Sei sulla pagina 1di 16

Materials Characterization 142 (2018) 406–421

Contents lists available at ScienceDirect

Materials Characterization
journal homepage: www.elsevier.com/locate/matchar

Microstructure and texture evolution of duplex stainless steels with different T


molybdenum contents

Paula O. Malta , Francislaynne L. Dias, Ana Clara M. de Souza, Dagoberto B. Santos
Metallurgical and Materials Engineering Department, Universidade Federal de Minas Gerais, Antônio Carlos Avenue, 6627, Pampulha, Belo Horizonte, MG CEP 31270-
901, Brazil

A R T I C LE I N FO A B S T R A C T

Keywords: The microstructure and texture evolution of duplex stainless steels with different molybdenum contents were
Duplex stainless steel studied. For this purpose, hot-rolled 2205 and 2304 samples were solution-annealed at 1100 °C for 1800 s and
Microstructure then cold-rolled to a 75% thickness reduction. This was followed by an isothermal annealing at 900 and 1100 °C
EBSD for 180 s. The EBSD technique was used to determine the phase ratio, crystallographic texture, and micro-
Annealing
structural characteristics. The hot-rolled steel had a work hardened microstructure and the samples submitted to
CSL
intermediate annealing presented mainly deformation-free grains. Cold rolling resulted in a substantial reduc-
tion in the phase spacing with the formation of the strain-induced α′-martensite (SIM) from the metastable
austenite. After annealing at 900 °C, the ferrite grains retained the elongated shape of the cold-rolled condition
and the primary recrystallization progressed in the austenite phase. Increasing the Mo content increased the
resistance to strain-induced martensitic transformation and inhibited the grain growth of austenite phase during
annealing at 900 °C. Increasing annealing temperature to 1100 °C promoted the coarsening of both bcc and fcc
structures with the formation of annealing twins in the austenitic grains. The recrystallization kinetics and
subsequent grain growth occur earlier in ferrite than in austenite phase. Differences in Mo content did not
significantly alter the evolution of austenite texture, whereas in the ferrite phase the increase in Mo content
suppressed the γ-fiber development. Oriented nucleation was the predominant mechanism observed during the
recrystallization of ferritic grains in the 2205 samples. The development of ∑19a boundaries was associated with
the selected growth of {554}〈225〉α grains in the annealed 2304 samples. Annealing twins were associated with
the formation of Σ3 boundaries after the heat treatments. Grain coarsening resulted in a larger fraction of the
special boundaries in ferrite phase but inhibited the CSL formation in the austenite phase.

1. Introduction molybdenum increases the yield strength of duplex stainless steels but
decreases their ultimate tensile stress (UTS) [18,19]. Molybdenum ad-
Duplex stainless steels are a category of high-alloyed steels char- ditions are responsible also by improving the stability of the passive
acterized by a biphasic austenitic-ferritic microstructure [1–5]. Che- film and the resistance of duplex stainless steels to chloride-induced
mical composition is usually adjusted to achieve the optimum phase corrosion (pitting, crevice corrosion, and stress-corrosion cracking), in
ratio of approximately 1:1, which provides an attractive combination of aqueous environments particularly [20–22].
mechanical properties and corrosion resistance [6–11]. The morphology and repartition of austenitic and ferritic phases, as
Duplex stainless steels, based on their chromium, nickel, and mo- well as conditions of plastic working and annealing treatment, are
lybdenum content, can be sub-divided into standard duplex, super du- important factors which are known to control the final properties of
plex, hyper duplex, 25 Cr duplex, and lean duplex stainless steels duplex alloys [23]. The presence of α/γ interfaces affects considerably
grades. Lean duplex stainless steels have been developed to minimize the deformation mechanisms of these materials, due to strain in-
cost fluctuations by reducing expensive elements like Ni and Mo compatibilities at phase boundaries and to crystallographic slip in the
[6,8,12–14]. Because of its lower Ni and Mo contents than standard bcc and fcc structures [24–27]. Apart from the mechanism of slip or
2205 grades, 2304 steels are less sensitive to precipitation of inter- twinning within the γ-phase and the multiple slips within the α-phase,
metallic compounds, which may seriously deteriorate the corrosion and plastic deformation can proceed by strain induced γ → α transforma-
mechanical properties of such alloys [15–17]. As a ferrite former, tion. Such transformation can be described by favored orientation


Corresponding author.
E-mail address: paula-malta@hotmail.com (P.O. Malta).

https://doi.org/10.1016/j.matchar.2018.06.006
Received 24 January 2018; Received in revised form 11 May 2018; Accepted 3 June 2018
Available online 05 June 2018
1044-5803/ © 2018 Elsevier Inc. All rights reserved.
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

relationships (OR's) between specific planes and directions of the


phases, which allows the best fit of their interfaces [28]. The OR models
were firstly investigated by Bain [29], Kurdjumov and Sachs (K–S) [30],
Nishiyama and Wasserman (N–W) [31,32], Greninger and Troiano
(G–T) [33], and Pitsch [34], and recently by Headley and Brooks (H–B)
[35].
Since for most end applications duplex structures are manufactured
from hot-rolled and annealed sheets, their crystallographic texture re-
sults from the deformation, recrystallization and phase transformation
phenomena occurring in austenite and ferrite phases [23,36]. Re-
crystallization of deformed metals gives rise to changes in the dis-
tribution of crystallographic orientations by nucleation of new non-
deformed grains and their subsequent growth into the deformed matrix
[28,37]. The evolution of the recrystallization texture during annealing
of duplex stainless steels has been the subject of several research studies
[23,36,38–44]. In this sense, oriented nucleation (ON) and selected
growth (SG) theories were associated with the preferential formation of
specific orientations owing to either a frequency (ON) or size advantage
(SG) [45–49]. According to Lee and Han [50], the Goss grains that
persisted after rolling of fcc alloys seem to act as nuclei during the re-
crystallization and develop at the expense of surrounding brass grains.
In the case of bcc alloys, the lower stored energy of the 〈110〉 com-
ponents compared to those of the γ-fibers leads to a favored nucleation
of {111} grains [49,51,52].
Annealing textures of bcc and fcc materials have been reasonably
explained by the correlation between the SG theory and the CSL type-
boundaries. It was established that certain misorientations, namely
~40° 〈111〉 in fcc and 26.5° 〈110〉 in bcc, resulted in high mobility
interfaces that growth preferentially during the recrystallization
[50,53–55]. The superior mobility of ∑7 CSL boundaries, misoriented in
38.2° about 〈111〉, is central to the SG model of fcc recrystallization
texture and seems to be associated with the replacement of deformed S-
grains by recrystallized cube grains [53]. With respect to bcc steels,
Malta et al. [56] observed that during annealing the {554}〈225〉
components formed at the expense of {112}〈110〉 under a 26.5° 〈110〉
misorientation relationship, the same of Σ19a CSL boundaries.
The aim of present work was to evaluate the effect of molybdenum
content and annealing temperature on microstructure and texture
evolution by comparing the results obtained for a standard 2205 alloy
and a lean 2304 alloy. To understand more deeply the morphological
and crystallographic changes of annealed duplex steels, it is essential to
study all the mechanisms involved in their recrystallization kinetics.
Fig. 1. Phase fraction as a function of temperature calculated by using the
Therefore, grain nucleation and growth, as well as the distribution of
Thermo-Calc software for the (a) 2205 and (b) 2304 duplex stainless steels used
phases were evaluated from the initial processing condition. in the present work.

2. Materials and Methods appeared during the last hot rolling passes.
The phase fraction as a function of temperature was calculated by
The duplex stainless steels investigated in the present work, desig- using the Thermo-Calc software for the steels studied in this work, as
nated as 2205 and 2304 alloys, were received in the form of hot-rolled shown in Fig. 1. Due to the high number of alloying elements, both
sheets of 4 mm thick. The chemical compositions of the standard and steels showed a rather complex precipitation behavior. In duplex
lean grades are given in Table 1. stainless steels, the precipitation of intermetallic phases is of greater
The industrial sequence for obtaining duplex stainless steels by interest since they seriously deteriorate the corrosion and mechanical
rolling includes two annealing steps: one intermediate after the hot properties of such alloys [15,16]. The χ-phase was observed only in
rolling, and the final one after the cold rolling. In the present work, the 2205 alloy at a temperature range of 550–750 °C, while the σ-phase was
parameters, i.e. temperatures and times, of both annealing treatments formed in both 2205 (500–950 °C) and 2304 (500–750 °C) grades.
are selected to avoid the formation of precipitates. Furthermore, in the Specimens aged between 500 and 900 °C were subjected to the pre-
case of the intermediate annealing, the temperature was higher and cipitation of hexagonal nitrides (Cr2N), which could lead to a possible
soaking time longer to dissolve the precipitates which may have

Table 1
Chemical composition of 2205 and 2304 duplex stainless steels (in wt%).
C Cr Ni Mo Mn Si P Cu S N Fe

Standard 2205 0.023 22.43 5.44 3.046 1.82 0.249 0.03 0.191 0.0002 0.169 Bal.
Lean 2304 0.011 22.87 4.20 0.275 1.45 0.201 0.02 0.453 0.0004 0.101 Bal.

