 3 shows
that a signicant amount of the Ti remains in the solid solution at
1200 C. This result is conrmed by the ndings of Lu et al. [26]
in which the Ti in the solid solution is measured as 0.014 wt% at
1000 C and also Comineli et al. [27] who performed deformation
at higher temperatures (750950 C). Since the amount of these
precipitates is remarkably higher in cool deformed steels, it can
be deduced that they are induced by deformation in spite of low
working temperature. For this to happen, Ti has to exist in solid
solution. Therefore there is always a possibility of Ti precipitation
during the deformation at lower temperatures. The experimental
ndings, with regard to age hardening in the Mo steel, also suggest
that part of the Ti value remains in the solid solution.
In the Hi-Nb steel, the increment in strength is most probably due to precipitation of niobium carbonitrides. It is generally
accepted that although precipitation of Nb(C,N) in austenite
enhances the mechanical properties, it does not strengthen the
matrix after transformation of austenite to ferrite as these precipitates coarsen [28]. On the other hand, precipitation in ferrite plays
an important role in precipitation strengthening because the carbonitrides formed are very small and therefore more effective for
strengthening.
The only difference between the Cu steel and Hi-Nb steel used
in this study is the copper content of about 0.4 wt%. Therefore, in
addition to Nb(C,N), precipitation of Cu is possible. According to the
FeCu phase diagram, the solubility of Cu in -Fe is about 10 wt%
at 1200 C and near 0 wt% below 500 C. This is consistent with
the thermodynamic calculations which predict Cu precipitation
(Fig. 9). Therefore, the copper content can either precipitate inside
the matrix mainly on the dislocations and other structural defects
[29] or exist in the ferrite forming a supersaturated solid solution.
Such speculations based on thermodynamics are not supported by
experimental reports. For instance, according to [30] 0.54 0.08 at%
Cu remains in solid solution in an Fe-1.1 at% Cu-1.4% Ni alloy after
the aging treatment of 100 h at 400 C. Gladman also considers Cu
contents up to 2 wt% soluble in ferrite [4]. Nonetheless, deformation can accelerate the precipitation via producing nucleation sites
and hence expedite the formation of equilibrium phases predicted
by thermodynamics.
In the present work, it was not possible to determine whether or
not Cu precipitation had taken place, because of the inuence of the
copper bearing specimen holder as well as resolution and detec-

wt% in solid solution

Precipitates vol. fraction

0
0

0.0283
0.0262
0.0209

3.98 104
3.98 104
1.59 104

46
35

0.0309
0.0368

8.55 104
9.56 104

at% Nb

tion limit in the TEM study. Cu precipitation is feasible in the Cu


contents close to and above 1 wt% but few studies have been done
on lower Cu contents as in this study. Deschamps et al. [31] show
that for non-deformed Fe0.8 wt% Cu alloy Cu precipitates sizes
range between 2 and 8 nm after 100 h aging at 500 C and between
5 and 15 nm after 1000 h aging at this temperature. Also, for 10%
deformed specimen the Cu precipitates sizes range between 2 and
10 nm after 100 h aging and between 5 and 20 nm after 1000 h aging
at 500 C. Furthermore, Liu et al. [32] found that in 1.7 wt% Cu containing IF steel, cold worked for 75% and overaged at 600 C for 1 h,
in only one third of grains dispersed ne Cu precipitates with diameters of 415 nm were observed. In case of specimens with 0.25 wt%
Cu content, 75% cold worked and aged at 600 C for 1 h, no Cu precipitates were observed. Therefore it seems that exact detection of
probable Cu precipitates in steel with such low Cu content requires
high resolution characterization methods such as HR-TEM, SANS
(small angle neutron scattering) and AP-FIM (atom probe-eld ion
microscopy).
Very ne precipitates of less than 10 nm as observed in Fig. 7 can
be associated with Cu, Nb or Ti precipitates. According to the literature [32,33], Cu precipitates of less than 5 nm in size are expected.
In the TEM dark eld image, not all the precipitates have been
illuminated which suggests that some of them are different in composition or orientation. Since the solubility of the Cu in Fe is very
low, there is always a possibility of it existing in the form of precipitates. Nonetheless, the size of these precipitates is not expected to
be easily observable because of low aging temperature leading to
low diffusivity.
5. Conclusions
Cool deformation increases the strength of the Mo, Hi-Nb and
Cu steels appreciably when compared with the aged-only samples.
The strengthening mechanisms due to cool deformation are precipitation, work hardening as well as phase transformation of the
retained austenite. After cool deformation, the Cu steel has the
highest strength for all aging times. No evidence of Cu precipitation is found. Dynamic precipitation appears to occur during cool
deformation. Cool deformation induces precipitates which can be
Ti-based in non-Nb-bearing steels, and Ti- and Nb-based in Nbbearing steels. The Mo steel responded best to cool deformation,
which is probably because this steel exhibits the lowest non-cool
deformed strength.
Acknowledgements
This work was nancially supported by the Canadian Steel
Industry Research Association (CSIRA) and the Natural Science and
Engineering Research Council of Canada (NSERC).
References
[1] W.-B. Lee, S.-G. Hong, C.-G. Park, S.-H. Park, Metall. Mater. Trans. A 33 (2002)
16891698.

