Sei sulla pagina 1di 14

Materials Science & Engineering A 689 (2017) 370–383

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Annealing behavior of a 304L stainless steel processed by large strain cold MARK
and warm rolling

Marina Odnobokovaa, Andrey Belyakova, , Nariman Enikeevb, Dmitri A. Molodovc,
Rustam Kaibysheva
a
Belgorod State University, Belgorod 308015, Russia
b
Ufa State Aviation Technical University, Ufa 450000, Russia
c
Institute of Physical Metallurgy and Metal Physics, RWTH Aachen University, Aachen 52056, Germany

A R T I C L E I N F O A BS T RAC T

Keywords: A 304L-type austenitic stainless steel was subjected to plate rolling at ambient temperature and at 573 K to total
Austenitic stainless steels strains of 3 and then annealed at temperatures of 873, 973 and 1073 K. The structural changes during annealing
Thermomechanical processing were associated with the austenite reversal (for the cold rolled samples), recrystallization and grain growth,
Electron microscopy which depended significantly on the annealing temperature. The grain growth exponent of 4 and 5 was obtained
Phase transformation
after annealing at 973 K/1073 K for the cold and warm rolled samples, whereas very sluggish grain coarsening
Recrystallization
Texture
took place at 873 K. The texture evolution during annealing in the austenite domain was characterized by
gradual weakness of cold/warm rolled textures, although the main texture components such as Brass, {110} <
112 > , and S, {123} < 634 > , remained in the annealed samples irrespective of annealing mechanisms of
microstructure evolution. The grain coarsening during annealing was accompanied by gradual softening. The
yield strength of ultrafine grained steel processed by cold/warm rolling followed by annealing could be
expressed by a Hall-Petch type relationship with σ0=160 MPa and ky=470 MPa m0.5.

1. Introduction cold worked austenitic stainless steels involves annealing treatment at


temperatures above austenite reversal. An appropriate combination of
Chromium-nickel austenitic stainless steels are one of the most cold rolling and heat treatment may result in beneficial mechanical
widely used structural materials for various engineering applications properties including high strength and sufficient ductility [10–12,14].
from kitchen stuff to elements of spaceships [1,2]. The austenitic The attractive mechanical properties in cold worked and annealed
stainless steels are frequently produced in forms of cold rolled semi- austenitic stainless steels have been attributed to ultrafine grained
products [3–6]. Among various advantages of cold rolling, one should (UFG) microstructures [15–18]. The latter ones are readily developed
be emphasized in regards to face centered cubic (fcc) austenitic in the cold worked steels during annealing as a result of inverse
stainless steels with low stacking fault energy (SFE), namely, it results (martensite to austenite) phase transformation followed by recrystalli-
in outstanding strengthening [7–14]. The yield strength can be zation of austenite. The development of UFG structure in largely
increased above 2000 MPa [8,11–13]. On the other hand, the large strained metallic materials during subsequent annealing is favored by
strain cold working leads to a drastic degradation of plasticity [10–14]. a specific recrystallization mechanism. The annealing behavior of
After rather large rolling reductions, total elongation in a tensile test submicrocrystalline or nanocrystalline metals and alloys processed by
may decrease to a few percent. This drawback limits the utilization of severe plastic deformation has been considered in terms of continuous
cold rolled austenitic stainless steels as semi-products for further post-dynamic recrystallization [19–22]. In contrast to ordinary dis-
processing, which involves various forming operations. Moreover, continuous recrystallization in cold worked materials, the continuous
austenitic stainless steels commonly experience strain-induced mar- recrystallization does not involve any nucleation stage and develops
tensitic transformation during cold working that alternates physical right on heating, when the ultrafine crystallites, which were created by
properties of the steels and may be quite detrimental for certain previous large strain deformation, start homogeneously to grow
applications. following spontaneous or transient recrystallization [19].
A common approach to recover plasticity and austenite structure in In spite of numerous studies on the UFG austenitic stainless steels


Corresponding author.
E-mail address: belyakov@bsu.edu.ru (A. Belyakov).

http://dx.doi.org/10.1016/j.msea.2017.02.073
Received 18 October 2016; Received in revised form 17 February 2017; Accepted 18 February 2017
Available online 20 February 2017
0921-5093/ © 2017 Elsevier B.V. All rights reserved.
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

Fig. 1. Initial (a) and deformation (b–d) microstructures evolved in a 304L-type stainless steel during warm (b) and cold (c, d) rolling to a total strain of 3. The black and white lines
indicate the high- and low-angle boundaries, respectively. The inverse pole figures are shown for the rolling direction (RD). (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)