407
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

embrittlement of the material. Moreover, up to temperatures of 1400 °C microstructure and texture can be ignored under these conditions [63].
there was the precipitation of MnS, whose inclusions have been defined As shown in Fig. 3(a) and (b), cold rolling resulted in a substantial
as good initiation sites for pitting corrosion of stainless steels [57–59]. reduction in the phase spacing with a considerable increase in grain
Therefore, in the present work, temperatures below 900 °C were not flattening, especially of ferrite grains, from the hot- to the cold-rolled
used to avoid the formation of these detrimental secondary phases. condition. Both phases were plastically deformed in the rolling direc-
Samples with dimensions of 15 mm in length, 10.5 in width and tion and a strong grain refining was observed owing to the elongation
4 mm in thickness were solution annealed at 1100 °C for 1800 s and and break of grains during deformation at room temperature. The cold
then quenched in water. Plastic deformation of the solution annealed rolling step possibly involved also the formation of the strain-induced
material was carried out by cold rolling to a thickness of 1 mm on a α′-martensite (SIM) from the metastable austenite. This inference is
laboratory rolling mill. The cold-rolled specimens were isothermally assigned to the presence of some bcc structure inside austenite grains,
annealed for 180 s at the temperatures of 900 and 1050 °C followed by which implies that martensite had formed during plastic deformation.
water quenching. Although diffraction patterns of ferrite grains (bcc) can easily be dis-
The EBSD technique was used to determine the phase ratio, crys- tinguished from austenite grains (fcc), EBSD technique is incapable of
tallographic texture, interface character and substructure character- separating martensite regions from the ferrite grains. In this context, the
istics. Specimens for EBSD were prepared by classical procedures of sharpness of image quality (IQ) maps may be a key to differentiate the
mechanical grinding and polishing. The final polishing was performed bcc phases since martensite has greater lattice distortion and hence a
with colloidal silica suspension using a Buehler Minimet 1000® auto- more unfocused image compared to ferrite.
matic machine. The EBSD scans were carried out using a FEI InspectTM Higher magnification images of Fig. 3(a) and (b) are shown in
S50 SEM by an accelerating voltage of 20 kV and a working distance Fig. 3(c) and (d), in which the image quality was averaged over each
close to 20.0 μm. A step size of 0.2 μm was used to scan an area of grain and a threshold value was defined. Grains below this threshold
approximately 45 × 120 μm. The raw data were post-processed using value, with poor diffraction, were distinguished as martensite from
EDAX OIM Analysis™ software after a 5–10% grain dilation clean-up. ferrite regions. Besides, the boundaries between the surrounding mar-
Microstructural investigations were carried out by image quality tensite and the retained austenite are often described in terms of K–S
(IQ) maps, which provided information about grain size and shape, as (90° 〈112〉) and N–W (95.3° 〈362〉) relationships [7,64,65], as shown
well as phase partition and twinning. The ODFs were calculated using in the right-side images in Fig. 3(c) and (d). In this sense, regions dis-
the discrete binning method of Bunge [60] with both bin size and played in blue were identified as martensite phase, since they presented
Gaussian smoothing of 5°. Orthotropic sample symmetry for rolled both low IQ and OR boundaries typical of martensite-austenite inter-
specimens was assumed in calculating the ODFs. CSL boundaries were faces. Further evidence of the correct separation of bcc structures can be
identified following the Brandon's criterion [61] for maximum per- given by the differences in grain size between the two phases, so that
missible deviation (i.e., Δθ = θ0 (∑)−n degrees, with θ0 = 15° and ferrite presented larger grains and martensite was characterized by
n = 0.5). Only the fraction of boundaries up to ∑29 was calculated since smaller grains [64]. The grains defined as belonging to the martensite
boundaries above this value are considered irrelevant in polycrystals. phase have been developed from the austenite phase, which corrobo-
First order twin boundaries (TBs) were defined as ∑3 (60° 〈111〉). rates the strain-induced γ → α′ transformation. The cold-rolled material
was mainly composed of ferrite (bcc) phase, especially the 2304 sam-
3. Results ples. Increasing the Mo content slightly suppressed the full formation of
the ferrite phase after rolling, possibly due to its effect on increasing the
3.1. Microstructural analysis strength of the structure [62] and, hence, reducing the degree of de-
formation and flattening of the grains. The uneven partitioning of
Fig. 2 shows the microstructure of the hot-rolled (Fig. 2(a) and (b)) phases may also be attributed to the replacement of austenite grains by
and solution-annealed (Fig. 2(c) and (d)) samples. In the hot-rolled plastically induced martensite. Moreover, austenite is commonly con-
condition, both phases had a partially work hardened microstructure sidered as a second phase in duplex steels, so that its deformation is
and developed a typical band-like arrangement consisting of alternating restricted to an area limited by the surrounding ferrite.
bands of austenite and ferrite aligned parallel to the rolling direction In Fig. 4 the microstructure evolution during the final annealing at
(RD). The elongated shape of the grains was more pronounced in the 900 and 1100 °C is presented. Annealing at 900 °C (Fig. 4(a) and (b))
2205 samples (Fig. 2(a)) compared to 2304 samples (Fig. 2(b)). This promoted the primary recrystallization of the austenite phase, com-
elongation probably resulted either from the non-homogeneous strain prising a fine grain structure with only some remains of deformed
distribution or from the intrinsic instability of the constituents during austenite, mainly in the 2304 samples. Increasing the Mo content had a
plastic deformation. Intermediate annealing promoted the development pronounced effect on the grain growth phenomenon of the austenite
of a deformation-free microstructure with austenitic islands forming phase. At 900 °C, the austenite grains of the 2205 samples showed a
from a coarse ferritic matrix. Besides, the recovery of ferrite phase (bcc) significantly smaller size when compared to the austenite grains of the
and recrystallization of austenite phase (fcc) appeared to be sufficient. 2304 samples. This may be related to the presence of Mo atoms at the
The microstructure of 2205 samples was formed by roughly equiaxed grain boundaries, which anchored the boundaries and reduced their
grains, especially the austenite phase (Fig. 2(c)). For 2304 samples mobility. At high annealing temperatures, the Mo retarding effect be-
(Fig. 2(d)), both phases presented a pancake structure with the grains came weak, so that an equiaxial austenite phase was achieved with the
remaining elongated along RD and with a clearly larger size than those grains of both samples having similar sizes and shapes. A bamboo-like
of 2205 steels. Both hot-rolled and solution-annealed materials ex- morphology was developed in the ferrite phase after the annealing at
hibited a well-balanced microstructure, composed of a slightly larger 900 °C, with the grains retaining the elongated shape of the cold-rolled
volume fraction of ferrite than of austenite. The effects of Mo partition condition. The retention of lamellar structure indicates that an ex-
did not seem to influence, therefore, the distribution of phase fractions tensive recovery took place in this constituent and that its re-
under these conditions. In contrast, the segregation and anchoring crystallization was not enough during the annealing at a low tem-
mechanisms of this element promoted a greater impact on the growth of perature. Increasing annealing temperature to 1100 °C promoted the
ferrite and austenite grains after the solution annealing. According to transformation of the bamboo structure to a more globular morphology.
Wang et al. [62], increasing the Mo content can prevent the growth of The breakdown of the flat layered structure into smaller pieces was
austenite and hence refine the steel structure. Since austenite fraction initiated by mutual inter-penetration of the phases along the grain
have remained stable during hot rolling and annealing, the effects of junctions. The triple junctions, marked by circles in Fig. 4, consist of
any stress-induced phase transformation on the development of the one grain boundary and two-phase boundaries [36,41,44]. Considering

408
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

RD

ND

Total Total
Phase Fracon Phase Fracon
Ferrite 0.539 Ferrite 0.535
Austenite 0.461 Austenite 0.465
(a) (b)

Total Total
Phase Fracon Phase Fracon
Ferrite 0.549 Ferrite 0.558
Austenite 0.451 Austenite 0.442

(c) (d)
Fig. 2. IQ combined with phase maps for the (a,b) hot-rolled and (c,d) solution-annealed (a,c) 2205 and (b,d) 2304 samples.

that the diffusion is much faster in the α-grains than in the γ-grains [3], precipitates were found in the annealed specimens.
it is expected that the recrystallization kinetic and the grain growth
occur earlier in the ferrite than in the austenite. In addition to this
coalescence of the ferrite grains, the increase in annealing temperature 3.2. Microtexture Analysis
intensified the formation of annealing twins in the austenite grains
(Fig. 4(b) and (c)). At 900 °C, the austenite phase reached a larger vo- Fig. 5 show the φ2 = 0, 45 and 65° sections of the orientation dis-
lume fraction than the ferrite phase because of the heat-induced re- tribution function (ODF) of austenite grains for the hot-rolled (Fig. 5(a)
version of martensite, generated in cold rolling, to austenite. Mean- and (b)) and solution-annealed (Fig. 5(c) and (d)) samples. The auste-
while, there was an increase of ferrite fraction with increasing nite texture of 2205 and 2304 hot-rolled steels showed strong S {123}
temperature. In agreement with the 1:1 phase ratio reached at 1100 °C 〈634〉γ and brass {110} 〈112〉γ components, the latter being more in-
by the Thermo-Calc calculations (Fig. 1), a nearly equal volume fraction tense in 2205 samples. A brass texture spread was observed along the
of the phases was observed in Fig. 4(c) and (d). Besides, no visible G/B {110}〈115〉γ to the Goss {110}〈001〉γ orientation in the 2304 hot-
rolled samples and in both solution-annealed samples (Fig. 5(b)–(d)).

