You are on page 1of 13

Acta Materialia 50 (2002) 4177–4189

www.actamat-journals.com

Ultragrain refinement of plain low carbon steel by cold-


rolling and annealing of martensite
R. Ueji a, N. Tsuji b,∗, Y. Minamino b, Y. Koizumi b
a
Graduate student of Osaka University, 2-1, Yamadaoka, Suita, Osaka, 565-0871, Japan
b
Department of Adaptive Machine Systems, Osaka University, 2-1, Yamadaoka, Suita, Osaka, 565-0871, Japan

Received 21 March 2002; accepted 22 May 2002

Abstract

Simple cold-rolling and annealing of martensite starting structure can produce ultrafine grained structure in carbon
steel. The microstructural evolution during the process was studied in a 0.13%C steel. The ultrafine lamellar dislocation
cells (LDCs) with mean thickness of 60 nm were mainly observed in a 50% cold-rolled specimen as well as the
irregularly bent lamellas (IBLs) and the kinked laths (KLs). The LDCs and the IBLs had large local misorientations.
The specimens annealed at temperatures from 723 to 773 K showed the multiphased ultrafine structure composed of
equiaxed ultrafine ferrite grains with the mean grain size of 180 nm, nano-carbides distributed uniformly and small
blocks of tempered martensite. The formation of the ultrafine grained structure was discussed from the viewpoint of
characteristics of the martensite starting structure. It was concluded that the fine grained structure of martensite play
an important role for ultrafine grain subdivision during plastic deformation.  2002 Acta Materialia Inc. Published by
Elsevier Science Ltd. All rights reserved.

Keywords: Steels; Annealing; Cold working; Microstructure; Martensite

1. Introduction milling for metallic powder [8,9], multiple com-


pression [10] and accumulative roll-bonding
In recent years, ultrafine grained (UFG) struc- (ARB) [11–17], have been established, and they
tural materials with mean grain size of less than 1 have made great progress in ultra-grain refinement.
µm have been studied aggressively, because they However, the SPD processes may be difficult to
are expected to provide superior mechanical apply to practical manufacturing, because of their
properties. In order to produce the UFG materials, complicated procedures and/or inapplicability to
severe plastic deformation (SPD) processes [1], bulky workpieces. On the other hand, the present
such as torsion straining under high pressure [2,3], authors have developed a new and simple way to
equal channel angular pressing [4–7], mechanical produce UFG low-carbon steels without SPD
[18,19]. The process is characterized by conven-
tional cold-rolling and annealing of a martensite

Corresponding author. Fax: +81-6-6879-7434. starting microstructure. Plain low-carbon steel JIS-
E-mail address: tsuji@ams.eng.osaka-u.ac.jp (N. Tsuji). SS400, for example, 50% cold-rolled and annealed

1359-6454/02/$22.00  2002 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.
PII: S 1 3 5 9 - 6 4 5 4 ( 0 2 ) 0 0 2 6 0 - 4
4178 R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189

