You are on page 1of 10

Acta Materialia 54 (2006) 1279–1288

www.actamat-journals.com

Crystallographic features of lath martensite in low-carbon steel


Hiromoto Kitahara a, Rintaro Ueji b, Nobuhiro Tsuji a,*
, Yoritoshi Minamino a

a
Department of Adaptive Machine Systems, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
b
Department of Advanced Materials Science, Faculty of Engineering, Kagawa University, 2217-20 Hayashi-cho, Takamatsu 761-0396, Japan

Received 14 January 2005; received in revised form 1 November 2005; accepted 1 November 2005
Available online 6 January 2006

Abstract

Electron backscattering diffraction with field-emission scanning electron microscopy was used to analyze crystallographically the lath
martensite structure in a 0.20% carbon steel. The crystallographic features of the lath martensite structure, of the order of the prior aus-
tenite grain size or larger, were clarified. Although the orientations of the martensite crystals were scattered around the ideal variant
orientations, the martensite in this steel maintained the Kurdjumov–Sachs (K–S) orientation relationship. The procedures of the crys-
tallographic analysis of the martensite (ferrite) phase with the K–S orientation relationship were explained in detail. Variant analysis
showed that all 24 possible variants did not necessarily appear within a single prior austenite grain and that all six variants did not nec-
essarily appear within each packet. Specific combinations of two variants appeared within local regions (sub-blocks), indicating a strict
rule for variant selection. Prior austenite grain boundaries and most of the packet boundaries were clearly recognized. However, it was
difficult to determine the block boundaries within the sub-blocks.
 2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Martensitic phase transformation; Packet; Block; Misorientation; Orientation relationship

1. Introduction i.e., laths, blocks, and packets [2]. The mechanical proper-
ties of martensitic steels are greatly affected by the micro-
When the austenite (c) phase with a face-centered cubic structures. For example, the strength of low-carbon steels
(fcc) structure in steels is quickly quenched to low temper- with lath martensite structure increases as the packet size
ature, a displacive transformation forms a new phase with decreases [2]. The packet size has also been reported to
body-centered cubic (bcc) or body-centered tetragonal affect the toughness of martensitic steels [5,6]. Therefore,
structure. The transformation and the resulting phase are studying the morphological and crystallographic character-
called the martensitic transformation and martensite, istics of martensite in steels is important.
respectively [1–3]. Numerous studies have been conducted Nevertheless, the crystallographic features of martensite
on the martensitic transformation and martensite in steels. of the order of the austenite grain size or larger have not
Furthermore, martensite in steels is of practical impor- yet been clarified due to the limitations in experimental
tance, especially for high-strength applications. When aus- techniques. Electron diffraction methods in transmission
tenite transforms into martensite, the crystallographic electron microscopy (TEM) are conventionally used to
orientation relationships are maintained [1–4]. The orienta- analyze the crystallography of martensite [1–3]. However,
tion relationship leads to a characteristic lath or lenticular TEM can only observe limited areas and it is difficult to
morphology [1,2]. For example, the lath martensite in low- determine accurately the orientation from electron diffrac-
carbon steels has a three-level hierarchy in its morphology, tion analysis, except when using Kikuchi-line diffraction
[7]. Consequently, crystallographic data for martensite
*
Corresponding author. Tel.: +81 6 6879 7434/7502; fax: +81 6 6879
structures have not yet been clarified for sizes of the order
7434/7500. of the prior austenite grains (several tens to about a hun-
E-mail address: tsuji@ams.eng.osaka-u.ac.jp (N. Tsuji). dred micrometers).

1359-6454/$30.00  2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2005.11.001
1280 H. Kitahara et al. / Acta Materialia 54 (2006) 1279–1288