409
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

Fig. 3. IQ combined with phase maps for


RD
the cold-rolled (a) 2205 and (b) 2304
samples; (c) and (d) higher magnification
images of Fig. 3(a) and (b), respectively,
displaying, in the right-side images, the
ND special axis-angle phase boundaries as fol-
lows: K–S (90° 〈112〉) in yellow and N–W
(95.3° 〈362〉) in pink with a tolerance of
2.5°. (Separation of ferrite and martensite
was based on grain average IQ parameter;
image quality property was poorer for
martensite than for ferrite; grain dilatation
cleanup was applied before the acquisition
of Fig. 3(c) and (d)). (For interpretation of
the references to colour in this figure le-
gend, the reader is referred to the web
version of this article.)

Total Total
Phase Fracon Phase Fracon
Ferrite 0.769 Ferrite 0.892
Austenite 0.230 Austenite 0.108
(a) (b)

Phase Boundary Phase Boundary


Ferrite K–S Ferrite K–S
Austenite N–W Austenite N–W
Martensite Martensite
(c) (d)

Lean steel in the hot-rolled condition also had a stronger copper {112} An extension of α-fiber in the direction of the {111} components was
〈111〉γ component compared to 2205 steel. Textures originated through further observed in the 2304 hot-rolled steel with a declension of γ-fiber
the solution annealing process are known as annealing textures, which from {111}〈011〉α to {554}〈225〉α component. After the annealing,
depend on the prior hot rolling textures but generally differ from them. there was a strengthening of the {112}〈110〉α orientation in the lean
Annealing the alloys after hot rolling shifted the copper orientation by alloy (Fig. 6(d)) and the displacement of the γ-fiber became more evi-
about 7° along Φ generating the Dillamore {4 4 11}〈11 11 8〉γ com- dent. On the other hand, the 2205 samples did not reach a complete
ponent, a twin of the Goss orientation. Except for this displacement, the development of the γ-fiber, even after the annealing (Fig. 6(c)), possibly
solution-annealed texture of 2304 samples remained as its hot-rolled due to their higher Mo content compared to 2304 samples. In the bcc
texture. In both conditions, a propagation of {110}〈001〉γ orientation structures, grains with {111} orientation are known to have a greater
along the 〈100〉-fiber towards the cube components was observed in capacity to store energy during plastic deformation than grains with
the lean steel. Goss texture, absent in the 2205 hot-rolled alloy 〈110〉 orientation [37]. As this stored energy acts as a driving force for
(Fig. 5(a)), was displayed after the annealing at 1100 °C for 30 min recrystallization and grain growth, the increase in Mo content was re-
(Fig. 5(c)). The different Mo contents of the two samples did not sig- sponsible for suppression of the γ-fiber evolution and, hence, for the
nificantly influence the development of the austenite texture during the inhibition of the ferrite grain growth observed in Fig. 2.
hot rolling and solution annealing. Fig. 7 shows the texture of austenite phase of the cold-rolled sam-
Fig. 6 shows the ferrite texture of hot-rolled (Fig. 6(a) and (b)) and ples. Cold rolling promoted a reduction in the texture intensity of
solution-annealed (Fig. 6(c) and (d)) specimens. In ODF representation austenite grains in relation to the hot-rolled and solution-annealed
the ferrite phase exhibited stronger texture intensities than austenite conditions (Fig. 5), especially of the 2205 samples. The brass and Goss
phase. Hot-rolled 2205 samples (Fig. 6(a)) revealed a strong cube tex- components were slightly more intense in the 2304 alloy than in the
ture, which became weaker in the solution-annealed condition 2205 alloy. The propagation of {110}〈001〉γ component towards the
(Fig. 6(c)) due to a slight propagation of this component to the {112} 〈100〉 direction, previously noted in the lean steel (Fig. 5 (b) and (d)),
〈110〉α and {111}〈121〉α orientations. In the case of 2304 hot-rolled was observed in the 2205 cold-rolled samples (Fig. 7(a)). Moreover,
samples (Fig. 6(b)) the ferrite grains showed a complete α-phase of such steel displayed a weak {111}〈110〉γ component accompanied by a
strong intensity from the {112}〈110〉α to the {111}〈110〉α orientation. spread orientation along the γ-fiber. The rolling texture of 2304 alloy

410
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

RD

ND

Total Total
Phase Fracon Phase Fracon
Ferrite 0.462 Ferrite 0.395
Austenite 0.548 Austenite 0.605
(a) (b)

Total Total
Phase Fracon Phase Fracon
Ferrite 0.521 Ferrite 0.519
Austenite 0.479 Austenite 0.481
(c) (d)
Fig. 4. IQ combined with phase maps for the (a,c) 2205 and (b,d) 2304 samples annealed at (a,b) 900 and (c,d) 1100 °C for 180 s.

(Fig. 7(b)) can be interpreted as being essentially a brass-type texture brass and Goss components comprising the intermediate G/B orienta-
with a scattering towards the Goss-type texture. Both steels also pre- tion. The annealing at 1100 °C promoted the strengthening of all
sented a low intensity of S-orientation. components of both alloys compared to the 900 °C-annealed condition.
Fig. 8 shows the φ2 = 45° section of the ODF of ferrite grains for the Summing up, the austenite texture of duplex stainless steels after an-
cold-rolled 2205 (Fig. 8(a)) and 2304 (Fig. 8(b)) samples. As in the hot nealing consisted primarily of the brass, Goss, Dillamore and S-or-
rolling condition, the 2205 cold-rolled specimens had strong cube ientations.
components that propagated to the {112}〈110〉α orientation. The α- Fig. 10 shows the texture of ferritic grains of the duplex alloys after
and γfibers, observed under hot rolling and annealing conditions final annealing. Increasing temperature reinforced the ferrite texture,
(Fig. 6(b) and (d)), presented lower intensity in the 2304 cold-rolled mainly of the components related to the γ-fiber and especially in the
steel (Fig. 8(b)). 2304 specimens. The development of γ-fiber was only initiated in the
Final annealing textures are shown in Fig. 9 for the austenite grains. steels annealed at 1100 °C (Fig. 10(c) and (d)), with no complete for-
Both specimens annealed at 900 (Fig. 9(a) and (b)) and 1100 °C mation of that fiber as observed in Fig. 6(b) and (d). The {112}〈110〉α
(Fig. 9(c) and (d)) presented similar austenitic textures, except for the and cube orientations were found in both alloys and annealing tem-
absence of the Cu and D-component in the 2205 samples annealed at peratures, being more intense in the 2304 samples annealed at 1100 °C
900 °C (Fig. 9(a)). The austenite phase was composed mainly of the (Fig. 10(d)). As in the solution-annealing condition (Fig. 6(c) and (d)),

411
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

1 (0.0°-90.0°)

0° 45° 65°
{110}<112>

(0.0°-90.0°) //<100>

{110}<001>

{4 4 11}<11 11 8>

(a) {112}<111>

0° 45° 65° {110}<115>

{123}<634>

(b)

0° 45° 65°

(c)

0° 45° 65°

(d)
Fig. 5. ODF of austenite phase for the (a,b) hot-rolled and (c,d) solution-annealed (a,c) 2205 and (b,d) 2304 samples.

the {111}〈112〉α component was observed in the 2205 alloy (Fig. 10(a) were quantified (Fig. 11(d) and (e)). A smaller fraction of special
and (c)) and was replaced by the {554}〈225〉α orientation in the 2304 boundaries was observed in the 2205 alloy in relation to the lean alloy.
alloy (Fig. 10(d)). Molybdenum atoms appeared to inhibit the dis- Intermediate and final annealing strengthened the CSL formation in the
placement of the γ-fiber and, hence, the formation of the {554}〈225〉α ferrite phase of 2304 specimens, except for the ∑19a boundaries in the
orientation. 900 °C-annealed condition.

3.3. Grain Boundary Character 4. Discussion

The numeric fraction of CSL boundaries up to ∑29 is shown in 4.1. Microstructure Evolution
Fig. 11(a) and (b). Some boundaries were separately calculated to allow
their individual distribution in each constituent. For the γ-phase, the It is known that the ferrite and austenite phases have their own
∑3, ∑7, and ∑9 boundaries were considered (Fig. 11(c) and (d)). In both characteristic flow stress level, due mainly to their different crystal-
steels, only the ∑3 boundary varied significantly from one condition to lographic structures and softening mechanisms. The partitioning of al-
the other. After solution annealing, the 2205 samples had a higher loying elements, especially the nitrogen and carbon to austenite, con-
density of ∑3 boundary compared to the 2304 samples, while after final siderably strengthens this phase and increases the flow stress gradients.
annealing at 1100 °C, both samples had a similar fraction of this CSL Such differences in hot strength are responsible for creating an uneven
boundary. Regarding the α-phase the ∑3, ∑13b, and ∑19a boundaries strain partitioning between the two phases, with the harder austenite

412
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

1 (0.0°-90.0°)
{001}<110>
45° 45°
{112}<110>

(0.0°-90.0°) {111}<110>

{111}<011>

{554}<225>

{111}<121>
(a) (b) {111}<112>

45° 45° -fiber

(c) (d)
Fig. 6. ODF of ferrite phase for the (a,b) hot-rolled and (c,d) solution-annealed (a,c) 2205 and (b,d) 2304 samples.

offering a greater resistance to deformation than the softer ferrite accommodation of most of the deformation by the α-grains possibly
[63,66,67]. As both phases coexisted in large volume fractions during resulted in a faster dynamic recovery of the ferrite compared to the
the hot rolling, it is expected that their deformation behavior was in- dynamic recrystallization of the austenite. Tsuzaki et al. [68], in their
fluenced by the presence of the other phase. In addition, since the strain studies on superplasticity of a duplex stainless steel, indicated that the
energy is the driving force for softening processes to occur, the role of dynamic recrystallization of austenitic grains is to make the fine

1 (0.0°-90.0°)
0° 45° 65° {110}<112>

//<100>
(0.0°-90.0°)
{110}<001>

{111}<110>

{110}<115>

{123}<634>
(a)
-fiber
0° 45° 65°

(b)
Fig. 7. ODF of austenite phase for the cold-rolled (a) 2205 and (b) 2304 samples.