at 823 K from a martensite starting structure a 10% HClO4+90% CH3COOH solution, and
revealed a multiphased UFG structure and exhib- observed with a Hitachi H-800 transmission elec-
ited superior strength and ductility (870 MPa of tron microscope operated at 200 kV. The precise
tensile strength and 20% of total elongation) crystallographic orientations of the UFG grains
[18,19]. It was surprising that only a 50% rolling were measured by Kikuchi-line analysis or detailed
reduction (equivalent strain of 0.8) could produce diffraction spot pattern analysis in a Philips
the UFG structure, since very high strain over 4 CM200 transmission electron microscope operated
has been generally applied to the materials in SPD at 200 kV. The program developed by Zaefferer
processes to obtain UFGs. The key of the process [20] was used for the semi-automatic on-line
must be to start from a martensitic microstructure. analysis of the Kikuchi-lines and the diffraction
The aim of this paper is to clarify the mechanism spot patterns. Thin specimens of the as-quenched
of the microstructural evolution during this new martensite prepared by the same method as the
type of thermomechanical processing. TEM specimens were provided for the investi-
gation of orientation mapping by the electron back
scattering diffraction (EBSD) method in a Philips
2. Experimental XL30 scanning electron microscope with LaB6
filament operated at 25 kV. The TSL-OIM
Hot-rolled sheets of a commercial plain low-car- software (ver. 3.03) [21,22] was used for automatic
bon steel JIS-SS400 (400 MPa class; 0.13wt%C- orientation mapping. The mapping was carried out
0.01%Si-0.37%Mn-0.020%P-0.0043%N) were in a 200 µm×200 µm area by 0.5 µm of step size
used. The sheets were 2 mm in thickness, 25 mm on a hexagonal grid, so that the scan contained
in width and 200 mm in length. The sheets were 185,000 orientations. The data were also analyzed
austenitized at 1273 K for 1.8 ks in Ar+10vol%H2 with the TSL-OIM software.
atmosphere, followed by water-quenching to get
the martensite structure.
The sheets with martensitic microstructure were 3. Results
cold-rolled by 50% reduction in thickness
(equivalent strain; ε ⫽ 0.8) in three passes by the Fig. 1 shows microstructures of the as-quenched
use of a two-high mill with a roll diameter of 310 martensite, which is the starting structure of the
mm. The rolling was performed at a roll peripheral present thermomechanical process. OM micro-
speed of 17.5 m·min⫺1 with machine-oil as lubri- structure (a) represents typical lath martensite in
cant. The cold rolling was successfully done with- low-carbon steels. Although a small amount of
out any cracking. The cold-rolled specimens were proeutectoid ferrite precipitated along some of the
subsequently annealed for 1.8 ks at various tem- prior austenite grain boundaries and the subsurface
peratures ranging from 473 to 973K. regions with thickness of 50 µm revealed equiaxed
Microstructural observations were carried out ferrite due to decarburization during austenitiz-
for the specimens at various stages of the process. ation, 95% of the sheet indicated lath martensite
All specimens were observed from the transverse after water-quench. This is because grain coarsen-
direction (TD) on the longitudinal section which is ing of austenite during austenitization increased
parallel to both the normal direction (ND) and the hardenability of the plain low-carbon steel [23].
rolling direction (RD) of the sheets. The micro- The mean grain size of austenite was evaluated as
structures for optical microscopy (OM) or scanning 270 µm. In the SEM microstructure (b), a fine and
electron microscopy (SEM) were revealed with 3% complicated morphology of packets and blocks of
nital. SEM observations were conducted on JEOL the martensite is observed. TEM microstructure (c)
JSM-5600 scanning electron microscope operated reveals that the martensite laths with width of 0.1–
at 15 kV. The thin foils, which were perpendicular 0.2 µm include large densities of dislocations.
to TD, were prepared for transmission electron Fig. 2(a) shows an OM microstructure of the
microscopy (TEM) by twin-jet electropolishing in 50% cold-rolled specimen. The horizontal and the
R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189 4179

Fig. 2. OM (a) and SEM (b) microstructures of the 0.13wt%C


steel 50% cold-rolled. The starting microstructure was marten-
site. Observed from TD.

Fig. 1. OM (a) , SEM (b) and TEM(c) microstructures of the


as-quenched martensite in the 0.13wt%C steel.
(270 µm), assuming that each austenite grain shape
changes in the same manner as the macroscopic
vertical directions in this figure are parallel to RD change of the rolled sheet. It suggests that the
and ND, respectively. A number of darkly etched darkly etched bands may correspond not only to
bands were observed. They seem to appear along the prior austenite grain boundaries but also to the
prior austenite grain boundaries. However, the packet and/or block boundaries in lath martensite.
mean interval of the bands along ND was 79 µm, At interior of the prior austenite grains, complex
which is smaller than the mean interval of the prior metal flow is observed. In SEM observations
austenite grain boundaries (135 µm) calculated (Fig.2(b)), the details of the morphology at interior
from the mean austenite grain size before rolling were shown. The structure of the cold-rolled speci-
4180 R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189

men could be classified into three kinds of micro-


structures, as represented by the alphabetical
characters (A, B and C in Fig. 2(b));

앫 A: very fine lamellar structure mainly elongated


parallel to RD,
앫 B: irregularly bent lamellar structure,
앫 C: lump of martensite laths with shear bands.