Orientation imaging microscopy (OIM) [8] using elec- a FE-SEM instrument (Philips XL30S-FEG) equipped
tron backscattering diffraction (EBSD) with scanning elec- with a FE-type gun operated at 20 kV. Orientation map-
tron microscopy (SEM) is a potential method for ping for the FE-SEM/EBSD data was performed with a
overcoming these problems. Recently, several investiga- 100 nm step size. TSL-OIM software was used for the ori-
tions have analyzed the martensite microstructures using entation measurement and analysis.
SEM/EBSD [9–11]. Ueda et al. [9] used EBSD to analyze Thin foils were prepared for TEM analysis by twin-jet
the crystallographic variants of the large martensite crys- electropolishing in the same solution described above.
tals with plate (or lenticular) morphology transformed The TEM observations were conducted using a Hitachi
from bicrystals of austenite in a Fe–32 at.% Ni alloy. Wil- H-800 microscope operated at 200 kV.
son et al. [10] applied EBSD to 0.12% C–3.28% Ni and
HSLA steels to determine the volume fraction of lath mar- 3. Procedure for crystallographic analysis
tensite, another common morphology of martensite in
steels [1–4], using a lower pattern quality from martensite Lath martensite appears in carbon steels with a carbon
due to the high density of lattice defects. Morito et al. content less than 0.6 wt.% [2]. The Kurdjumov–Sachs
[11] systematically studied the crystallography of martens- (K–S) orientation relationship between lath martensite
ite in various carbon steels. However, these studies are just and austenite is known in plain carbon steels [1–3],
the beginning and the complete crystallographic view of although other orientation relationships have also been
martensite in steels is still unclear. reported [3,11]. It has been shown that the K–S orientation
We recently studied the crystallography of plate (or len- relationship is true for 0.18 wt.% C –0.006 wt.% Si –
ticular) martensite with the Nishiyama–Wassermann (N– 0.02 wt.% Mn steel, which has a similar carbon content
W) orientation relationship in a 28.5 at.% Ni steel using to the steel used in the present investigation [11]. Therefore,
field-emission SEM (FE-SEM) with EBSD. The morphol- the 0.20 wt.% C steel under investigation is expected to
ogy and crystallography of the plate martensite crystals have a lath martensite structure with a K–S orientation
were successfully obtained for a large area [12]. The present relationship. The K–S orientation relationship between
investigation involves the crystallographic analysis of typi- austenite (c) and martensite (a 0 ) is expressed as
cal lath martensite in a low-carbon steel using FE-SEM/
ð1 1 1Þ ==ð0 1 1Þ 0 ½1 0 1 ==½1 1 1 0 .
c a c a
EBSD. Lath martensite in steels is of practical importance
since nearly all martensitic steels have the lath-type mor- Because of symmetry, cubic systems have 24 equivalent
phology. However, lath martensite includes a high density crystallographic variants in martensite, which evolve from
of lattice defects, especially dislocations [1–3,11], which a single crystal (grain) of austenite if the K–S orienta-
makes EBSD analysis difficult. The present study aims tion relationship is maintained. Table 2 summarizes the
firstly to confirm the applicability of EBSD for the study
of the crystallography of lath martensite, secondly to Table 2
describe the detailed procedure for crystallographic analysis The 24 crystallographic variants for the K–S orientation relationship
of the lath martensite from the EBSD data, and thirdly to Variant Plane parallel Direction parallel
clarify the crystallographic and morphological characteris- V1 ð1 1 1Þc ==ð0 1 1Þa0 ½1 0 1c //½1 1 1a0
tics of the lath martensite crystals on a relatively large scale. V2 ½1 0 1c //½1 1 1
 0
a
V3 ½0 1 1c //½1 1 1a0
V4  //½1 1 1
½0 1 1  0
2. Experimental c a
V5 ½1 1 0c //½1 1 1a0
V6 ½1 1 0 //½1 1 1
c
 0
a
A plain low-carbon steel (0.20 wt.% C steel), whose
V7 ð1 1 1Þc ==ð0 1 1Þa0 ½1 0 1c //½1 1 1a0
chemical composition is shown in Table 1, was used. Steel V8 ½1 0 1c //½1 1 1a0
sheets (2 mm thick) were austenitized at 1473 K for 1.8 ks V9 ½1 1 0c //½1 1 1a0
and subsequently quenched in ice brine (258–263 K) to V10 ½1 1 0c //½1 1 1a0
cause the martensite transformation. This procedure V11 [0 1 1]c//½1 1 1a0
resulted in sheets having a martensite structure, which were V12 [0 1 1]c//½1 1 1a0
used for the EBSD measurements. The observed section of V13 ð1 1 1Þc ==ð0 1 1Þa0 ½0 1 1c //½1 1 1a0
the specimen was mechanically polished and then electro- V14 ½0 1 1c //½1 1 1a0
polished. The electropolishing was conducted using a V15 ½1 0 1c //½1 1 1a0
V16 ½1 0 1c //½1 1 1a0
900 ml CH3COOH + 100 ml HClO4 solution at 284 K V17 [1 1 0]c//½1 1 1a0
and 20 V. The EBSD measurements were conducted using V18 [1 1 0]c//½1 1 1a0
V19 ð1 1 1Þc ==ð0 1 1Þa0 ½1 1 0c //½1 1 1a0
Table 1 V20 ½1 1 0c //½1 1 1a0
Chemical composition (wt.%) of the 0.20 wt.% carbon steel studied V21 ½0 1 1c //½1 1 1a0
V22 ½0 1 1c //½1 1 1a0
C Si Mn P S sol. Al Fe V23 [1 0 1]c//½1 1 1a0
0.195 0.19 0.35 0.016 0.01 0.021 Bal. V24 [1 0 1]c//½1 1 1a0
H. Kitahara et al. / Acta Materialia 54 (2006) 1279–1288 1281