413
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

(0.0°- [79], in their studies about an austenitic stainless steel, reported that
1
the K-S and N-W relationships described very well the crystallographic
45° orientation relations between austenite and martensite phases during
the cold rolling.
{001}<110> The effect of annealing temperature on the microstructure and
properties of duplex alloys has been widely discussed [10,80–85]. The
(0.0°-90.0°)
{112}<110> microstructural changes that occur during annealing are based on two
fundamental recrystallization mechanisms, namely: the formation of
nuclei in specific regions of the work-hardened material and the sub-
{111}<011>
sequent growth of these nuclei from the deformed regions [86]. During
the cold rolling stage, a larger number of dislocations were generated
-fiber near the grain boundaries and between the phase interfaces, thereby
resulting in optimal conditions for nucleation and growth in these re-
gions. It was clear that the band-shaped ferrite grains became wider and
the island-shaped austenite grains grew larger with the enhancement of
(a)
annealing temperature. Besides that, the thermally activated grain
boundary migration took place during the annealing at 1100 °C, re-
45° sulting in the formation of annealing twins in the austenite phase. In the
present work, the austenite fraction decreased with increasing tem-
perature due to the diffusion-controlled γ → α transformation, typical
of high annealing temperatures [80]. Tan et al. [81] reported that, since
raising the annealing temperature resulted in a higher ferrite fraction,
the original α-forming elements (Cr and Mo) were spread over a larger
volume and diluted, leading to a decrease in their concentrations in the
ferrite phase. In another front, Naghizadeh and Mirzadeh [77] reported
that the grain growth of austenite phase was severely retarded in AISI
316 stainless steel containing about 2.0 Mo wt%. Suppression of grain
growth is a phenomenon highly sensitive to temperature, being more
pronounced at low annealing temperatures. This predicts that the
(b) concentration of solute atoms (e.g., Mo) in the boundaries decreases
with increasing temperature, so that the recrystallization and growth of
Fig. 8. ODF of ferrite phase for the cold-rolled (a) 2205 and (b) 2304 samples. the austenite grains occur easily due to the free movement of the
boundaries.
structure suitable for sliding the grain boundary in the early stages of
plastic deformation. After hot rolling, the microstructure of duplex 4.2. Texture and Grain Boundary Evolution
steels assumed a linear planar configuration with a fibrous appearance
throughout the longitudinal section. Besides that, some annealing twins The crystallographic behavior of duplex alloys results from the de-
were observed in austenite phase, especially of the 2205 samples formation, recrystallization and phase transformation phenomena oc-
(Fig. 2(a)). The low stacking fault energy (SFE) of the austenite phase is curring in austenite and ferrite at each processing stage. Several in-
known to be the main reason for the formation of twinning during hot vestigations have been argued that the two phases deform
rolling, so that the fraction of twins increases with the amount of dy- independently and that their texture evolution can be inferred from the
namic-recrystallized grains [69–73]. Intermediate annealing further behavior of single-phase austenitic and ferritic stainless steels [23, 36].
increased the formation of twins in the γ-phase because of the stacking The texture components of deformed austenite produced by hot rolling
errors generated by migrating grain boundaries during recrystalliza- were described by Hutchinson et al. [87] and Watershoot et al. [88] as
tion. Jin et al. [71] in their studies about a 304 L austenitic stainless the β-fiber stretching from the copper orientation and via S- to the brass
steel observed that the higher the prior strain level on annealing twin, orientation. Properly, these texture components were observed in both
the higher the velocity of grain boundary migration and the higher the hot-rolled specimens (Fig. 5 (a) and (b)). Based on the K–S relationships
annealing twin density in the recrystallized grains. between the parent austenite phase and its transformation products, the
During cold rolling, the original grains became more elongated β-fiber components of austenite can transform into α- and γ-fibers
along RD and, consequently, received a larger grain refinement effect. components of ferrite [88]. In this sense, the {112}〈111〉γ transfor-
Plastic deformation led also to the replacement of a fraction of me- mation in austenite phase was probably responsible for the appearance
tastable austenite by deformation-induced martensite. The predisposi- of the {112}〈110〉α orientation in ferrite phase of both hot-rolled
tion to γ → α′ transformation is strongly related to the stability of the samples (Fig. 6(a) and (b)). Resulting from the brass transformation
austenite phase and, hence, increases with the decrease of its SFE during hot rolling, the {001}〈110〉α and {111}〈110〉α components was
[4,74–76]. According to Naghizadeh and Mirzadeh [77], increasing the observed in the 2205 (Fig. 6(a)) and 2304 (Fig. 6(b)) specimens, re-
content of alloying elements generally increases the austenite stability spectively.
with respect to strain-induced transformation. The higher Mo content of Annealing at high temperatures often promotes the microstructure
the 2205 alloy than the 2304 alloy was therefore the likely reason for its recrystallization, which may lead to an absence of texture, to the de-
greater resistance to this transformation. Tavares et al. [78] defined velopment of intense texture components, or even to no modification in
that the lean duplex steel 2304 is much more susceptible to martensitic the deformation texture [89]. Despite the changes in the micro-
induced transformation than duplex 2205 because of its lower Ni and structure, the austenite phase of the solution-annealed samples pre-
Mo contents. Such transformation was defined in Fig. 3(c) and (d) by sented similar orientations from those of the hot-deformed material.
the K–S and N–W relationships, which are the most frequently reported The Goss texture in the 2205 alloy (Fig. 5(c)) and the Dillamore com-
relationships for bcc-fcc systems and are the only ones that have been ponent in the 2304 alloy (Fig. 5(d)) were the only changes observed.
reported for ferrite-austenite microstructures [35]. Kurc-Lisiecka et al. According to Barella and Mapelli [90], the Goss and Dillamore com-
ponents are probably produced by the twinning of the austenite phase

414
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

1 (0.0°-90.0°)

0° 45° 65°
{110}<112>

(0.0°-90.0°) {110}<001>

{4 4 11}<11 11 8>

{112}<111>

(a) {110}<115>

45° {123}<634>
0° 65°

(b)

0° 45° 65°

(c)

0° 45° 65°

(d)
Fig. 9. ODF of austenite phase for the (a,c) 2205 and (b,d) 2304 samples annealed at (a,b) 900 and (c,d) 1100 °C for 180 s.

after the solution annealing. In ferrite phase, the texture changes were orientations with a common 〈110〉 direction parallel to the sheet
more significant, with the weakening of the cube component in the normal [91]. Therefore, the standard duplex steel was deformed by
standard alloy (Fig. 6(c)) and the strengthening of the {112}〈110〉α both twinning and slip mechanisms, while in the lean steel (Fig. 7(b))
orientation and of the displacement of γ-fiber in the lean alloy the twinning was the predominant deformation mechanism for auste-
(Fig. 6(d)). Such displacement of the γ-fiber, from {111}〈112〉α to nite phase. Jia et al. [92] studied the texture evolution of a 2507 duplex
{554}〈225〉α, was identified by Malta et al. [52,56] in the re- stainless steel after cold rolling over a range of 25–85% of thickness
crystallization texture of a ferritic stainless steel. The authors attributed reduction. At the highest reduction, very strong brass and Goss com-
the presence of this latter component to the mechanism of selected ponents were identified, i.e., the typical ones found in the deformed fcc
growth of ferritic grains. materials with low SFE [63,91]. Ryś and Zielińska-Lipiec [93] observed
Plastic deformation occurs in metals crystals by slip or twinning that the {110}〈001〉γ component is a relatively stable orientation in
systems, which comprises certain atomic planes and specific crystal- cold-rolled duplex steels, in cases where it appears as the strongest
lographic directions. In bcc metals, the slip direction is always (111)α austenite texture after the preliminary thermomechanical treatment.
and the slip planes are mainly {110}α and {112}α, while the fcc metals For the ferrite phase, the cold-rolled texture of both samples (Fig. 8(a)
have the slip systems in {111}(110)γ and the mechanical twinning and (b)) was very similar to that of the hot rolling condition, so that the
systems in {111}(112)γ [86]. In the present work, the austenite phase of {001}〈110〉α, {112}〈110〉α and {111}〈110〉α components may have
the 2205 cold-rolled specimens (Fig. 7(a)) presented, in addition to the been inherited from the previous thermomechanical processing. It is
tube-shaped β-fiber, the so-called α-fiber which is formed by suggested, therefore, that there was a significant influence of the initial