The broken lines in Fig. 2(b) are the boundaries


between the different structures. The area fractions
of the microstructures A, B and C, in the 50%
rolled specimen were 50, 43 and 7%, respectively.
In the SEM micrograph (Fig. 2(b)), the regions cor-
responding to the darkly etched bands in OM (Fig.
2(a)) could be observed along the arrowed bound-
aries. The vicinity of the boundaries exhibited
much more complicated metal flow than other
areas.
The three kinds of microstructures (A, B and C)
could be also distinguished in TEM observations.
Fig. 3(a), (b) and (c) show the TEM morphologies
of the microstructures A, B and C, respectively.
The figures include selected area diffraction (SAD)
patterns taken from the circled areas by the use of
an aperture with a diameter of 1.6 µm. Fig. 3(a)
corresponds to the microstructure A in Fig. 2(b).
This microstructure is composed of the ultrafine
lamella dislocation cells (LDCs). The LDCs elong-
ate mainly along RD, but they have often wavy
flow possibly due to micro shear-banding. The
mean interval of the lamellar boundaries is only 60
nm. The SAD patterns from the LDC structure are
arc-like, which suggests that a number of different
orientations exist within the selected small areas.
Similar lamellar structures have been previously
observed in materials intensely strained by conven- Fig. 3. TEM microstructures (a,b,c) of the 0.13wt%C steel
tional rolling [24–26] or accumulative roll-bonding 50% cold-rolled, corresponding to the microstructure A
(ARB) [14–16,27–29]. Table 1 summarizes the (lamellar dislocation cells; LDC), B (irregularly bent laths; IBL)
characteristics of the lamellar structures in various and C (kinked laths; KL) in Fig. 2, respectively. The starting
microstructure was martensite. Observed from TD.
kinds of deformed materials. The lamellar bound-
aries in the heavily deformed Ni [25] and Al
[14,15,24,26,28,29] have been proved to be the boundaries in the present case was smaller than the
high-angle boundaries (HABs) whose misorien- others formed by strains larger than 4.
tations are larger than 15 °. It should be noted that A prior austenite grain boundary indicated by
the ultrafine LDC structure in the present study was arrows is included in Fig. 3(b). The narrow region
obtained after conventional cold-rolling by a strain just below the boundary reveals the complicated
of only 0.8, and that the mean interval between the morphology and darker contrast which suggests
Table 1
Characteristics of the lamellar deformation structures in various kinds of deformed materials

Material Deformation process and Total strain Boundary spacing Mean misorientation References
temperature

Ni(99.99%) rolling, RT 4.5 about 100 nm about 20 ° [25]


Al(A1200) rolling, RT 5.0 about 300 nm 29.2 ° [24,26]
Al(A1100) ARB, 473 K 4.8 270 nm 37.3 ° [14,15,28,29]
0.13wt%C-steel (ferrite+pearlite starting structure) ARB, RT 4.0 110 nm ring-like SAD pattern [27]
0.13wt%C-steel (Martensite starting structure) rolling, RT 0.8 60 nm ring-like SAD pattern this work
R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189
4181
4182 R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189