crystallographic variants. The variants are denoted as V1– in the K–S orientation relationship can be calculated and
V24. Fig. 1 illustrates the six variants (V1–V6) that form on are given in Table 3.
an identical (1 1 1) austenite plane. The triangle and rectan- Using the orientation transformation matrices T1–T24,
gles indicate the (1 1 1) plane of austenite (fcc) and the the orientations of the 24 different variants of martensite
(0 1 1) plane of martensite (bcc), respectively. There are 24 (M1–M24) that are transformed from a single crystal
variants since austenite has four different {1 1 1} planes. (grain) austenite with any orientation can be calculated.
The change in the crystallographic orientation through Fig. 2 plots the 24 martensite variants assuming that the
the martensite transformation is expressed by a matrix orientation of the prior austenite grain is (0 0 1)[1 0 0]. Each
equation: number in Fig. 2 indicates the particular martensite
M ¼ TA; ð1Þ variant.
The misorientation angle and rotation axis between two
where T is the orientation transformation matrix, and M different variants are determined using the orientations for
and A are the orientation matrices composed of three the 24 martensite variants (M1–M24) calculated from Eq.
orthogonal unit vectors that show crystallographic orienta- (1). The rotation matrix (R) between the different variants
tions for the austenite and martensite phases, respectively is calculated from the following:
[12].
M2 ¼ RM1 . ð3Þ
As an example, Table 2 shows that for V1 the directions
[1 1 1]c and ½1 0 1c in austenite are transformed into ½0 1 1a0 The misorientation angle and rotation axis are calculated
and ½ 1
1 1a0 in martensite, respectively, which can be from the elements of the rotation matrix, R [12,13]. Table
expressed as 4 summarizes the axis/angle relationships between V1
0 1 0 1
0:707 0:408 0:577 0:577 0:816 0:000
B C T1 B C Table 3
@ 0:000 0:816 0:577 A ! @ 0:577 0:408 0:707 A
The 24 orientation transformation matrices for the K–S orientation
0:707 0:408 0:577 c 0:577 0:402 0:707 a0 relationship
0 1 0 1
0:742 0:667 0:075 0:667 0:742 0:075
Therefore, the orientation transformation matrix, T1, is T 1 ¼ @ 0:650 0:742 0:167 A T 13 ¼ @ 0:742 0:650 0:167 A
calculated by 0:167 0:075 0:983 0:075 0:167 0:983
0 1 0 1
0:075 0:667 0:742 0:667 0:075 0:742
T1 ¼ M1 A1
1 . ð2Þ T 2 ¼ @ 0:167 0:742 0:650 A T 14 ¼ @ 0:742 0:167 0:650 A
0:983 0:075 0:167 0:075 0:983 0:167
Similarly, the 24 different orientation transformation 0 1 0 1
0:667 0:075 0:742 0:075 0:667 0:742
matrices that correspond to the crystallographic variants @
T3 ¼ 0:742 0:167 0:650 A T 15 ¼ @ 0:167 0:742 0:650 A
0:075 0:983 0:167 0:983 0:075 0:167
0 1 0 1
0:667 0:742 0:075 0:742 0:667 0:075
T 4 ¼ 0:742 0:650 0:167 A
@ T 16 ¼ 0:650 0:742 0:167 A
@
0:075 0:167 0:983 0:167 0:075 0:983
0 1 0 1
0:075 0:742 0:667 0742 0:075 0:667
T 5 ¼ 0:167 0:650 0:742 A
@ T 17 ¼ 0:650 0:167 0:742 A
@
0:983 0:167 0:075 0:167 0:983 0:075
0 1 0 1
0:742 0:075 0:667 0:075 0:742 0:667
T 6 ¼ @ 0:650 0:167 0:742 A T 18 ¼ @ 0:167 0:650 0:742 A
0:167 0:983 0:075 0:983 0:167 0:075
0 1 0 1
0:075 0:667 0:742 0:742 0:075 0:667
T 7 ¼ @ 0:167 0:742 0:650 A T 19 ¼ @ 0:650 0:167 0:742 A
0:983 0:075 0:167 0:167 0:983 0:075
0 1 0 1
0:742 0:667 0:075 0:075 0:742 0:667
T 8 ¼ @ 0:650 0:742 0:167 A T 20 ¼ @ 0:167 0:650 0:742 A
0:167 0:075 0:983 0:983 0:167 0:075
0 1 0 1
0:742 0:075 0:667 0:667 0:742 0:075
@
T 9 ¼ 0:650 0:167 0:742 A T 21 ¼ @ 0:742 0:650 0:167 A
0:167 0:983 0:075 0:075 0:167 0:983
0 1 0 1
0:075 0:742 0:667 0:667 0:075 0:742
T 10 ¼ 0:167 0:650 0:742 A
@ T 22 ¼ 0:742 0:167 0:650 A
@
0:983 0:167 0:075 0:075 0:983 0:167
0 1 0 1
0:667 0:742 0:075 0:075 0:667 0:742
T 11 ¼ @ 0:742 0:650 0:167 A T 23 ¼ 0:167 0:742 0:650 A
@
0:075 0:167 0:983 0:983 0:075 0:167
Fig. 1. Six crystallographic variants (V1–V6) for the K–S orientation 0 1 0 1
0:667 0:075 0:742 0:742 0:667 0:075
relationship that evolves on a (1 1 1) austenite plane. The triangle and T 12 ¼ @ 0:742 0:167 0:650 A T 24 ¼ @ 0:650 0:742 0:167 A
rectangles indicate the (1 1 1) plane of austenite (c: fcc) and the (0 1 1) plane 0:075 0:983 0:167 0:167 0:075 0:983
of martensite (a 0 : bcc), respectively.
1282 H. Kitahara et al. / Acta Materialia 54 (2006) 1279–1288