415
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

1 (0.0°-90.0°)
45° 45°
{001}<110>

{112}<110>
(0.0°-90.0°)
{111}<121>

{111}<112>

{554}<225>

(a) (b) -fiber

45° 45°

(c) (d)
Fig. 10. ODF of ferrite phase for the (a,c) 2205 and (b,d) 2304 samples annealed at (a,b) 900 and (c,d) 1100 °C for 180 s.

grain orientation on the texture evolution of the duplex steels in the except for the noticeable strengthening of the copper and Dillamore
subsequent mechanical process. Raabe and Lücke [51] evaluated the components, especially in the 2205 alloy. Since the {4 4 11}〈11 11 8〉γ
cold deformation texture of single-phase ferritic stainless steels and orientation is a first order twin of the deformation texture component
found that high levels of reduction lead to the strengthening of the α- {110}〈001〉γ [44], its increased intensity appeared to result from the
fiber, with maxima occurring at {001}〈110〉α and {112}〈110〉α. This is intense twinning during recrystallization favored by the low SFE energy
consistent with the results obtained for the 2205 specimens as shown in of austenite phase. Additionally, the displacement of the local max-
Fig. 8(a). imum intensity from the Cu orientation towards the D-orientation was
After annealing, the austenitic texture of 2304 samples (Fig. 9(b) also observed by Blicharski et al. [42] during the annealing of duplex
and (d)) seemed to be a random sampling of its cold rolling texture with steel.
the nucleation of grains from pre-existing orientations in the deformed The gradual transition from a lamellar morphology (Fig. 4(a) and
matrix, i.e. the brass, Goss, and S-orientations. Retention of deforma- (b)) towards a comparatively more globular morphology (Fig. 4(c) and
tion texture during recrystallization has been attributed either to a (d)) with increasing annealing temperature was indicative of the re-
discontinuous recrystallization without preferential orientation selec- covery in the ferrite phase [41] and affected the evolution of re-
tion [36,40,41] or to a suppressed growth selection, e.g. by a low grain crystallization texture in this constituent. In single-phase ferritic steels,
boundary mobility [44,94]. Discontinuous recrystallization and growth the typical rolling orientations {001}〈110〉α and {112}〈110〉α mainly
were also supported by the presence of annealing twins in austenite undergo recovery while the {111}〈112〉α and {111}〈121〉α orienta-
[44]. Lindell [95] investigated the texture evolution of warm-rolled and tions preferentially recrystallize during annealing [36,41,44]. It is well
annealed 304 L (0.5 Mo wt%) and 316 L (2.3 Mo wt%) austenitic stain- known that the stored energy of the α-fiber components is lower than
less steels. The rolling texture was maintained in the 304 L alloy, which that of the γ-fiber components, which makes the {111} nuclei faster in
was attributed by the author to differences in grain boundary mobility forming than those of 〈110〉 orientations. Since the {112}〈110〉α de-
by solute drag effects, mainly from molybdenum. On the other hand, formed grains are preferentially consumed by the {111}〈112〉α re-
the annealing textures of 2205 samples (Fig. 9(b) and (d)) differed crystallized grains [49,96], their weakening from the cold-rolled
strongly from their cold rolling textures. The disappearance of the (Fig. 8) to the annealed condition (Fig. 10) proves the favored nu-
{111} component was correlated by Donadille et al. [94] to oriented cleation of the γ-fiber grains during recrystallization. In this context, the
nucleation in shear bands of {110}〈001〉γ grains invading the {111} annealing textures of the 2205 samples (Fig. 10(a) and (c)) can be
〈uvw〉γ domains in which no nucleation was observed. The annealing therefore explained in terms of the ON theory by the joint occurrence of
textures of both alloys did not show the brass recrystallization (BR) the recovery (retention of α-fiber) and recrystallization (development
orientation {236}〈385〉γ, which is typically developed during re- of γ-fiber) mechanisms. While only recovery was observed in the 2304
crystallization of single-phase austenitic steels with low SFE [40,44,94]. samples annealed at 900 °C (Fig. 10(c)), increasing the temperature
Since the BR-oriented grains come from rotations of 40° 〈111〉 with developed further recrystallization behavior with the displacement of
respect to the brass (rolling) orientation, their boundaries result in high the {111}〈112〉α orientation by 5° along Φ (Fig. 10(d)). Previously
mobility interfaces that growth preferentially during recrystallization observed in the annealing texture of bcc steels [46,50,52,56,96,97],
[28,44,50,53–55,94]. Apparently, the inhibition of shear bands for- such displacement gave rise to the {554}〈225〉α component, whose
mation in the deformed matrix was the main reason for not developing grains could (in principle) participate in the selective growth of their
the BR orientation. Increasing annealing temperature did not sig- nuclei. Yantaç et al. [97] affirmed that subgrains with {554}〈225〉α
nificantly alter the recrystallization texture of the austenite phase, orientations have a high nucleation probability, which ensures an early

416
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

Fig. 11. Numeric fraction of (a,b) CSL boundaries up to ∑29, (c,d) ∑3, ∑7, and ∑9 in austenite phase and (e,f) ∑3, ∑13b, and ∑19a in ferrite phase of (a.c,e) 2205 and
(b,d,f) 2304 samples. (HR – hot-rolled; SA – solution-annealed; CR – cold-rolled; 900 °C – annealed at 900 °C for 180 s; 1100 °C - annealed at 1100 °C for 180 s).

size advantage and a continuous growth during the annealing cycle. heat treatments was closely associated with the formation of annealing
Kestens and Jonas [98] stipulated, however, that the growth advantage twins resulting from a thermally activated grain boundary migration
of a specific component is attributed only to the selection mechanism so [99,100]. Fig. 12 corresponds to higher magnification images of
that the first formed nuclei do not necessarily grow faster than the Fig. 2(c) and (d), highlighting the distribution and fraction of Σ3
others. This latter condition seemed to be in greater agreement with boundaries (60° 〈111〉 misorientation relationship) in the austenitic
texture formation of the 2304 samples since the {554}〈225〉α orienta- grains. During the recrystallization development, the enhanced mobi-
tion grow preferentially in bcc steels due to an enhanced interface lity of Σ3 boundaries usually increases the twin density, whereas the
mobility [46], which was only achieved during annealing at superior subsequent grain growth is generally accompanied by an apparent de-
temperatures. crease of the density of these first-order twin boundaries [100,101].
Although the Mo atoms directly influenced the growth of austenite Therefore, the marked growth of the austenitic grains, clearly observed
grains, differences in the Mo content among the samples did not sig- in Fig. 12(b), was possibly the main reason for the smaller fraction of
nificantly alter the development of the austenite texture in the different the Σ3 boundaries in the 2304 samples (Fig. 11(b)) compared to 2205
conditions evaluated. Regarding the ferrite phase, the increase in Mo samples (Fig. 11(a)) after solution-annealing. Tikhonova et al. [102]
content was responsible for a reduction of the favorable conditions for suggested that the annealing twins are generated mostly during the
recrystallization and grain growth, leading to the suppression of the γ- beginning of recrystallization regime and that their development slows
fiber development and the selective growth effect in 2205 samples. down as the recrystallized grains coarsen. In addition, the limited and
The high mobility of certain grain boundaries has been directly slow grain growth in 2205 steel could result from the segregation of Mo
associated with the development of specific CSL type-boundaries during atoms to the austenite moving boundaries [103], further inducing the
processing of metallic materials [50,53–56]. In austenite phase dragging of the solute due to its large atomic mismatch with the aus-
(Fig. 11(a) and (b)), the increasing intensity of Σ3 boundaries after the tenite matrix [104]. The low intensity of ∑9 boundaries discarded the

417
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

CSL Total CSL Total


Boundary Fracon Boundary Fracon
∑3 0.386 ∑3 0.285

(a) (b)
Fig. 12. Higher magnification images of (a) Fig. 2(c) and (b) Fig. 2(d) showing the distribution and fraction of ∑3 boundaries (60° 〈111〉) in the austenitic grains of
solution-annealed (a) 2205 and (b) 2304 samples.