high dislocation density. These regions correspond well. Clear grain boundaries surround the ultrafine
with the darkly etched regions in OM (Fig. 2). That grains. On the other hand, some of the regions do
is, the darkly etched bands in Fig. 2(a) are presum- not show the ultrafine grains like the upper-right
ably the regions severely and inhomogeneously area of Fig. 4(b). The block-like region without
deformed due to constraint by pre-existing grain ultrafine grains corresponds to the KL structure.
boundaries. The soft phase (proeutectoid ferrite) Ultrafine grains exist only along the line with the
sometimes existing at prior austenite grain bound- slope of about 45 ° to RD, which was micro-shear
aries would also enhance such concentration of band in the as-deformed state. That is to say, the
severe deformation. LDC structure is seen above KL structure becomes tempered martensite by
the arrowed boundary in Fig. 3(b), while the lower annealing. It can be concluded that the LDC and
region mainly showed the irregularly bent lamella the IBL structures with large local misorientations
(IBL) structure. The IBL structure corresponds to turn into the ultrafine ferrite grains. In addition to
the microstructure B in Fig. 2(b). In the IBL struc- the ultrafine ferrite grains, nano-carbides precipi-
ture, a bunch of laths is bent to orient in various tate uniformly in Fig. 4(b). This is reasonable
directions. The SAD patterns taken from the IBL because martensite (the starting microstructure) is
structure are ring-like, indicating that large local a supersaturated solid solution of carbon. The
misorientations also exist in the IBL. multiphased nano-structure composed of ultrafine
Fig. 3(c) is a TEM micrograph of the kinked lath ferrite grains, nano-carbides and tempered marten-
(KL) structure which corresponds with the micro- site blocks were also observed in the specimen
structure C in Fig. 2(b). The KL is a martensite annealed at 823 K (Fig. 4(c)). Annealing at and
block kinked by micro shear bands. The directions above 873 K causes grain growth to result in
of the laths showing the KL structure tend to be coarse ferrite grains with spheroidized coarse car-
nearly parallel to ND and the shear bands in KL bides (Fig. 4(d)). Only annealing at intermediate
structure are inclined at about 30 ° to RD. This (warm) temperatures produces the UFG structure
suggests that the martensite blocks composed of in the present processing. It should be noted that
the laths parallel to ND are hard to deform during the UFG structure similar to that in intensely
rolling. The SAD pattern taken from the KL struc- strained materials [14,15,24] could be obtained
ture (I) in Fig. 3(c) is nearly a single net pattern, without severe plastic deformation.
reflecting that the KL maintains the martensite lath Detailed crystallographic analysis of the
morphology. It is noteworthy, by the way, that the obtained UFG structure was performed in TEM for
KL structure naturally has a zig-zag shaped outline the specimen annealed at 773 K. Fig. 5 shows the
(prior block boundary). The vicinity of the outline TEM microstructure of the analyzed area (a), and
must suffer complicated plastic deformation to the corresponding map indicating boundary mis-
satisfy compatibility. Actually the lower-right orientations (b). Nearly equiaxed ultrafine ferrite
regions in Fig. 3(c) suggests the presence of a very grains and fine carbides are seen in Fig. 5(a). In
high density of dislocations and the SAD pattern Fig. 5(b), low-angle boundaries whose misorien-
from the region (II) is accordingly ring-like. tations are smaller than 15 degrees are drawn by
Fig. 4 shows TEM microstructures of the speci- narrow lines, while high-angle boundaries with
mens 50% rolled and annealed at various tempera- misorientations larger than 15 degrees are drawn
tures for 1.8 ks. The microstructures of the speci- by bold lines. The dotted lines are the boundaries
mens annealed at 673 K are not so different from whose misorientations were not measured. The
the as-deformed one. The LDC structure can be gray particles in Fig. 5(b) are the carbides. Each
observed in Fig. 4(a), for example. In the specimen misorientation angle (deg.) is also superimposed at
annealed at 773 K (Fig. 4(b)), the equiaxed each boundary. The result indicates that large por-
ultrafine ferrite grains with the mean diameter of tion of the boundaries surrounding the UFGs are
180 nm form in the most of the areas. The shape high-angle boundaries, although there are many
of the ultrafine grains is dominantly equiaxed but low-angle ones as well. The mean misorientation
the elongated grains are sometimes observed as and the fraction of the high-angle boundaries are
R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189 4183

Fig. 4. TEM microstructures of the 0.13wt%C steel 50% cold-rolled and subsequently annealed at 673 K(a), 773 K(b), 823 K(c)
and 873 K(d) for 1.8 ks. Starting microstructure was martensite. Observed from TD.

summarized in Table 2. Seventy percent of the structed from the orientational data of the UFGs in
measured boundaries are high-angle ones, and the Fig. 5. Although the orientation data were obtained
mean misorientation of all the observed boundaries from the limited area (13 µm×7 µm), the pole fig-
is 28.6 °. The boundaries are also classified into ure does not show any strong texture but scattered
two categories in Table 2: the boundaries nearly orientation distribution. However, the grains with
parallel to RD and those nearly perpendicular to similar orientations tend to make small colonies,
RD. 13 boundaries with a slope of about 45 ° to as is shown in Fig. 6(b). This presumably reflects
RD were not classified. The fraction of the high- the block/packet structure of the starting marten-
angle boundaries among the boundaries nearly par- site.
allel to RD was higher than that among the bound-
aries nearly perpendicular to RD. Similar results
have been reported in the elongated ultrafine grains 4. Discussion
or the lamellar boundary structures in heavily
deformed materials [14,15,17,25,27–29]. It was clearly shown in the present study that
Fig. 6(a) shows the {001} pole figure con- simple cold-rolling and warm temperature
4184 R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189