Morito et al. [11] reported the axis/angle pairs between


different variants of the K–S orientation relationship,
which are consistent with the present results. In the K–S
relationship, there are certain combinations of the variants
that show special coincidence site lattice (CSL) relation-
ships, such as R3 and R11. Table 4 also indicates the
CSL relationships when BrandonÕs criterion is satisfied
[14]. There are 10 possible misorientation angles, i.e.,
10.53, 14.88, 20.61, 21.06, 47.11, 49.47, 50.51,
51.73, 57.21, and 60.00, between the variants that satisfy
the K–S relationship. Most of the angles exceed 15, which
is typically the criterion between low- and high-angle
boundaries. Even the minimum misorientation, 10.53, is
relatively large, which indicates that the neighboring mar-
tensite variants have large-angle boundaries in the K–S ori-
entation relationship.

4. Results and discussion

Fig. 3 shows a TEM image of the lath martensite struc-


Fig. 2. {0 0 1} Pole figure showings the orientations of the 24 martensite
variants transformed from a single crystal austenite with a (0 0 1)[1 0 0] ture in the 0.20 wt.% C steel. Typically, the widths of the
orientation that maintain the K–S orientation relationship. The numbers martensite laths are less than 1 lm. The TEM image con-
indicate the martensite variants. firms that the lath martensite structure has a high density
of dislocations. Fig. 4 illustrates a typical lath martensite
structure with a three-level hierarchy in its morphology
and other variants in the K–S orientation relationship [2]: martensite lath, block, and packet. The martensite lath
using the procedure previously described for the N–W ori- is a single crystal of martensite with a high density of lattice
entation relationship [12]. defects. The block is aggregations of the laths with the

Table 4
Crystal orientations of 24 martensite variants transformed from a (0 0 1)[1 0 0] oriented austenite maintaining the K–S orientation relationship, and
crystallographic relationships between V1 and other variants
Variant no. Crystal orientations of martensite variants Misorientation Misorientation axis between V1 CSL*
( h k l ) [ u v w ] angle from V1 () and other variants