processed by cold working followed by annealing [23–27], the mechan- deformation microstructures developed during rolling at room tem-
isms of UFG microstructure formation, i.e., austenite reversal and perature and 573 K on the annealed microstructures. Particular focus
recrystallization, the UFG structure stability during annealing as well was put on the grain coarsening kinetics and the texture evolution.
as the texture evolution have not been elaborated in sufficient detail.
The operating mechanism of austenite reversal, i.e., shear or diffusion, 2. Experimental
has been suggested being dependent on the difference in the Gibbs free
energy, although the both mechanisms have been frequently consid- A 304L-type austenitic steel (Fe-0.04C-18.2Cr-8.8Ni-1.65Mn-
ered as concurrently operating ones [11,28,29]. In addition work 0.43Si-0.05P-0.04S, all in wt%) was investigated. The steel was hot
hardened stainless steels with fully austenitic microstructure can be forged at 1100°С followed by air cooling to produce a uniform
obtained by warm rolling at elevated temperatures ranging approx. microstructure with the mean grain size of 24 µm. The plate rolling
from 0.3 to 0.4 Tm (homologous temperature/melting point). Then, was carried out using a 2 High Rolling Mill (Hankook M-Tech Co., Ltd)
UFG microstructures and improved mechanical properties could be with a line speed of 5 m/min at room temperature (293 K) and at
expected after subsequent annealing. However, the regularities of UFG 573 K under isothermal conditions when both the rolls and samples
microstructure/texture evolution during annealing and their effect on were preheated to designated temperature. (Note here that possible
the mechanical properties of austenitic stainless steels subjected to strain-induced heating was disregarded in multiple rolling with about
large strain warm rolling are still unclear. 10% reduction in each rolling pass for both cold and warm treatments).
The aim of the present work is to study the annealing behavior and The samples were rolled to total true strains of ε=3 (thickness
mechanical properties of an austenitic stainless steel with enhanced reduction from 30 mm to 1.5 mm). The rolled samples were annealed
corrosion resistance subjected to cold and warm rolling to large total in a conventional muffle furnace at various temperatures in the range
strain. The paper presents comparative analysis of the effect of from 873 to 1073 K followed by water quenching. The annealing

371
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

Fig. 2. Location of the main texture components of austenite and martensite phases in Euler space and the orientation distribution function (ODF) sections (φ2=45°) of a 304L-type
stainless steel after warm and cold rolling to a total strain of 3.

Table 1 technique [32]. The mechanical properties of processed samples were


Definition of texture components. evaluated by means of tensile tests using flat specimens with a gauge
length of 12 mm and cross section of 3.0×1.5 mm2. The specimens
Component {hkl} < uvw > Euler angles
were tested at ambient temperature and at a crosshead rate of 2 mm/
φ1 Φ φ2 min using an Instron 5882 testing machine. The tensile axis was
parallel to the rolling direction (RD).
Cube (C) {001} < 100 > 45 0 45
Rotated cube (H) {001} < 110 > 0/90 0 45
E {111} < 110 > 0/60 55 45 3. Results and discussion
F {111} < 112 > 30/90 55 45
I* {223} < 110 > 0 43 45
Goss (G) {110} < 001 > 90 90 45 3.1. Deformation microstructures and textures
Rotated Goss (RtG) {110} < 110 > 0 90 45
Brass (B) {110} < 112 > 55 90 45 The deformation microstructures developed in the investigated
A {110} < 111 > 35 90 45
steel after warm and cold rolling are shown in Fig. 1 along with the
Copper (Cu) {112} < 111 > 90 35 45
S {123} < 634 > 59 37 63 initial hot forged microstructure. The warm rolled microstructure
γ-fiber < 111 > parallel to ND consists of flattened wavy austenitic grains/subgrains, which are highly
ζ-fiber < 110 > parallel to ND elongated along the rolling direction. The wavy character of the rolled
α-fiber < 110 > parallel to RD microstructure results from high density of microshear bands crossing
τ-fiber < 110 > parallel to TD
η-fiber < 001 > parallel to ND
over the flattened grains/subgrians. On the other hand, cold rolling was
accompanied by strain-induced martensitic transformation. The mar-
tensite fraction comprises 0.75 in the present sample. Therefore, the
softening was studied by Vickers hardness tests with a load of 3 N. The cold rolled microstructure consists mainly of highly elongated (lamel-
microstructure and texture characterization was performed using a lar-type) strain-induced martensite grains interleaved with retained
JEM-2100 transmission electron microscope (TEM) and a Nova austenite. Similar to warm rolling, the microshear bands frequently
Nanosem 450 scanning electron microscope equipped with electron developed during cold rolling.
back-scatter diffraction (EBSD) analyzer on the sample sections normal The corresponding textures after warm and cold rolling are shown
to the transverse direction (TD). The volume fractions of the strain- in Fig. 2 as representative sections of orientation distribution functions
induced martensite were determined by averaging the data obtained (ODF). The colors in the schematic representation of texture compo-
through X-ray analysis, magnetic induction method and EBSD techni- nents (see Table 1) correspond to those in the inverse pole figures in
que [30,31]. The transverse grain size was measured on orientation Fig. 1. The warm rolled texture is mainly characterized by two texture
imaging microscopy micrographs by a linear intercept method along components, i.e., Brass ({110} < 112 > ) and S ({123} < 634 > ) compo-
the normal direction (ND). In addition to orientation imaging micro- nents, with nearly the same intensity. These components are corre-
scopy, the misorientations between grains/subgrains were analyzed by sponded with purple color, which is dominant for wavy flattened
conventional TEM Kikuchi line method with a converged beam grains/subgrains in Fig. 1b. This is typical of deformation texture for
fcc metals with low to medium SFE [33–39]. Besides the Brass and S