Fig. 5. TEM microstructure (a) and corresponding misorientation map (b) of the 0.13wt%C steel 50% cold-rolled and annealed at
773 K for 1.8 ks. Starting microstructure was martensite. Observed from TD. Misorientation angles (deg.) of the boundaries are also
written in (b).

annealing of martensite can produce a multiphased by severe plastic deformation [6,12–14,27]. The
nano-structure with large misorientations in a low- key of the easy formation of the UFGs in the
carbon steel. It should be noted that not severe present thermomechanical processing must be in
plastic deformation but only 50% rolling reduction the starting structure.
(equivalent strain of 0.8) was applied to the The starting microstructure in the present pro-
material. It has been reported that very high strain cess was a typical lath martensite, as was shown
over strain of 4 is necessary to form UFG structure in Fig. 1. As is well known, lath martensite has a
R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189 4185

Table 2
Mean misorientation and fraction of HAB in the UFG microstructure of the 0.13wt%C steel 50% cold-rolled and annealed at 773
K. The starting microstructure was martensite. Corresponding to the data of Fig. 5

Mean misorientation Fraction of HAB

Whole boundaries 28.6 ° 70% (73/104)


Near parallel to RD 35.0 ° 96% (48/50)
Near perpendicular to RD 20.2 ° 20% (19/41)

Fig. 6. Crystallographic features of the 0.13wt%C steel 50% cold-rolled and annealed for 1.8 ks at 773 K. Starting microstructure
was martensite. (a) {001} pole figure of all the orientations measured in Fig. 5. (b) Orientation map indicating the orientation of
each ultrafine grain by {001} pole figure.

three-level hierarchy in its morphology: (I) Lath; quantitatively, orientation mapping of the as-
single crystal of martensite including high density quenched martensite in the present steel was car-
of lattice defects, (II) Block; aggregation of laths ried out by EBSD method. Fig. 7 shows the results.
with the same crystallographic orientation The measurement was successfully done, and the
(variant), and (III) Packet; aggregation of the image quality (IQ) [22] map (Fig. 7(a)), which
blocks having the same habit plane (for example, reflects quality of the EBSD pattern at each point,
{111}g). Within one prior austenite grain, there can well reconstructs the lath-martensite structure like
appear several packets, because there are four dif- Fig. 1. Fig. 7(b) displays the orientation color map
ferent {111}g planes in austenite. Each packet of the as-quenched martensite. The colors determ-
includes several blocks composed of laths with dif- ined in the stereographic triangle indicated the
ferent variants lying parallel to the same {111}g crystallographic orientation parallel to the plane
plane. According to Kurdjumov and Sachs (K–S) normal in Fig. 7(b). Before making this color map,
relationship, it is possible to have 24 different vari- a clean-up procedure of the data were carried out
ants (orientations) in an identical prior austenite by the grain dilation method in order to eliminate
grain. As Morito et al. [30] pointed out, 83% of the bad points where the analysis failed. In the pro-
the block boundaries and the packet boundaries can cedure of the grain dilation method, first of all,
be high-angle boundaries. It means that martensite “grain” was determined as an aggregate of the
is a kind of fine grained structure subdivided by a measured points surrounded by high-angle bound-
number of high-angle boundaries (block and aries with misorientations larger than 15 °. If the
packet boundaries). number of the measured points constructing an
In order to evaluate the fineness of martensite identical grain was smaller than three (one or two),
4186 R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189