occurrence of multiple twinning events, which are believed to be pro- samples compared to 2205 samples in the same condition. Grain
moted by strain-induced boundary migration [105]. The negligible coarsening resulted in a larger fraction of the ∑3, ∑13b, and especially
fraction of the ∑7 boundaries, which is believed to be dominant for the of ∑19a boundaries in ferrite phase (Fig. 11(f)) as opposed to austenite,
development of the SG mechanism in fcc recrystallization textures [53], where such coalescence appeared to inhibit the CSL formation.
corroborated the absence of preferential growth of specific γ-grains
during annealing. 5. Conclusions
In bcc steels, the development of special misorientations relation-
ships has been reasonably related to the favored nucleation and growth (i) The band-like morphology of the hot-rolled specimens was re-
of specific orientations during the recrystallization process [46,52,56]. placed by a pancake structure after solution annealing. Annealing
The apparent numerical disparity between the CSL boundaries in the twin development was especially marked in the austenite phase of
ferrite phase of the heat-treated steels (Fig. 11(a) and (b)) could also be solution-annealed samples. Despite the changes in the micro-
associated with their distinct Mo contents. Since molybdenum usually structure, the crystallographic texture of this phase was similar to
diffuses from ferrite to austenite in the absence of carbide formation the hot rolling and solution annealing. In the ferrite phase of lean
[106], there would probably be a depletion in its concentration in the alloy, the long-time annealing promoted the displacement of γ-
ferrite near the interface during annealing. In the standard steel, the fiber from {111}〈112〉α to {554}〈225〉α owing to the SG me-
superior Mo content may have increased the depletion rate of the ferrite chanism.
interfaces and generated unfavorable conditions for the formation of (ii) Cold rolling promoted a reduction in the phase spacing accom-
special grain boundaries. Yan et al. [106], in their studies on the panied by a strong grain refining. The γ → α′ transformation
crystallographic texture of single-phase ferritic steels, proposed that the during plastic deformation was based on the K–S and N–W or-
∑3 boundaries have low energy and mobility even at high temperatures. ientation relationships. The cold-rolled texture of austenitic grains
In this sense, its formation in the α-phase is possible only by the was governed by the α-fiber and the tube-shaped β-fiber. Twinning
combination of the driving force coming from grain rotation with the was the predominant deformation mechanism in the austenite
microstructural changes caused by plastic deformation and annealing. phase of the 2304 steel, while the 2205 steel was deformed by both
The authors established also a near misorientation relationship between twinning and slip mechanisms. For the ferrite phase, the de-
the ∑13b boundaries (27.8° 〈111〉) and the main γ-fiber components, formation texture of both samples was inherited from hot rolling
namely the {111}〈110〉α and {111}〈112〉α, which are related by a condition.
rotation of 30° 〈111〉 [104]. Additionally, the 26.5° 〈110〉 mis- (iii) After annealing at 900 °C, the austenite microstructure was formed
orientation relation, referring to the ∑19a boundaries, was closely re- by fine grains, which coarsened with increasing annealing tem-
lated to the SG mechanism in the final stages of recrystallization of the perature to 1100 °C. Extensive recovery took place in the ferritic
ferrite phase. As mentioned earlier, grains with {554}〈225〉α orienta- grains and recrystallization was not enough during the annealing
tion were developed after annealing at 1100 °C for 30 (Fig. 6(d)) and at 900 °C. Increasing annealing temperature did not significantly
3 min (Fig. 10(d)) due possibly to a preferential growth of their nuclei alter the recrystallization texture of the γ-phase in both steels.
in the ferritic matrix. Several authors [46,52,56,96,97,107,108] corre- Discontinuous recrystallization was responsible for the develop-
lated the growth advantage of {554}〈225〉α grains with either the de- ment of the austenitic texture in the 2304 samples after final an-
crease or the disappearing of the {112}〈110〉α deformation orientation nealing with the retention of deformation texture and formation of
through the 26.5° 〈110〉 rapid growth relationship. According to Ver- annealing twins. In the 2205 and 2304 alloys, the annealing tex-
beken et al. [46], such crystallographic rotation between these com- ture of ferritic grains was developed according to the ON and SG
ponents enhances the interface mobility of the {554}〈225〉α grains and, theories, respectively.
hence, favors their growth during annealing. In this context, a specific (iv) Increasing the Mo content increased the resistance to strain-in-
region of Fig. 2(d) was magnified in Fig. 13 to display the ∑19a duced martensitic transformation, as well as suppressed the grain
boundary typical of interfaces between the {112}〈110〉α and {554} growth of austenite phase during annealing at 900 °C. In addition,
〈225〉α orientations. As in the austenite phase, there was a more differences in the texture evolution, especially ferritic, in both
marked growth of the ferritic grains in the solution-annealed 2304 steels were associated with the amount of solute Mo and the

418
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

Orientaon Total CSL Total


{hkl}<uvw> Min Max Fracon Boundary Fracon
{112}<110> 0° 10° 0.328 ∑19a 0.124
{554}<225> 0° 10° 0.125

Fig. 13. – Higher magnification image of Fig. 2(d) showing the distribution of ∑19a boundary (26.5° 〈110〉) between ferritic grains with the {112}〈110〉α and {554}
〈225〉α orientations of solution-annealed 2304 samples.

temperature dependence of its retarding effect on recrystallization [9] Q. Ran, Y. Xu, J. Li, J. Wan, X. Xiao, H. Yu, et al., Effect of heat treatment on
and grain growth. transformation-induced plasticity of economical Cr19 duplex stainless steel,
Mater. Des. 56 (2014) 959–965.
(v) Σ3 boundaries were associated with the formation of annealing [10] W. Zhang, J. Hu, Effect of annealing temperature on transformation induced
twins in the austenite phase after the heat treatments. For ferrite plasticity effect of a lean duplex stainless steel, Mater. Charact. 79 (2013) 37–42.
phase, the development of ∑19a boundaries in 2304 samples was [11] J.-O. Nilsson, Super duplex stainless steels, Mater. Sci. Technol. 8 (1992) 685–700.
[12] H. Sieurin, R. Sandström, E.M. Westin, Fracture toughness of the lean duplex
associated with the favored growth of {554}〈225〉α grains during stainless steel LDX 2101, Metall. Mater. Trans. A 37 (2006) 2975–2981.
the recrystallization. Grain coarsening resulted in a larger fraction [13] S. Baldo, I. Mészáros, Effect of cold rolling on microstructure and magnetic
of the special boundaries in ferrite phase but inhibited the CSL properties in a metastable lean duplex stainless steel, J. Mater. Sci. 45 (2010)
5339–5346.
formation in the austenite phase. The higher Mo content of 2205 [14] Y.L. Fang, Z.Y. Liu, W.Y. Xue, H.M. Song, L.Z. Jiang, Precipitation of secondary
samples compared to the 2304 samples may have been the reason phases in lean duplex stainless steel 2101 during isothermal ageing, ISIJ Int. 50
for their smaller fraction of CSL boundaries in the ferrite phase. (2010) 286–293.
[15] M. Pohl, O. Storz, T. Glogowski, Effect of intermetallic precipitations on the
properties of duplex stainless steel, Mater. Charact. 58 (2007) 65–71.
Acknowledgments [16] B. Deng, Z. Wang, Y. Jiang, T. Sun, J. Xu, J. Li, Effect of thermal cycles on the
corrosion and mechanical properties of UNS S31803 duplex stainless steel, Corros.
The authors are grateful to CAPES-PROEX (0878/2016), FAPEMIG Sci. 51 (2009) 2969–2975.
[17] K.T. Kim, K.W. Cho, J.S. Kim, S.K. Ahn, Y.D. Lee, A new lean duplex stainless steel
(APQ-01905-16), and CNPq (404303/2016-1) for research fellowships and its properties, 8th Duplex Stainless Steels Conference, 2010 (Beaune).
made available to students and for their financial support. Thanks, are [18] J.A. Jiménez, G. Frommeyer, M. Carsí, O.A. Ruano, Superplastic properties of a δ/
expressed to the UFMG Microscopy Center for providing excellent sci- γ stainless steel, Mater. Sci. Eng. A 307 (2001) 134–142.
[19] C.L. Beech, Effect of Temperature and Strain Rate on the Mechanical Properties
entific support. We also thank the company Aperam South America SA and Deformation Behavior of a Duplex Stainless Steel, M.S. thesis Colorado School
for the supply of samples.Data availability of Mines, Golden, CO, 1989.
[20] J.H. Potgieter, P.A. Olubambi, L. Cornish, C.N. Machio, E.S. Sherif, Influence of
nickel additions on the corrosion behaviour of low nitrogen 22% Cr series duplex
The raw/processed data required to reproduce these findings cannot stainless steels, Corros. Sci. 50 (2008) 2572–2579.
be shared at this time as the data also forms part of an ongoing study. [21] P. Schweitzer, Metallic Materials: Physical, Mechanical, and Corrosion Properties,
Marcel Dekker, Inc., New York, 2003.
[22] P. Roberge, Handbook of Corrosion Engineering, McGraw-Hill, New York, 1999.
References [23] R. Badji, B. Bacroix, M. Bouabdallah, Texture, microstructure and anisotropic
properties in annealed 2205 duplex stainless steel welds, Mater. Charact. 62
[1] G. Krauss, Steels: Processing, Structure, and Performance, ASM Int, Ohio, 2005. (2011) 833–843.
[2] J.R. Davis (Ed.), Stainless steel, ASM Int, Ohio, 1994. [24] J. Keichel, J. Foct, G. Gottstein, Deformation and annealing behavior of nitrogen
[3] I. Alvarez-Armas, S. Degallaix-Moreuil, Duplex Stainless Steels, Wiley, Hoboken, alloyed duplex stainless steels. Part I: rolling, ISIJ Int. 43 (2003) 1781–1787.
2009. [25] N. Akdut, J. Foct, Phase boundaries and deformation in high nitrogen duplex
[4] M. Breda, K. Brunelli, F. Grazzi, A. Scherillo, I. Calliari, Effects of cold rolling and stainless steels I.-rolling texture development, Scr. Metall. Mater. 32 (1995)
strain-induced martensite formation in a SAF 2205 duplex stainless steel, Metall. 103–108.
Mater. Trans. A 46 (2015) 577–586. [26] J. Ryś, W. Ratuszek, M. Witkowska, Rolling texture development in duplex type
[5] J. Michalska, M. Sozańska, Qualitative and quantitative analysis of σ and χ phases steel with strong initial texture, Arch. Metall. Mater. 50 (2005) 857–870.
in 2205 duplex stainless steel, Mater. Charact. 56 (2006) 355–362. [27] N. Akdut, J. Foct, Microstructure and deformation behavior of high nitrogen du-
[6] J.Y. Choi, J.H. Ji, S.W. Hwang, K.-T. Park, Strain induced martensitic transfor- plex stainless steels, ISIJ Int. 36 (1996) 883–892.
mation of Fe–20Cr–5Mn–0.2Ni duplex stainless steel during cold rolling: effects of [28] O. Engler, V. Randle, Introduction to Texture Analysis: Macrotexture, Microtexture
nitrogen addition, Mater. Sci. Eng. A 528 (2011) 6012–6019. and Orientation Mapping, CRC Press, Boca Raton, 2000.
[7] C. Herrera, D. Ponge, D. Raabe, Design of a novel Mn-based 1 GPa duplex stainless [29] E. Bain, N. Dunkirk, The nature of martensite, Trans.TMS-AIME, 70 1924, pp.
TRIP steel with 60% ductility by a reduction of austenite stability, Acta Mater. 59 25–47.
(2011) 4653–4664. [30] G. Kurdjumov, G. Sachs, Über den mechanismus der Stahlhärtung, Z. Phys. 64
[8] J.Y. Kang, H. Kim, K.I. Kim, C.H. Lee, H.N. Han, K.H. Oh, et al., Effect of austenitic (1930) 325–343.
texture on tensile behavior of lean duplex stainless steel with transformation in- [31] Z. Nishiyama, X-ray Investigation of the Mechanism of the Transformation From
duced plasticity (TRIP), Mater. Sci. Eng. A 681 (2017) 114–120. Face Centered Cubic Lattice to Body Centered Cubic, 23 Scientific Repors Tohoku