V1 ( 0.075 0.167 0.983 ) [ 0.742 0.650 0.167 ] – [ – – – ] –


V2 ( 0.650 0.167 0.742 ) [ 0.167 0.983 0.074 ] 60.00 [ 0.577 0.577 0.577 ] R3
V3 ( 0.742 0.167 0.650 ) [ 0.667 0.075 0.742 ] 60.00 [ 0.000 0.707 0.707 ] –
V4 ( 0.075 0.167 0.983 ) [ 0.667 0.742 0.075 ] 10.53 [ 0.000 0.707 0.707 ] R1
V5 ( 0.667 0.075 0.742 ) [ 0.075 0.983 0.167 ] 60.00 [ 0.000 0.707 0.707 ] –
V6 ( 0.667 0.075 0.742 ) [ 0.742 0.167 0.650 ] 49.47 [ 0.000 0.707 0.707 ] R11
V7 ( 0.167 0.650 0.742 ) [ 0.983 0.167 0.075 ] 49.47 [ 0.577 0.577 0.577 ] R19b
V8 ( 0.167 0.075 0.983 ) [ 0.650 0.742 0.167 ] 10.53 [ 0.577 0.577 0.577 ] R1
V9 ( 0.075 0.667 0.742 ) [ 0.167 0.742 0.650 ] 50.51 [ 0.615 0.186 0.767 ] –
V10 ( 0.075 0.742 0.667 ) [ 0.983 0.167 0.075 ] 50.51 [ 0.739 0.462 0.490 ] –
V11 ( 0.167 0.075 0.983 ) [ 0.742 0.667 0.075 ] 14.88 [ 0.933 0.354 0.065 ] R1
V12 ( 0.742 0.167 0.650 ) [ 0.667 0.075 0.742 ] 57.21 [ 0.357 0.603 0.714 ] –
V13 ( 0.167 0.075 0.983 ) [ 0.742 0.667 0.075 ] 14.88 [ 0.354 0.933 0.065 ] R1
V14 ( 0.167 0.650 0.742 ) [ 0.075 0.742 0.667 ] 50.51 [ 0.490 0.462 0.739 ] –
V15 ( 0.167 0.742 0.650 ) [ 0.983 0.075 0.167 ] 57.21 [ 0.738 0.246 0.628 ] –
V16 ( 0.167 0.075 0.983 ) [ 0.650 0.742 0.167 ] 20.61 [ 0.659 0.659 0.363 ] –
V17 ( 0.075 0.742 0.667 ) [ 0.167 0.650 0.742 ] 51.73 [ 0.659 0.363 0.659 ] –
V18 ( 0.075 0.667 0.742 ) [ 0.983 0.075 0.167 ] 47.11 [ 0.719 0.302 0.626 ] –
V19 ( 0.742 0.075 0.667 ) [ 0.650 0.167 0.742 ] 50.51 [ 0.186 0.767 0.615 ] –
V20 ( 0.667 0.075 0.742 ) [ 0.075 0.983 0.167 ] 57.21 [ 0.357 0.714 0.603 ] –
V21 ( 0.075 0.167 0.983 ) [ 0.667 0.742 0.075 ] 20.61 [ 0.955 0.000 0.296 ] –
V22 ( 0.650 0.167 0.742 ) [ 0.742 0.075 0.667 ] 47.11 [ 0.302 0.626 0.719 ] –
V23 ( 0.650 0.167 0.742 ) [ 0.167 0.983 0.075 ] 57.21 [ 0.246 0.628 0.738 ] –
V24 ( 0.075 0.167 0.983 ) [ 0.742 0.650 0.167 ] 21.06 [ 0.912 0.410 0.000 ] –
*
Coincidence site lattice is indicated when BrandonÕs criterion [14] is satisfied.
H. Kitahara et al. / Acta Materialia 54 (2006) 1279–1288 1283