the orientation of the points was changed into the


orientation of the majority of the surrounding
points. This procedure can remove small grains
composed of one or two measured points with little
reliance, and any new high-angle boundary seg-
ments are not made by this method. Fig. 7(b)
clearly shows the fine and complicated mor-
phology of the blocks and the packets with differ-
ent orientations (colors). The boundary map (Fig.
7(c)) was constructed from the data of Fig. 7(b).
The black lines in the figure are the high-angle
boundaries whose misorientations are larger than
15 °. This figure quantitatively indicates that the
lath martensite structure is a kind of fine grained
structure. The total length of the high-angle bound-
aries in this region (200 µm×200 µm) is 24.8 mm.
Assuming that the measured section is the rep-
resentative of the three-dimensional morphology of
this martensite structure, the area of the high-angle
boundaries per unit volume, Sv is 0.62 µm⫺1. It is
known from quantitative microscopy [31] that
there is a simple relationship between Sv and the
mean intercept length (L) for any three-dimen-
sional shape of masses,
2
SV ⫽ (1)
L
From Eq. (1), L ⫽ 3.2 µm was obtained in the
present microstructure. This can be considered as
the mean grain size of the lath martensite structure.
A number of papers have been published about
the formation process of the ultrafine grains during
SPD [27–29,32–34]. Some of these works [27–29]
clearly reported that the original grains are subdiv-
ided into fine crystals with submicrometer sizes
during intense straining. Several researchers
showed that the misorientation of the deformation
induced boundaries (DIBs), which subdivide the
original grain, increases with increasing strain and
finally most of the DIBs become high-angle ones
at high strain [27–29]. That is, grain subdivision
is the essentially important process for ultra-grain
refinement by intense plastic deformation. Tsuji et
al. [27] additionally pointed out the importance of
recovery. The as-deformed DIBs are composed of
or at least decorated by dislocations. Recovery
Fig. 7. Image quality (IQ) map (a), orientation color map (b) turns such unclear boundaries into clear grain
and boundaries map showing high angle boundaries (c) of the
as-quenched martensite in the 0.13wt%C steel.
boundaries, resulting in obvious UFG structure
R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189 4187

[27]. The formation of the ultrafine ferrite grains in would be one of the main reasons for the quick
the present study is the same as that in the intensely grain subdivision.
strained materials described above. That is, the as- In addition to the fine-grained structure, marten-
deformed specimen showed finely subdivided site includes high density of dislocations as well
structure including large local misorientations (Fig. as a number of solute carbon atoms in the as-trans-
3). Annealing at warm temperatures produced the formed state. The dislocation substructures would
clear UFGs (Fig. 4). The difference is the amount also play a role to cause inhomogeneous defor-
of total strain. Only 0.8 of equivalent strain was mation. Solute carbon atoms inhibit dislocation
applied to the present martensite, although very motions and may enhance inhomogeneous defor-
high strain over 4 is necessary to form UFGs in mation. Such effects may be taken into account for
the SPD processes [6,12–14,27]. Apparently, mar- quick grain subdivision as well. Further, solute car-
tensite transformation is not equivalent to SPD. bon atoms uniformly precipitate as fine carbides
The present results suggest that the martensite during warm annealing, which inhibits grain
starting microstructure is effective to subdivide the growth of the ultrafine ferrite. The obtained multi-
grains finely by any means. phase nano-structure in the present process is also
Hansen [35] classified the deformation induced favorable to realize excellent strength-ductility bal-
boundaries into two categories: incidental dislo- ance of the material [38].
cation boundaries (IDBs) and geometrically neces-
sary boundaries (GNBs). The GNBs are more
important than the IDBs for ultra-grain refinement, 5. Conclusions
because it is difficult for the IDBs to get large mis-
orientations even at high strain level [35]. Microstructural evolution during martensite
Although the detailed mechanism of the grain sub- cold-rolling and annealing process was studied in
division by the GNBs is still unclear, it might be an 0.13wt%C steel. The ultrafine grained structure
was formed without severe plastic deformation.
reasonable to consider an analogy to deformation
Formation mechanism of the ultrafine multiphased
banding. Higashida et al. [36] clarified the forma-
structure was discussed mainly from a viewpoint
tion mechanism of well-defined deformation bands
of characteristic features of the martensite starting
(kink band and band of secondary slip) in tensile
structure. The major results are summarized below.
deformation of f.c.c. single crystals. They clearly
showed that the initiation of deformation banding 1. Ultrafine inhomogeneous deformation micro-
is local lattice curvature due to constraint. Though structures were observed in the 50% cold-rolled
they directly showed the effect of the constraint by specimens. The deformation microstructures
the grip of tensile test, constraint effect can be were categorized into three kinds of morpho-
caused by another factors. For example, grain logies: ultrafine lamella dislocation cell (LDC),
boundaries would restrict the plastic deformation irregularly bent lath (IBL) and kinked lath (KL).
near them. Jago and Hansen [37] studied the grain The ultrafine LDCs, which are similar to the
size effects on the deformation microstructure of severely deformed microstructures, with mean
polycrystalline iron and showed that the degree of thickness of 60 nm were the major component
the microstructural inhomogeneities including of the microstructure. Most of the areas with the
deformation bands, slip bands and micro shear LDCs or IBLs have large local misorientations,
bands is higher in fine grains than that in coarse which indicates that ultrafine grain subdivision
ones at the same level of strain. This suggests that was achieved only by 50% cold-rolling.
fine grained structure in the starting material would 2. Annealed specimens showed the equiaxed
accelerate grain subdivision possibly due to con- ultrafine ferrite grains with mean grain size of
straint by the initial grain boundaries. That is, the 180 nm in most areas. The boundary misorien-
fine grained structure with L ⫽ 3.2 µm, as shown tations of the ultrafine grains were quantitatively
in Fig. 7, of the present lath martensite structure evaluated and it was shown that 70% of the
4188 R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189