372
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

decreases the hardness. The hardness drop during annealing for


30 min at 973 K and 1073 K comprises 20–30% and 40–50%, respec-
tively. Following a rapid softening during 30 min annealing, the
hardness steadily decreases with an increase in the annealing duration.
Typical annealing microstructures evolved in the warm rolled
samples are shown in Fig. 4. The pancake morphology of warm rolled
microstructure does not change significantly during annealing at
873 K, although the recrystallized grains develop like UFG chains at
the boundaries of highly elongated along the rolling direction grains
and in the microshear bands after sufficient annealing times. It is
clearly seen in Fig. 4 that recrystallization readily occurs in the warm
rolled samples during annealing at 973–1073 K. Annealing at 973 K
for 30 min corresponds to early recrystallization, when the chains of
new grains just appeared. Then, the recrystallization rapidly propagates
through the sample, leading to complete development of UFG micro-
structure. An increase in the annealing temperature to 1073 K accel-
erates significantly the recrystallization kinetics. After 30 min anneal-
ing, only normal grain growth operates during annealing at 1073 K.
The strain-induced martensite does not completely transform to
austenite at 873 K. Therefore, two-phase microstructure of approx.
65% of austenite and 35% of ferrite evolves during annealing at 873 K
in the cold rolled samples (Fig. 5). In contrast to the warm rolled
samples, which are characterized by large portions of work hardened
remnants after annealing at 873 K, the partial austenite reversal and
recrystallization result in the development of UFG microstructure in
the cold rolled sample right after 30 min annealing. The presence of
ferrite grains stabilizes the duplex (austenite-ferrite) microstructure
against grain coarsening during annealing at 873 K. On the other hand,
the strain-induced martensite in the cold rolled samples completely
transforms to austenite upon heating up to 973–1073 K. The long time
annealing of the cold rolled samples at these temperatures, therefore, is
accompanied by a gradual grain growth similar to that for the warm
rolled samples. The structural changes in the cold rolled samples
during annealing can be summarized as a quick austenite reversal with
concurrent recrystallization followed by normal grain growth.
The fine structures developed upon annealing of the cold rolled
samples are represented in Fig. 6. These micrographs illustrate some
details of the UFG microstructures developed in the cold rolled samples
after annealing. The austenite reversal takes place homogeneously
resulting in the austenite and untransformed ferrite grains at 873 K
Fig. 3. Effect of annealing temperature/time on the hardness (Hv) of 304L-stainless
being homogeneously distributed throughout the microstructure
steel processed by warm and cold rolling.
(Fig. 6a). The network of phase/grain/subgrain boundaries makes
uniform microstructure consisting of ultrafine austenite and ferrite
components, the warm deformation texture includes a relatively strong
grains/subgrains, which are fairly similar in appearance. These ultra-
Goss texture component, {110} < 100 > , which is associated with the
fine austenite and ferrite crystallites contain dislocations in their
ultrafine grains within the microshear bands (red in Fig. 1b). The cold
interiors and, therefore, are characterized by high internal stresses as
rolled austenite is characterized by almost the same texture compo-
suggested by elastic bending contours on the TEM image (enlarged
nents with a predominance of the Brass component that may be
portion in Fig. 6a). It is clearly seen in Fig. 6b that the austenite
attributed to a decrease in the SFE with a decrease in the rolling
reversal and recrystallization at 973 K result in almost equiaxed fine
temperature. The strain-induced martensite exhibits a strong I* texture
grains, which are entirely bounded by high-angle boundaries. It should
component, {223} < 110 > (corresponds to green color in Fig. 1c),
also be noted that some annealed grains (Fig. 6d) are still characterized
along with remarkable γ fiber, < 111 > ∥ND. Similar textures have been
by a high dislocation density. This feature differentiates the present
frequently observed in body centered cubic metals after large strain
microstructures from conventional annealing ones developed by pri-
cold rolling and commonly associated with the main operative slip
mary recrystallization [42].
systems of the type {110} < 111 > [40,41].
Fig. 7 represents quantitative changes in the grain sizes during
annealing of the cold and warm rolled samples. The filled symbols in
3.2. Annealing softening and microstructures Fig. 7 correspond to the grain sizes, which were evaluated as a distance
between ordinary grain boundaries with misorientations of θ≥15°, i.e.,
Remarkable softening occurs during 30 min annealing, and the twin-related boundaries were omitted from measurements, and the
softening level depends substantially on the annealing temperature. open symbols represent the grain size as the mean spacing of all high-
Moreover, it is somewhat affected by the temperature of previous angle boundaries including twin-related ones. The grain coarsening
rolling, i.e., the cold rolled samples exhibit higher hardness values than during annealing at 873 K is characterized by a quite slow kinetics
the warm rolled ones (Fig. 3). Annealing at 873 K does not lead to during annealing at 873 K irrespective of single-phase (austenite) in
significant softening; the hardness decreases by about 5–10% after the warm rolled samples or two-phase (austenite-ferrite) microstruc-
30 min annealing followed by sluggish softening upon further anneal- ture in the cold rolled samples. The fraction of twin-related boundaries
ing. In contrast, an increase in the annealing temperature significantly in these samples after annealing is negligibly small. Thus, the same

373
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

Fig. 4. Typical microstructures in 304L-type stainless steel subjected to warm rolling to a total strain of 3 and then annealed at temperatures of 873 K, 973 K and 1073 K. The black,
white and red lines indicate the high-angle, low-angle and Σ3 CSL boundaries, respectively. The inverse pole figures are shown for the rolling direction (RD). (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)

grain sizes were measured irrespective of the type of the counted (i.e., the deviation of open from filled symbols in Fig. 7) increases as the
boundaries. In contrast, the austenite reversal and recrystallization at grains coarsen. This suggests that annealing twins appear during grain
an early stage of annealing at 973 K and 1073 K result in rapid increase boundary migration as a result of grain growth accidents [46,47].
in the transverse grain size; the higher the temperature, the larger the The annealing behavior of the present cold or warm rolled steel is
grain size. Then, the grains gradually grow with an increase in the quite similar to continuous recrystallization of severely strained
annealing time and demonstrate almost the same coarsening behavior metallic materials. Namely, the hardness of cold/warm worked samples
in the range of 973–1073 K. The grain growth at 973 K and 1073 K quickly drops to certain level depending on the annealing temperature
follows a power law function of annealing time with a grain growth and then smoothly decreases with an increase in the annealing time
exponent n of approx. n=4 and 5 for cold and warm rolled samples, (Fig. 3) that is essential feature of continuous recrystallization [20].
respectively. It is to mention here that n values of 2–10 have been Correspondingly, the grain size rapidly increases at an early annealing
frequently reported as grain growth exponents in various experimental followed by gradual coarsening upon further annealing (Fig. 7).
studies [42–45]. Finally, it should be noted that the number of Another specific feature of continuous recrystallization is a presence
annealing twins increases with an increase of the grain size during of dislocation substructures in the growing grains [19,22]. In fact,
annealing. The difference between the boundary spacing as measured Fig. 6 unambiguously suggests that the present annealed steel micro-
including twin-related boundaries or without twin-related boundaries structures comprise grains with high dislocation density in the course