4240

A. Fatehi et al. / Materials Science and Engineering A 527 (2010) 42334240

[2] R.W.K. Honeycombe, in: e.a. J.M. Gray (Ed.), HSLA Steels: Metallurgy and Applications, ASM, Metals Park, OH, 1986, pp. 243250.
[3] A.J. DeArdo, Mater. Sci. Forum 284286 (1998) 1526.
[4] T. Gladman, The Physical Metallurgy of Microalloyed Steels, The Institute of
Materials, London, 1997.
[5] L. Meyer, F. Heisterkamp, K. Hulka, W. Mschenborn, in: T.S.T. Chandra (Ed.),
THERMEC 97, TMS, Warrendale, PA, 1997, pp. 8797.
[6] B. Dutta, E. Palmiere, Metall. Mater. Trans. A 34 (2003) 12371247.
[7] A. Elwazri, R. Varano, S. Yue, D. Bai, F. Siciliano, Metall. Mater. Trans. A 36 (2005)
29292936.
[8] S.S. Ghasemi, D. Banadouki, D. Yu, D.P. Dunne, ISIJ Int. 36 (1996) 6167.
[9] M. Mujahid, A.K. Lis, C.I. Garcia, A.J. DeArdo, J. Mater. Eng. Perform. 7 (1996)
247257.
[10] S.R. Goodman, S.S. Brenner, J.R. Low, Metall. Trans. 4 (1973) 2363.
[11] E. Hornbogen, R.C. Glenn, Trans. AIME 218 (1960) 10641070.
[12] N. Maruyama, N. Sugiyama, T. Hara, H. Tamehiro, Mater. Trans. JIM 40 (1999)
268.
[13] S.K. Dhua, A. Ray, D.S. Sarma, Mater. Sci. Eng. A 318 (2001) 197210.
[14] G.C. Hwang, S. Lee, J.Y. Yoo, W.Y. Choo, Mater. Sci. Eng. A 252 (1998) 256268.
[15] A. Ghosh, S. das, S. Chatterjee, Mater. Sci. Eng. A 486 (2008) 152157.
[16] C. Zhang, M. Enomoto, T. Yamashita, N. Sano, Metall. Mater. Trans. A 35 (2004)
12631272.
[17] T. Ishizaki, T. Yoshiie, K. Sato, S. Yanagita, Q. Xu, M. Komatsu, M. Kiritani, Mater.
Sci. Eng. A 350 (2003) 102107.

[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]

K.J. Irvine, F.B. Pickering, T. Gladman, J. Iron Steel Inst. 205 (1967) 161.
A.M. Elwazri, R. Varano, P. Wanjara, S. Yue, Can. Met. Quar. 45 (2006) 33.
T. Araki, Atlas for Bainitic Microstructures, ISIJ, Tokyo, Japan, 1992.
P.C.M. Rodrigues, E.V. Pereloma, D.B. Santos, Mater. Sci. Eng. A 283 (2000)
136143.
S.-C. Wang, R.-I. Hsieh, H.-Y. Liou, J.-R. Yang, Mater. Sci. Eng. A 157 (1992) 29
36.
A.M. Elwazri, D. Bai, F. Siciliano, S. Yue, Can. Metall. Quar. 45 (2006) 441449.
Q. Wei, J. Qu, A.M. Elwazri, D. Bai, S. Yue, in: MS&T 2004 Conference Proceedings,
New Orleans, LA, USA, 2004.
D.V. Shtansky, G. Inden, Acta Mater. 45 (1997) 28612878.
J. Lu, H. Henein, D.G. Ivey, in: IPC2006-10600, ASME, Faireld, 2006.
O. Comineli, A. Tuling, B. Mintz, L.P. Karjalainen, in: 62nd International Congress
of the ABM, Vitria, ES, Brazil, 2007.
E.V. Pereloma, B.R. Crawford, P.D. Hodgson, Mater. Sci. Eng. A 299 (2001) 2737.
M.G. Akben, B. Bacroix, J.J. Jonas, Acta Metall. 31 (1983) 161174.
M. Miller, Atom Probe Tomography, Kluwer Academic/Plenum, New York,
2000.
A. Deschamps, M. Militzer, W.J. Poole, ISIJ Int. 41 (2001) 196205.
J. Liu, D. Xie, D. Wang, L. Gao, M. Wen, S. Luo, P. Li, J. Gui, R. Wang, ISIJ Int. 39
(1999) 614616.
P.J. Othen, M.L. Jenkins, G.D.W. Smith, W.J. Phthian, Philos. Mag. Lett. 64 (1991)
383391.

You might also like