boundaries are high-angle ones. In addition to [8] Takaki S, Kawasaki K, Kimura Y. In: Mishra RS, Semi-
the ultrafine ferrite grains, nano-carbides pre- ation SL, Suryanarayana C, Thadhani NN, Lowe TC, edi-
tors. Ultrafine Grained Materials. TMS; 2000. p. 247.
cipitated uniformly and tempered martensite [9] Belyakov A, Sakai Y, Hara T, Kimura Y, Tsuzaki K. Met-
blocks were also observed in some areas. all Mater Trans 2001;32A:1769.
3. The main reason for the quick grain subdivision [10] Belyakov A, Sakai T, Miura H, Kaibyshev R. ISIJ Int
in the present thermomechanical process is that 1999;39:593.
the martensite starting structure is itself a fine [11] Saito Y, Tsuji N, Utsunomiya H, Sakai T. Acta Mater
1999;47:579.
grained structure. The mean grain size of the [12] Saito Y, Tsuji N, Utsunomiya H, Sakai T, Hong RG.
martensite blocks/packets surrounded by high- Scripta Mater 1998;39:1221.
angle boundaries was quantitatively evaluated [13] Tsuji N, Saito Y, Utsunomiya Y, Tanigawa S. Scripta
as 3.2 µm by means of EBSD analysis. The fine Mater 1999;40:795.
grained structure causes constraint effect of [14] Tsuji N, Saito Y, Ito Y, Utsunomiya H, Sakai T. In:
Mishra RS, Semiatin SL, Suryanarayana C, Thadhani NN,
plastic deformation to enhance inhomogeneous
Lowe TC, editors. Ultrafine Grained Materials. TMS;
deformation, in other words, grain subdivision. 2000. p. 207.
High dislocation density and a number of solute [15] Ito Y, Tsuji N, Saito Y, Utsunomiya H, Sakai T. J of Jpn
carbon atoms in martensite are also expected to Inst Metals 2000;64:429.
play a role to enhance grain subdivision. [16] Tsuji N, Shiotsuki K, Saito Y. Mater Trans JIM
1999;40:765.
[17] Tsuji N, Ueji R, Saito Y. Materia Japan 2000;39:961.
[18] Tsuji, N., Ueji, R., Saito, Y., Koizumi, Y., Minamino, Y.,
6. Acknowledgments 2001. Proc. of the 22nd Risø Int. Symp. on Materials
Science, edited by Dinesen, A., Eldrup. M., Juul Jensen,
D., Linderoth, S., Pederson, T.B., Pryds, N. H., Schrøder
The present study was financially supported by Pedersen, A. and Wert, J. A., Risø National Laboratory,
Industrial Technology Research Grant Program in Denmark, 407.
’01 from New Energy and Industrial Technology [19] Tsuji N, Ueji R, Minamino Y, Saito Y. Scripta Mater
Development Organization (NEDO) of Japan 2002;46:305.
[20] Zaefferer S. J of Appl Crystall 1999;33:10.
(project ID 01A23025d) and 2001 Steel Research [21] Adams BL, Wright SI, Kunze K. Metall Trans
Grant from Iron and Steel Institute of Japan (ISIJ). 1993;24A:819.
The steel used in this study was provided by NKK [22] Wright SI. J Comput Assist Microsc 1993;5(3):207.