374
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

Fig. 5. Typical microstructures in a 304L-type stainless steel subjected to cold rolling to a total strain of 3 and then annealed at temperatures of 873 K, 973 K and 1073 K. The black,
white and red lines indicate the high-angle, low-angle and Σ3 CSL boundaries, respectively. The inverse pole figures are shown for the rolling direction (RD). (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)

375
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

Fig. 6. Fine structures of a 304L-type stainless steel subjected to cold rolling and annealing for 30 min at 873 K (a) and 973 K (b). The numbers at the enlarged portions indicate the
boundary misorientations in degrees.

of recrystallization. It is interesting to note that the present cold rolled of continuous recrystallization including transient recrystallization
steel exhibit annealing behavior that is much similar to continuous followed by grain growth takes place in the present cold/warm rolled
recrystallization irrespective of partial austenite reversal. In other steels during annealing.
words, the partial phase transformation does not alter the necessary
structural conditions for continuous recrystallization in strain-induced
ultrafine grained materials. This phenomenon can be interpreted as 3.3. Annealing textures
follows (Fig. 8). The obtained cold rolled microstructure consists of
ultrafine grains of austenite and martensite, which are uniformly Annealing textures developed in the warm rolled samples are
distributed throughout. The austenite reversal is not accompanied by shown in Fig. 9 as ODF sections at φ2=45°, and relative fractions of
an increase in the fraction of twin-related boundaries (Fig. 7). This the main texture components are quantitatively represented in Fig. 10.
suggests that the austenite reversal occurs by a shear mechanism Generally, the deformation textures tend to weaken and randomize
without remarkable diffusion assisted grain boundary migration [48], during annealing; and increasing the annealing temperature hastens
i.e., without noticeable change in the grain size. Therefore, upon a the changes (cf. Figs. 2 and 9). It is worth noting that the most rapid
heating, the strain-induced martensite rapidly transforms to austenite weakening is recorded for the strongest deformation texture compo-
grains, the size of which inherits the size of preceding martensite nents, i.e., Brass and S components; whereas relatively small fractions
crystallites, which in turn is almost the same as that of the austenite of the other texture components do not significantly change during
grains in the deformation microstructure. Thus, the uniform ultrafine annealing (Fig. 10). Therefore, the texture evolution during annealing
grained austenite microstructure, which is quite similar to those is characterized by an equalization of the fractions of various texture
frequently observed in other metals and alloys subjected to severe components. The decrease in the fractions of rather strong Brass and S
plastic deformation [49], develops in the cold rolled steel upon heating. components, which are inherent in the warm rolled samples, should
The uniform ultrafine-grained microstructure was indentified to be a lead to an apparent increase in the fraction of the other texture
prerequisite for continuous recrystallization during annealing [19,22]. components. Indeed, annealing of the warm rolled samples is accom-
Therefore, since the necessary structural conditions are attained, a kind panied by a gradual increase in the fractions of Goss and Copper, {112}
< 111 > , components. The latter becomes comparable with Brass

376
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

other hand, the strong Brass component in the cold rolled state
decreases during annealing at 873 K, leading to fractional increase in
the Goss, Copper and S components (Figs. 11 and 12). An increase in
the annealing temperature to 973 K results in the full austenite
reversal. However, this ferrite-to-austenite phase transformation does
not lead to any specific changes in the final austenite texture after
annealing at 973–1073 K. It is clearly seen in Figs. 11 and 12 that the
Brass component fraction quickly weakens during annealing of the cold
rolled samples, and the annealed austenite involves the Goss, Copper
and S components in addition to the rest of Brass component. In spite
of the phase variations, therefore, the annealed steel textures developed
in the warm and cold rolled samples are quite similar to each other.
In previous studies, the occurrence of continuous recrystallization
in largely strained steels was not accompanied by a significant change
in the deformation texture. This behavior was discussed as texture
invariant recrystallization [50]. Since the new grains that resulted from
continuous recrystallization are essentially the pre-existing deforma-
tion grains, the uniform grain coarsening should not alter the
deformation texture. The present results reveal that the microstructure
evolution during annealing of both cold and warm rolled steels is
accompanied by a randomization of deformation texture, while any
new texture components scarcely develop that, therefore, points to
continuous recrystallization phenomenon (cf. Figs. 2, 9, 11).
Again, it is worth noting that the deformation texture remains
qualitatively almost the same in the cold or warm rolled samples
irrespective of the partial phase transformation. Assuming that the
austenite reversal takes place by a shear mechanism, the developing
austenite orientations should commonly follow Kurdjumov-Sachs or
Nishiyama-Wasserman orientation relationships [51,52]. Note here
that both of them predict similar orientations, with the difference
between them of about 5° [53]. The deformation martensite in the cold
rolled samples is characterized by a strong I* texture component
(Fig. 2). Fig. 13 shows the austenite orientations, which should result
from I* oriented martensite through Kurdjumov-Sachs or Nishiyama-
Wasserman transformations. It is clearly seen in Fig. 13 that the
austenite transformed from I* textured martensite may obtain the
orientations that are close to Goss or Copper or S components. The S
and Goss components were present in the cold rolled austenite. The
difference between cold rolling and annealing texture, therefore, is the
Copper component, which indeed appears and has the forth rank
Fig. 7. Grain coarsening in the warm (a) and cold (b) rolled stainless steel samples among the annealing texture fractions after S, Brass and Goss
during annealing at the indicated temperatures. components (Fig. 12).