419
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

Imperial University, 1934, pp. 637–664. [64] L. Ryde, Application of EBSD to analysis of microstructures in commercial steels,
[32] G. Wassermann, Über den mechanismus der α-γ umwandlung des eisens, in Mater. Sci. Technol. 22 (2006) 1297–1306.
Mitteilungen aus dem Kaiser-Wilhelm-Institut für Eisenforschung, Verlag [65] R. Petrov, L. Kestens, A. Wasilkowska, Y. Houbaert, Microstructure and texture of
Stahleisen, Düsseldorf, 1935. a lightly deformed TRIP-assisted steel characterized by means of the EBSD tech-
[33] A.B. Greninger, A.R. Troiano, The mechanism of martensitic transformation, J. nique, Mater. Sci. Eng. A 447 (2007) 285–297.
Met. Trans. 185 (1949) 590–598. [66] I. Duprez, B. De Cooman, N. Akdut, Deformation behaviour of duplex stainless
[34] W. Pitsch, The martensite transformation in thin foils of iron-nitrogen alloys, steel during industrial hot rolling, Steel Res. 73 (2002) 531–538.
Philos. Mag. 4 (1959) 577–584. [67] L. Duprez, B. De Cooman, N. Akdut, Flow stress and ductility of duplex stainless
[35] T.J. Headley, J.A. Brooks, A new bcc-fcc orientation relationship observed be- steel during high-temperature torsion deformation, Metall. Mater. Trans. A 33
tween ferrite and austenite in solidification structures of steels, Metall. Mater. (2002) 1931–1938.
Trans. A Phys. Metall. Mater. Sci. 33 (2002) 5–15. [68] K. Tsuzaki, H. Matsuyama, M. Nagao, T. Maki, High-strain rate superplasticity and
[36] M.Z. Ahmed, P.P. Bhattacharjee, Microstructure, texture, and tensile properties of role of dynamic recrystallization in a superplastic duplex stainless steel, Mater.
a severely warm-rolled and annealed duplex stainless steel, Steel Res. Int. 87 Trans. 31 (1990) 983–994.
(2016) 472–483. [69] R. Wang, Deformation and restoration behaviour of ferrite-austenite two-phase
[37] F.J. Humphreys, M. Hatherly, Recrystallisation and Related Annealing structures, Mater. Sci. Eng. A 165 (1993) 19–27.
Phenomena, Pergamom, Oxford, 1995. [70] G. Reis, A. Jorge Jr., O. Balancin, Influence of the microstructure of duplex
[38] M. Witkowska, J. Rys, W. Ratuszec, A. Zielinska-Lipiec, Annealing textures and stainless steels on their failure characteristics during hot deformation, Mater. Res.
precipitation behavior in ferritic-austenitic duplex type steels, Arch. Metall. Mater. 3 (2000) 31–35.
50 (2005) 471–478. [71] Y. Jin, M. Bernacki, G. Rohrer, A. Rollett, B. Lin, N. Bozzolo, Formation of an-
[39] J.-I. Hamada, E. Ishimaru, H. Inoue, Texture and planar anisotropy in lean duplex nealing twins during recrystallization and grain growth, 5th International
stainless steel sheet, ISIJ Int. 56 (2016) 1831–1839. Conference on Recrystallization and Grain Growth, 2013 Sydney.
[40] M. Zaid, P. Bhattacharjee, Microtexture of constituent phases in a heavily warm- [72] A. Asli, A. Zarei-Hanzaki, The dynamic recrystallization behavior of a Fe-Cr-Ni
rolled and annealed duplex stainless steel, IOP Conf. Ser.: Mater. Sci. Eng, 82 super-austenitic stainless steel, J. Mater. Sci. Technol. 25 (2009) 603–606.
2015. [73] H. Sun, Y. Sun, R. Zhang, M. Wang, R. Tang, Z. Zhou, Hot deformation behavior
[41] M. Ahmed, P. Bhattacharjee, Evolution of microstructure and texture during iso- and microstructural evolution of a modified 310 austenitic steel, Mater. Des. 64
thermal annealing of a heavily warm-rolled duplex steel, ISIJ Int. 54 (2014) (2014) 374–380.
2844–2853. [74] I. Altenberger, B. Scholtes, U. Martin, H. Oettel, Cyclic deformation and near
[42] M. Blicharski, J. Jura, T.B.R. Penelle, J. Bornarski, M. Kowalski, Development of surface microstructures of shot peened or deep rolled austenitic stainless steel AISI
the orientation relationship between ferritic and austenitic phases during long 304, Mater. Sci. Eng. A 264 (1999) 1–16.
time annealing of duplex stainless steel, Arch. Metall. Mater. 50 (2005) 495–502. [75] W. Hübner, Phase transformations in austenitic stainless steels during low tem-
[43] J.-I. Hamada, H. Inoue, Texture and planar anisotropy of r-value in duplex perature tribological stressing, Tribol. Int. 34 (2001) 231–236.
stainless steel sheet, Mater. Trans. 51 (2010) 644–651. [76] S.S. Tavares, M.R. Silva, J.M. Pardal, H.F. Abreu, A.M. Gomes, Microstructural
[44] J. Keichel, J. Foct, G. Gottstein, Deformation and annealing behavior of nitrogen changes produced by plastic deformation in the UNS S31803 duplex stainless steel,
alloyed duplex stainless steels. Part II: annealing, ISIJ Int. 43 (2003) 1788–1794. J. Mater. Process. Technol. 180 (2006) 318–322.
[45] K.M. Lee, M.Y. Huh, O. Engler, Quantitative analysis of micro-textures during [77] M. Naghizadeh, H. Mirzadeh, Microstructural evolutions during reversion an-
recrystallization in an interstitial-free steel, Steel Res. Int. 83 (2012) 919–926. nealing of cold-rolled AISI 316 austenitic stainless steel, Metall. Mater. Trans. A 49
[46] K. Verbeken, L. Kestens, J.J. Jonas, Microtextural study of orientation change (2018) 2248–2256.
during nucleation and growth in a cold rolled ULC steel, Scr. Mater. 48 (2003) [78] S.S. Tavares, J.M. Pardal, M.R. Silva, C.A. de Oliveira, Martensitic transformation
1457–1462. induced by cold deformation of lean duplex stainless steel UNS S32304, Mater.
[47] I. Samajdar, R.D. Doherty, Role of S[{123}〈634〉] orientations in the preferred Res. 17 (2014) 381–385.
nucleation of cube grains in recrystallization of FCC metals, Scr. Metall. Mater. 32 [79] A. Kurc-Lisiecka, W. Ozgowicz, W. Ratuszek, K. Chruściel, Texture and structure
(1995) 845–850. evolution during cold rolling of austenitic stainless steel, J. Achiev. Mater. Manuf.
[48] A. Elsner, R. Kaspar, D. Ponge, D. Raabe, S. Van Der Zwaag, Recrystallisation Eng. 52 (2012) 22–30.
texture of cold rolled and annealed IF steel produced from ferritic rolled hot strip, [80] Y. Guo, J. Hu, J. Li, L. Jiang, T. Liu, Y. Wu, Effect of annealing temperature on the
Mater. Sci. Forum 467–470 (2004) 257–262. mechanical and corrosion behavior of a newly developed novel lean duplex
[49] F. Gao, F. Yu, R.D. Misra, X. Zhang, S. Zhang, Z. Liu, Microstructure, texture, and stainless steel, Dent. Mater. 7 (2014) 6604–6619.
deep drawability under two different cold-rolling processes in ferritic stainless [81] H. Tan, Y. Jiang, B. Deng, T. Sun, J. Xu, J. Li, Effect of annealing temperature on
steel, J. Mater. Eng. Perform. 24 (2015) 3862–3880. the pitting corrosion resistance of super duplex stainless steel UNS S32750, Mater.
[50] D.N. Lee, H.N. Han, Recrystallization textures of metals and alloys, in: P. Wilson Charact. 60 (2009) 1049–1054.
(Ed.), Recent Dev. Study Recryst, InTech, 2013, pp. 3–59. [82] G. Fargas, M. Anglada, A. Mateo, Effect of the annealing temperature on the
[51] D. Raabe, K. Lücke, Textures of ferritic stainless steels, Mater. Sci. Technol. 19 mechanical properties, formability and corrosion resistance of hot-rolled duplex
(1993) 302–312. stainless steel, J. Mater. Process. Technol. 209 (2009) 1770–1782.
[52] P.O. Malta, C.M. Gonçalves, D.S. Alves, A.O. Ferreira, I.D. Moutinho, D.B. Santos, [83] N. Saenarjhan, J.-H. Kang, S.C. Lee, S.-J. Kim, Influence of annealing temperature
The influence of hot band annealing on recrystallization kinetics and texture on deformation behavior of 329LA lean duplex stainless steel, Mater. Sci. Eng. A
evolution in a cold-rolled Nb-stabilized ferritic stainless steel during isothermal 679 (2017) 531–537.
annealing, J. Mater. Res. 31 (2016) 2838–2849. [84] L. Zhang, W. Zhang, Y. Jiang, B. Deng, D. Sun, J. Li, Influence of annealing
[53] R.D. Doherty, D.A. Hughes, F.J. Humphreys, J.J. Jonas, D. Juul Jensen, treatment on the corrosion resistance of lean duplex stainless steel 2101,
M.E. Kassner, et al., Current issues in recrystallization: a review, Mater. Sci. Eng. A Electrochim. Acta 54 (2009) 5387–5392.
238 (1997) 219–274. [85] Z. Zhang, D. Han, Y. Jiang, C. Shi, J. Li, Microstructural evolution and pitting
[54] J. Hjelen, R. Ørsund, E. Nes, On the origin of recrystallization textures in alumi- resistance of annealed lean duplex stainless steel UNS S32304, Nucl. Eng. Des. 243
nium, Acta Metall. Mater. 39 (1991) 1377–1404. (2012) 56–62.
[55] J.W. Slakhorst, The development of recrystallization textures in F.C.C. metals with [86] H. Hu, Texture of metals, Text 1 (1974) 233–258.
a low stacking fault energy, Acta Metall. 23 (1975) 301–308. [87] B. Hutchinson, L. Ryde, E. Lindh, K. Tagashira, Texture in hot rolled austenite and
[56] P.O. Malta, D.S. Alves, A.O. Ferreira, I.D. Moutinho, C.A. Dias, D.B. Santos, Static resulting transformation products, Mater. Sc. Eng. A 257 (1998) 9–17.
recrystallization kinetics and crystallographic texture of Nb-stabilized ferritic [88] T. Waterschoot, L. Kestens, B. Cooman, Hot rolling texture development in
stainless steel based on orientation imaging microscopy, Metall. Mater. Trans. A. CMnCrSi dual-phase steels, Metall. Mater. Trans. A 33 (2002) 1091–1102.
48 (2017) 1288–1309. [89] R. Doherty, Recrystallization and texture, Prog. Mater. Sci. 42 (1997) 39–58.
[57] G. Wranglen, Pitting and sulphide inclusions in steel, Corros. Sci. 14 (1974) [90] S. Barella, C. Mapelli, Mechanical properties and textures of duplex stainless steel
331–349. rolled worked by asymmetric rolling, Int. J. Mater. Form. 4 (2011) 379–388.
[58] H. Krawiec, V. Vignal, O. Heintz, R. Oltra, Influence of the dissolution of MnS [91] J. Hirsch, K. Lücke, Mechanism of deformation and development of rolling tex-
inclusions under free corrosion and potentiostatic conditions on the composition tures in polycrystalline f.c.c. metals–I. Description of rolling texture development
of passive films and the electrochemical behaviour of stainless steels, Electrochim. in homogeneous CuZn alloys, Acta Metall. 36 (1988) 2863–2882.
Acta 51 (2006) 3235–3243. [92] N. Jia, R. Lin Peng, Y.D. Wang, S. Johansson, P.K. Liaw, Micromechanical behavior
[59] P. Paulraj, R. Garg, Effect of intermetallic phases on corrosion behavior and me- and texture evolution of duplex stainless steel studied by neutron diffraction and
chanical properties of duplex stainless steel and super-duplex stainless steel, Adv. self-consistent modeling, Acta Mater. 56 (2008) 782–793.
Sci. Technol. Res. J. 9 (2015) 87–105. [93] J. Ryś, A. Zielińska-Lipiec, Deformation of ferrite-austenite banded structure in
[60] H. Bunge, Mathematische Methoden der Texturanalyse, Akademie-Verlag, Berlin, cold-rolled duplex steel, Arch. Metall. Mater. 57 (2012) 1041–1053.
1969. [94] C. Donadille, R. Valle, P. Dervin, R. Penelle, Development of texture and micro-
[61] D. Brandon, The structure of high-angle grain boundaries, Acta Metall. 14 (1966) structure during cold-rolling and annealing of F.C.C. Alloys: example of an aus-
1479–1484. tenitic stainless steel, Acta Metall. 37 (1989) 1547–1571.
[62] K. Wang, B. Chang, J. Chen, H. Fu, Y. Lin, Y. Lei, Effect of molybdenum on the [95] D. Lindell, Texture evolution of warm-rolled and annealed 304L and 316L auste-
microstructures and properties of stainless steel coatings by laser cladding, Appl. nitic stainless steels, IOP Conf. Ser. Mater. Sci. Eng. 82 (2015).
Sci. 7 (2017) 1–15. [96] M.-Y. Huh, O. Engler, Effect of intermediate annealing on texture, formability and
[63] P. Bhattacharjee, M. Zaid, G. Sathiaraj, B. Bhadak, Evolution of microstructure and ridging of 17%Cr ferritic stainless steel sheet, Mater. Sci. Eng. A 308 (2001) 74–87.
texture during warm rolling of duplex steel, Metall. Mater. Trans. A 45 (2014) [97] M. Yantaç, W. Roberts, D. Wilson, Texture development in ferritic stainless steel
2180–2191. sheet, Text 1 (1972) 71–86.