Co. Ltd. The semi-automatic online analysis of [23] Honeycombe RWK, Bhadeshia HKDH. STEELS Micro-
Kikuchi-line in TEM was done by the aid of Dr. structure and Properties, 2nd ed. Edward Arnold, 1995.
[24] Hansen N, Juul Jensen D. Phil Trans R Soc Lond
S.Morito, Prof. T. Furuhara, and Prof. T. Maki of
1999;A357:1447.
Kyoto University. These supports are gratefully [25] Hansen N, Huang X, Hughes DA. Mat Sci Engng
appreciated by the authors. 2001;A317:3.
[26] Liu, Q., Huang, X., Lloyd, D.J., Hansen, N., 1999. Proc.
of the 4th Int. Conf. on Recrystalization and Related
Phenomena, edited by T. Sakai and Suzuki, H. G., JIM,
References Japan, 315.
[27] Tsuji, N., Ueji, R., Ito, Y., Saito, Y., 2000. Proc. of the
[1] Shaw LL. J of Metals 2001;52(12):41. 21st Risø Int. Symp. on Mater. Sci., edited by Hansen, N.,
[2] Bridgman PW. Studies in large plastic flow and fracture. Huang, X., Juul Jensen, D., Lauridsen, E.M., Leffers, T.,
York: McGraw-Hill, 1952. Pantleon, W., Sabin, T.J. and Wert, J.A., Risø National
[3] Horita Z, Smith D, Furukawa M, Nemoto M, Valiev RZ, Laboratory, 607.
Langdon TG. J of Mater Res 1996;11:1880. [28] Huang, X., Tsuji, N., Minamino, Y., Hansen, N., 2001.
[4] Segal VM. Mater Sci Engng 1995;A197:157. Proc. of the 22nd Risø Int. Symp. on Materials Science,
[5] Valiev RZ, Ivanisenko YV, Rauch EF, Baudelet B. Acta edited by Dinesen, A., Eldrup. M., Juul Jensen, D., Linder-
Mater 1996;44:4705. oth, S., Pederson, T.B., Pryds, N. H., Schroder Pedersen,
[6] Iwahashi Y, Horita Z, Nemoto M, Lasngdon TG. Acta A. and Wert, J. A., Risø National Laboratory, Denmark,
Mater 1997;45:4733. 255.
[7] Shin DH, Kim BC, Park K, Kim Y-S, Park K. Acta [29] Huang, X., Tsuji, N., Hansen, N., Minamino, Y., Mater.
Mater 2000;48:2247. Sci. Engng. A 2002, in press.
R. Ueji et al. / Acta Materialia 50 (2002) 4177–4189 4189

[30] Morito, S., Tanaka, H., Furuhara, T., Maki, T., 1999. Proc. [34] Belyakov A, Sakai T, Miura H, Kaibyshev R. Phil Mag
of the 4th Int. Conf. on Recrystallization and Related Lett 2000;80:771.
Phenomena (Rex99), edited by Sakai, T. and Suzuki, H. [35] Hansen N. Metall Mater Trans A 2001;32A:2917.
G., Jpn. Inst. of Metals, 295. [36] Higashida K, Takamura J, Narita N. Mater Sci Engng
[31] DeHoff RT, Rhines FN. Quantitative Microscopy. 1986;81:239.
McGraw-Hill Publishing, 1968. [37] Jago RA, Hansen N. Acta Metall 1986;34:1711.
[32] Shin DH, Kim BC, Park K, Choo WY. Acta Mater [38] Tsuji N, Ito Y, Koizumi Y, Minamino Y, Saito Y. In: Zhu
2000;48:3245. YT, Langdon TG, Mishra RS, Semiatin SL, Saran ML,
[33] Shin DH, Kim Y-S, Lavernia EJ. Acta Mater Lowe TC, editors. Ultrafine Grained Materials II. TMS;
2001;49:2387. 2002. p. 389.

You might also like