component upon annealing at 1073 K (Fig. 9). 3.4. Tensile behavior (in relation to microstructure)
The texture evolution in the cold rolled samples during annealing at
873 K can be subdivided into ferrite and austenite constituents Representative tensile stress-strain curves of annealed steel sam-
(Fig. 11). Annealing at 873 K does not alter remarkably the ferrite ples are shown in Fig. 14. The both sets of samples, i.e., the warm and
texture, which looks like that in the cold rolling sample and does not cold rolled ones, demonstrate a common feature for the tensile
vary with an increase in the annealing time (cf. Figs. 2 and 11). On the behavior after annealing. Namely, the yield strength, ultimate tensile

Fig. 8. Schematic illustration of the uniform ultrafine grained austenite microstructure development after cold rolling and subsequent annealing.

377
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

Fig. 9. ODF sections (φ2=45°) of the steel samples subjected to warm rolling and annealing at temperatures of 873 K, 973 K and 1073 K.

strength, uniform and total elongation depend significantly on the steels [56,57]. The data in Fig. 15 suggest that there is additional
annealing temperature in the investigated temperature range of 873– strength contributor in the present ultrafine grained steels.
1073 K, whereas the annealing time varying from 1.8 ks to 7.2 ks has In contrast to discontinuous recrystallization, continuous one that
much weaker effect on the mechanical properties. Therefore, the tensile develops in UFG metals and alloys processed by large strain deforma-
flow curves of the annealed samples can be grouped in three domains tion involves homogeneous coarsening of strain-induced crystallites,
depending on the annealing temperature. Annealing at relatively low which are originally characterized by large internal distortions includ-
temperature of 873 K does not lead to any substantial changes in the ing a high density of interior dislocations and strain-induced non-
deformation microstructures. The corresponding tensile behavior is equilibrium grain/subgrain boundaries [49]. It has been suggested that
characterized by a sharp stress maximum well above 1000 MPa at very the internal stresses, which are associated with strain-induced bound-
small strains followed by a pronounced necking and failure. An aries, rapidly release during the transient recrystallization right on
increase in the annealing temperature up to 973 K results in a drastic heating, whereas the dislocation substructures remain over a long
increase in the grain size (Fig. 7). After yielding at 600–1000 MPa, the annealing time leading to incomplete softening in apparently recrys-
tensile stress-strain curves of the samples annealed at 973 K exhibit tallized microstructures [58]. The same discussion can be applied to the
distinct strain hardening, which extends up to rather large strains of structural strengthening observed in the current work. Namely, the
about 20%. Both the engineering stress and the elongation notably annealed UFG steel samples are characterized by high dislocation
depend on the annealing duration at 973 K. Annealing at the highest densities (Fig. 6), which contribute to strengthening and are respon-
studied temperature of 1073 K produces the largest grain size of about sible for a deviation of Hall-Petch type relationship from that observed
2–3 µm (Fig. 7). The tensile stress-strain curves for these samples look for coarse grained steels, similar to the work hardening effect on the
like usual ones, which have regularly appeared in austenitic stainless strength [59].
steels with recrystallized microstructures [11,54]. Finally, it should be noted that the cold rolled samples exhibit
The relationship between the yield strength (σ0.2) and the grain size higher strength than the warm rolled samples in spite of the same grain
is represented in Fig. 15. The present UFG steel samples processed by size range after annealing at 973 K. This may result from the originally
cold/warm rolling and annealing obey the following Hall-Petch type higher dislocation density in the cold rolled samples as compared to the
relationship. warm rolled ones. The dislocation densities in cold and warm rolled
samples become similar to each other after annealing at relatively high
σ0.2 = 160 + 470 D−0.5 (1) temperature of 1073 K, when significant grain coarsening occurs. It can
Almost the same relationship has been reported for a wide range of be concluded, therefore, that the high dislocation density gradually
grain sizes including nanoscale domain in austenitic stainless steels decreases during continuous recrystallization and provides an addi-
[11,55]. In contrast, the yield strength of coarse grained steel after tional strengthening of UFG steels processed by large strain cold/warm
annealing can be related to their grain sizes as follows [30]. rolling followed by annealing.

σ0.2 = 180 + 240 D−0.5 (2)


4. Conclusions
Note here that similar tendency, i.e., increasing the first term in Eq.
(2) and decreasing the grain size strengthening factor as the grain size The annealing behavior of a 304L-type austenitic stainless steel
increases above 5 µm, has been observed in other studies on austenitic subjected to large strain cold and warm rolling at 293 K and 573 K,

378
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

respectively, was studied. The main results can be summarized as


follows.