420
P.O. Malta et al. Materials Characterization 142 (2018) 406–421

[98] L. Kestens, J.J. Jonas, Deep drawing textures in low carbon steels, Met. Mater. 5 bay in an Fe–C–Mo alloy, Mater. Sci. Eng. A 343 (2003) 151–157.
(1999) 419–427. [104] S. Mandal, P. Sivaprasad, V. Sarma, Microstructural modification in a 15Cr-15Ni-
[99] X. Fang, K. Zhang, H. Guo, W. Wang, B. Zhou, Twin-induced grain boundary en- 2.2Mo-Ti modified austenitic stainless steel through twin induced grain boundary
gineering in 304 stainless steel, Mater. Sci. Eng. A 487 (2008) 7–13. engineering, in: A.D. Rollett (Ed.), Application of Texture Analysis, John Wiley &
[100] Y. Jin, B. Lin, M. Bernacki, G. Rohrer, A. Rollett, N. Bozzolo, Annealing twin de- Sons, Inc., Hoboken, NJ, 2008.
velopment during recrystallization and grain growth in pure nickel, Mater. Sci. [105] E. Humphreys, H. Fletcher, J. Hutchins, A. Garratt-Reed, W. Reynolds Jr.,
Eng. A 597 (2014) 295–303. H. Aaronson, et al., Molybdenum accumulation at ferrite:austenite interfaces
[101] S. Dépinoy, B. Marini, C. Toffolon-Masclet, F. Rock, A.-F. Gourgues-Lorenzon, during isothermal transformation of an Fe-0.24 Pct C-0.93 Pct Mo alloy, Metall.
Austenite grain growth in a 2.25Cr-1Mo vanadium-free steel accounting for Zener Mater. Trans. A 35 (2004) 1223–1235.
pinning and solute drag: experimental study and modeling, Metall. Mater. Trans. A [106] H. Yan, H. Bi, X. Li, Z. Xu, Microstructure and texture of Nb+Ti stabilized ferritic
48 (2017) 2289–2300. stainless steel, Mater. Charact. 59 (2008) 1741–1746.
[102] M. Tikhonova, Y. Kuzminova, X. Fang, W. Wang, R. Kaibyshev, A. Belyakov, Σ3 [107] Y. Park, D. Lee, G. Gottstein, The evolution of recrystallization textures in body
CSL boundary distributions in an austenitic stainless steel subjected to multi- centered cubic metals, Acta Mater. 46 (1998) 3371–3379.
directional forging followed by annealing, Philos. Mag. 94 (2014) 4181–4196. [108] T. Urabe, J. Jonas, Modeling texture change during the recrystallization of an IF
[103] M. Enomoto, N. Maruyama, K. Wu, T. Tarui, Alloying element accumulation at steel, ISIJ Int. 34 (1994) 435–442.
ferrite/austenite boundaries below the time–temperature–transformation diagram

421

Potrebbero piacerti anche