1. The austenite reversal, recrystallization and grain growth took place


during annealing of the cold rolled samples, while recrystallization
followed by grain growth occurred in the warm rolled steel, leading
to ultrafine grained structures with a grain size of about 0.5 µm and
upwards after annealing at 873 K, 973 K and 1073 K in the both cold
and warm rolled samples. The grain coarsening during annealing at
973 K and 1073 K was characterized by grain growth exponents of 4
and 5 for the cold and warm rolled steel samples.
2. The deformation textures tended to weaken and randomize during
annealing, although their main components did not alter qualita-
tively irrespective of phase transformation. The fractions of strong
Brass ({110} < 112 > ) and S ({123} < 634 > ) components in the
warm rolled austenite and Brass component in the cold rolled
austenite decreased during annealing, whereas the fractions of other
relatively weak austenite texture components like Goss ({110} < 100
> ), Copper ({112} < 111 > ) in cold/warm rolled samples and S
component in cold rolled sample apparently increased. The strong
ferrite I* ({223} < 110 > ) texture component, along with γ fiber ( <
111 > ∥ND), which was developed by cold rolling did not change
remarkably during annealing.
3. The observed annealing behavior can be interpreted in terms of
continuous recrystallization irrespective of partial austenite reversal.
The strain-induced martensite rapidly transformed to austenite
grains, the size of which was almost the same as that of the austenite
grains in the deformation microstructure, leading to the uniform
ultrafine grained austenite microstructure upon heating. Then, the
uniform ultrafine-grained austenite microstructure gradually coar-
sened during annealing that is the essence of continuous recrystalli-
zation.
4. The ultrafine grained steel processed by large strain cold/warm
rolling followed by annealing exhibited superior yield strength as
compared to that extrapolated from coarse grained domain by Hall-
Petch type relationship. The values of σ0=160 MPa and
ky=470 MPa m0.5 were obtained for the ultrafine grained samples.

Acknowledgments

The authors gratefully acknowledge the financial support from the


Ministry of Education and Science, Russia (A.B., Belgorod State
University project No. 1683); the Russian Foundation for Basic
Research (M.O. and N.E., Grant 16–38-50051-mol-nr); the Deutsche
Forschungsgemeinschaft (D.A.M., Collaborative Research Centre (SFB)
program 761: “Steel—ab initio; quantum mechanics guided design of
new Fe based materials”).

Fig. 10. The influence of annealing temperature/time on the fractions of the main
texture components in the steel samples processed by warm rolling to a total strain of 3.

379
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

Fig. 11. ODF sections (φ2=45°) of the steel samples subjected to cold rolling and annealing at temperatures of 873 K, 973 K and 1073 K.

380
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

Fig. 13. Orientations of austenite transformed from {223} < 110 > oriented ferrite
through Kurdjumov-Sachs (a) and Nishiyama-Wasserman (b) orientation relationships.
The texture component, i.e., Brass (B), S, Goss (G), Rotated Goss (RtG), Copper (Cu),
which are peculiar to austenite are also indicated.

Fig. 12. The influence of annealing temperature/time on the fractions of the main
texture components in the steel samples processed by cold rolling to a total strain of 3.

381
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

References

[1] K.H. Lo, C.H. Shek, J.K.L. Lai, Recent developments in stainless steel, Mater. Sci.
Eng. R. 65 (2009) 39–104.
[2] C.G. de Andres, C. Capdevila, D. San Martin, Structural steels, in: Encycopedia of
Iron, Steel, and Their Alloys, Taylor & Francis, UK, 2016.
[3] S. Watanabe, Technological progress and future outlook for stainless steel, Nippon
Steel Tech. Rep. 71 (1996) 1–9.
[4] S.G. Chowdhury, S. Das, P.K. De, Cold rolling behaviour and textural evolution in
AISI 316L austenitic stainless steel, Acta Mater. 53 (2005) 3951–3959.
[5] K. Omura, S. Kunioka, M. Furukawa, Product development on market trends of
stainless steel and its future prospects, Nippon Steel Tech. Rep. 99 (2010) 9–19.
[6] N. Nakada, H. Ito, Y. Matsuoka, T. Tsuchiyama, S. Takaki, Deformation-induced
martensitic transformation behavior in cold-rolled and cold-drawn type 316
stainless steels, Acta Mater. 58 (2010) 895–903.
[7] M. Hadji, R. Badji, Microstructure and mechanical properties of austenitic stainless
steels after cold rolling, J. Mater. Eng. Perform. 11 (2002) 145–151.
[8] Y. Nakao, H. Miura, Nano-grain evolution in austenitic stainless steel during multi-
directional forging, Mater. Sci. Eng. A 528 (2010) 1310–1317.
[9] H. Ueno, K. Kakihata, Y. Kaneko, S. Hashimoto, A. Vinogradov, Martensite/
austenite interfaces in ultrafine grained Fe–Ni–C alloy, J. Mater. Sci. 46 (2011)
4276–4283.
[10] A. Rezaee, A. Kermanpur, A. Najafizadeh, M. Moallemi, Production of nano/
ultrafine grained AISI 201L stainless steel through advanced thermo-mechanical
treatment, Mater. Sci. Eng. A 528 (2011) 5025–5029.
[11] I. Shakhova, V. Dudko, A. Belyakov, K. Tsuzaki, R. Kaibyshev, Effect of large strain
cold rolling and subsequent annealing on microstructure and mechanical proper-
ties of an austenitic stainless steel, Mater. Sci. Eng. A 545 (2012) 176–186.
[12] A. Belyakov, A. Kipelova, M. Odnobokova, I. Shakhova, R. Kaibyshev, Development
of ultrafine grained austenitic stainless steels by large strain deformation and
annealing, Mater. Sci. Forum 783–786 (2014) 651–656.
[13] A. Belyakov, M. Odnobokova, A. Kipelova, K. Tsuzaki, R. Kaibyshev,
Nanocrystalline structures and tensile properties of stainless steels processed by
severe plastic deformation, in: IOP Conference Series: Materials Science and
Engineering, vol. 63, 2014, (Paper No. 012156).
[14] M. Odnobokova, A. Belyakov, A. Kipelova, R. Kaibyshev, Formation of ultrafine-
grained structures in 304L and 316L stainless steels by recrystallization and reverse
phase transformation, Mater. Sci. Forum 838–839 (2016) 410–415.
[15] K. Tomimura, S. Takaki, S. Tanimoto, Y. Tokunaga, Optimal chemical composition
in Fe-Cr-Ni Alloys for ultra grain refining by reversion from deformation induced
martensite, ISIJ Int. 31 (1991) 721–727.
[16] A. Di Schino, M. Barteri, J.M. Kenny, Effects of grain size on the properties of a low
nickel austenitic stainless steel, J. Mater. Sci. 38 (2003) 4725–4733.
[17] M. Eskandari, A. Kermanpur, A. Najafizadeh, Formation of nanocrystalline
structure in 301 stainless steel produced by martensite treatment, Metall. Trans. A
40 (2009) 2241–2249.
[18] S.V. Dobatkin, O.V. Rybalchenko, N.A. Enikeev, A.A. Tokar, M.M. Abramova,
Formation of fully austenitic ultrafine-grained high strength state in metastable Cr-
Ni-Ti stainless steel by severe plastic deformation, Mater. Lett. 166 (2016)
276–279.
[19] A. Belyakov, T. Sakai, H. Miura, R. Kaibyshev, K. Tsuzaki, Continuous recrystalli-
zation in austenitic stainless steel after large strain deformation, Acta Mater. 50
Fig. 14. Engineering σ-ε curves for a 304L-type stainless steel processed by warm/cold
(2002) 1547–1557.
rolling to a total strain of 3 and then annealed under the indicated conditions.
[20] A. Belyakov, K. Tsuzaki, Y. Kimura, Y. Mishima, Annealing behavior of a ferritic
stainless steel subjected to large-strain cold working, J. Mater. Res. 22 (2007)
3042–3051.
[21] B.R. Kumar, S. Sharma, Recrystallization behavior of a heavily deformed austenitic
stainless steel during iterative type annealing, Metall. Trans. A 45A (2014)
6027–6038.
[22] T. Sakai, A. Belyakov, R. Kaibyshev, H. Miura, J.J. Jonas, Dynamic and post-
dynamic recrystallization under hot, cold and severe plastic deformation condi-
tions, Prog. Mater. Sci. 60 (2014) 130–207.
[23] D.L. Johannsen, A. Kyrolainen, P.J. Ferreira, Influence of annealing treatment on
the formation of nano/submicron grain size AISI 301 austenitic stainless steels,
Metall. Trans. A 37 (2006) 2325–2338.
[24] C.S. Yoo, Y.M. Park, Y.S. Jung, Y.K. Lee, Effect of grain size on transformation-
induced plasticity in an ultrafine-grained metastable austenitic steel, Scr. Mater. 59
(2008) 71–74.
[25] M. Karimi, A. Najafizadeh, A. Kermanpur, M. Eskandari, Effect of martensite to
austenite reversion on the formation of nano/submicron grained AISI 301 stainless
steel, Mater. Charact. 60 (2009) 1220–1223.
[26] S. Takaki, K. Tomimura, S. Ueda, Effect of pre-cold-working on diffusional
reversion of deformation induced martensite in metastable austenitic stainless
steel, ISIJ Int. 34 (1994) 522–527.
[27] F. Forouzan, A. Najafizadeh, A. Kermanpur, A. Hedayati, R. Surkialiabad,
Production of nano/submicron grained AISI 304L stainless steel through the
martensite reversion process, Mater. Sci. Eng. A 527 (2010) 7334–7339.
[28] K. Tomimura, S. Takaki, Y. Tokunaga, Reversion mechanism from deformation
induced martensite to austenite in metastable austenitic stainless steels, ISIJ Int.
31 (1991) 1431–1437.
Fig. 15. Relationship between the offset yield strength and the grain size in a 304L-type [29] R.D.K. Misra, Z. Zhang, P.K.C. Venkatasurya, M.C. Somani, L.P. Karjalainen,
stainless steel subjected to cold rolling (CR) or warm rolling (WR) followed by annealing. Martensite shear phase reversion-induced nanograined/ultrafine-grained Fe–
16Cr–10Ni alloy: the effect of interstitial alloying elements and degree of austenite
stability on phase reversion, Mater. Sci. Eng. A 527 (2010) 7779–7792.
[30] M. Odnobokova, A. Belyakov, R. Kaibyshev, Development of nanocrystalline 304L

382
M. Odnobokova et al. Materials Science & Engineering A 689 (2017) 370–383

stainless steel by large strain cold working, Metals 5 (2015) 656–668. Steel Inst. 208 (1970) 475–481.
[31] M. Odnobokova, A. Belyakov, R. Kaibyshev, Effect of severe cold or warm [46] M. Shimada, H. Kokawa, Z.J. Wang, Y.S. Sato, I. Karibe, Optimization of grain
deformation on microstructure evolution and tensile behavior of a 316L stainless boundary character distribution for intergranular corrosion resistant 304 stainless
steel, Adv. Eng. Mater. 17 (2015) 1812–1820. steel by twin-induced grain boundary engineering, Acta Mater. 50 (2002)
[32] D.B. Williams, C.D. Carter, Transmission Electron Microscopy, Plenum Press, New 2331–2341.
York, 1996. [47] S. Mahajan, Critique of mechanisms of formation of deformation, annealing and
[33] J. Hirsch, K. Lucke, Mechanism of deformation and development of rolling textures growth twins: face-centered cubic metals and alloys, Scr. Mater. 68 (2013) 95–99.
in polycrystalline F.C.C. metals – I. description of rolling texture development in [48] M. Tikhonova, Y. Kuzminova, X. Fang, W. Wang, R. Kaibyshev, A. Belyakov, Σ3
homogeneous CuZn alloys, Acta Metall. 36 (1988) 2863–2882. CSL boundary distributions in an austenitic stainless steel subjected to multi-
[34] B.R. Kumar, B. Mahato, N.R. Bandyopadhyay, D.K. Bhattacharya, Comparison of directional forging followed by annealing, Philos. Mag. 94 (2014) 4181–4196.
rolling texture in low and medium stacking fault energy austenitic stainless steels, [49] R.Z. Valiev, R.K. Islamgaliev, I.V. Alexandrov, Bulk nanostructured materials from
Mater. Sci. Eng. A 394 (2005) 296–301. severe plastic deformation, Prog. Mater. Sci. 45 (2000) 103–189.
[35] Y. Lü, D.A. Molodov, G. Gottstein, Correlation between microstructure and texture [50] K. Tsuzaki, A. Belyakov, F. Yin, Texture invariant annealing in severely deformed
development in a cold-rolled TWIP steel, ISIJ Intern. 51 (2011) 812–817. steel, Mater. Sci. Forum 558–559 (2007) 101–106.
[36] C. Haase, S.G. Chowdhury, L.A. Barrales-Mora, D.A. Molodov, G. Gottstein, On the [51] H. Kitahara, R. Ueji, M. Ueda, N. Tsuji, Y. Minamino, Crystallographic analysis of
relation of microstructure and texture evolution in an austenitic Fe-28Mn-0.28C plate martensite in Fe–28.5 at% Ni by FE-SEM/EBSD, Mater. Charact. 54 (2005)
TWIP steel during cold rolling, Metall. Mater. Trans. A 44A (2013) 911–922. 378–386.
[37] M. Nezakat, H. Akhiani, M. Hoseini, J. Szpunar, Effect of thermo-mechanical [52] H. Kitahara, R. Ueji, N. Tsuji, Y. Minamino, Crystallographic features of lath
processing on texture evolution in austenitic stainless steel 316L, Mater. Charact. martensite in low-carbon steel, Acta Mater. 54 (2006) 1279–1288.
98 (2014) 10–17. [53] Y. He, S. Godet, J.J. Jonas, Observations of the Gibeon meteorite and the inverse
[38] C. Haase, M. Kühbach, L.A. Barrales-Mora, S.L. Wong, F. Roters, D.A. Molodov, Greninger–Troiano orientation relationship, Appl. Crystall. 39 (2006) 72–81.
G. Gottstein, Recrystallization behavior of a high-manganese steel: experiments [54] B.R. Kumar, S. Sharma, B. Mahato, Formation of ultrafine grained microstructure
and simulations, Acta Mater. 100 (2015) 155–168. in the austenitic stainless steel and its impact on tensile properties, Mater. Sci. Eng.
[39] Z. Yanushkevich, A. Belyakov, C. Haase, D.A. Molodov, R. Kaibyshev, Structural/ A 528 (2011) 2209–2216.
textural changes and strengthening of an advanced high-Mn steel subjected to cold [55] C.X. Huang, G. Yang, C. Wang, Z.F. Zhang, S.D. Wu, Mechanical behaviors of
rolling, Mater. Sci. Eng. A 651 (2016) 763–773. ultrafine-grained 301 austenitic stainless steel produced by equal-channel angular
[40] R.K. Ray, J.J. Jonas, R.E. Hook, Cold rolling and annealing textures in low carbon pressing, Metall. Mater. Trans. A 42A (2011) 2061–2071.
and extra low carbon steels, Int. Mater. Rev. 39 (1994) 129–172. [56] B.P. Kashyap, K. Tangri, On the Hall-Petch relationship in type 316L stainless steel
[41] H.-R. Wenk, P. Van Houtte, Texture and anisotropy, Rep. Prog. Phys. 67 (2004) at room temperature, Scr. Metall. Mater. 24 (1990) 1777–1782.
1367–1428. [57] A. Di Schino, J.M. Kenny, Grain refinement strengthening of a micro-crystalline
[42] F.J. Humphreys, M. Hatherly, Recrystallization and Related Annealing high nitrogen austenitic stainless steel, Mater. Lett. 57 (2003) 1830–1834.
Phenomena, Pergamon Press, Oxford, 1996. [58] A. Belyakov, Y. Kimura, K. Tsuzaki, Recovery and recrystallization in ferritic
[43] P.A. Beck, M.L. Holtzworth, H. Hu, Instantaneous rates of grain growth, Phys. Rev. stainless steel after large strain deformation, Mater. Sci. Eng. A 403 (2005)
73 (1948) 526–527. 249–259.
[44] H. Hu, B.B. Rath, On the time exponent in isothermal grain growth, Metall. Trans. [59] B.P. Kashyap, K. Tangri, On the Hall-Petch relationship and substructural
1 (1970) 3181–3184. evolution in type 316L stainless steel, Acta Metall. Mater. 43 (1995) 3971–3981.
[45] N.E. Hannerz, F. De Kazinczy, Kinetics of austenite grain growth in steel, J. Iron

383

Potrebbero piacerti anche