ture should be a fine-grained structure that is divided by


many large-angle boundaries (packet and block
boundaries).
Fig. 5(a) shows an image quality (IQ) map and Fig. 5(b)
an inverse pole figure (IPF) color map obtained from the
FE-SEM/EBSD measurements over a large area
(55 · 65 lm2) of the 0.20 wt.% C steel after the martensite
transformation. The IQ map (Fig. 5(a)) reflects the quality
of the Kikuchi line for each measurement. For example, the
image quality is low, and thus the gray-scale level is dark at
the grain boundaries, because it is difficult to obtain clear
Kikuchi-line diffraction in these regions. Therefore, the
IQ map looks like a conventional SEM image. The IQ
map (Fig. 5(a)) clearly reveals a typical lath martensite
Fig. 3. TEM image of the lath martensite in the 0.20 wt.% C steel. structure and prior austenite grain boundaries, which
implies that FE-SEM/EBSD provides reliable orientation
mapping despite the high density of dislocations in the lath
martensite structure (Fig. 3). Thus, FE-SEM/EBSD can be
a useful tool for crystallographic analysis even for lath
martensite in low-carbon steels. The colors in Fig. 5(b) cor-
respond to the crystallographic orientation normal to the
observed plane, as indicated by the stereographic triangle
in the inset. Fig. 5(b) clearly shows the detailed morphol-
ogy and crystallographic features of the lath martensite
structure. The boundaries in Fig. 5(b) are drawn for misori-
entation between adjacent points greater than 10 since the
misorientation calculations (Table 4) imply that all the
packets and block boundaries should have misorientations
larger than 10. The figure indicates that martensite has a
fine-grained structure. The mean grain size of the lath mar-
tensite structure is evaluated by the equivalent grain size,
which is calculated by the total length of the high-angle
boundaries, whose misorientations are larger than 15,
per unit area [15]. Consequently, the equivalent grain size
of the lath martensite structure is 2.1 lm, which is similar
to the previous result for 0.13% C steel martensite
(3.2 lm) [15].
Fig. 6(a) shows an IPF color map of the lath martensite
structure for a limited area of Fig. 5(b). The boundaries
(black line) are drawn when the misorientation between
adjacent points is greater than 10. The martensite variants
Fig. 4. Microstructural hierarchy of the lath martensite structure. are accurately analyzed to identify the prior austenite grain
boundaries. The white and red lines in Fig. 6(a) indicate the
prior austenite grain boundaries and packet boundaries,
respectively. Fig. 6(b) shows a {0 0 1} pole figure that
same crystallographic orientation (variant). The packet is depicts the orientation of the martensite crystals from the
aggregations of the blocks with the same {1 1 1}c plane in prior austenite grain surrounded by the white line in
austenite. Several packets can appear within a single prior Fig. 6(a). The pole figure (Fig. 6(b)) indicates the orienta-
austenite grain since there are four different {1 1 1}c planes. tion spreads, which are consistent with the color gradient
Each packet includes several blocks that are composed of in Fig. 6(a). The orientation spread from the ideal variant
the martensite laths with different variants, but the same orientations was also observed in the plate martensite
{1 1 1}c planes. According to Fig. 1 and Table 2, a maxi- structure with the N–W orientation relationship in an
mum of six different variants (blocks) can appear within Fe–28.5 at.% Ni steel [12], but the orientation distribution
an identical packet. The misorientation of the packet in the present lath martensite is much more scattered than
boundaries and block boundaries should be larger than that in the plate martensite. The ideal orientations of the 24
10.53, the minimum misorientation angle between the dif- variants (Fig. 2) are rotated to coincide with the actual ori-
ferent variants (Table 4). Briefly, the lath martensite struc- entations of the measured martensite and are superimposed
1284 H. Kitahara et al. / Acta Materialia 54 (2006) 1279–1288

Fig. 5. Typical orientation imaging maps obtained from the FE-SEM/EBSD measurement of the lath martensite structure in the 0.20 wt.% C steel.
(a) Image quality (IQ) map and (b) inverse pole figure (IPF) color map. Colors of the martensite crystals in (b) agree with the orientations perpendicular to
the observed plane, which are indicated in the stereographic triangle. Black lines in (b) show the boundaries with misorientation angles larger than 10.

as the diamond symbols in Fig. 6(b). The concentrations aries are not recognized from the 10 boundary criterion
near the ideal orientations based on the K–S orientation within the sub-blocks, which suggests orientation scattering
relationship are obvious, which confirms that the martens- within the sub-block region that includes two specific vari-
ite in the 0.20 wt.% C steel under investigation maintains ants. Although most block boundaries are not indicated
the K–S orientation relationship. However, the orientation within the sub-blocks, most of the packet boundaries could
distribution is not discrete, but has a certain degree of scat- be drawn from the variant notations and Table 2. The red
tering around the ideal orientations. In addition, it is deter- lines in Fig. 6(a) are the packet boundaries. However, the
mined that all 24 variants do not necessarily appear within region between V2 and (V20 and V23) sub-block at the
the austenite grain. For example, V16 and V18 do not right corner of the prior austenite that must have a packet
appear in Fig. 6. boundary is not drawn in the 10 criterion, which indicates
Variant analysis of the lath martensite in Fig. 6(a) was that the orientation does not suddenly change between the
conducted using Fig. 6(b). For example, for the violet different packets. Fig. 6(a) also shows that all six possible
region (the white arrow) in Fig. 6(a), the orientation is also variants do not necessarily appear within each packet.
drawn in violet in Fig. 6(b). The spread of the orientation is Morito et al. [11] reported that all six variants are present
consistent with V1 and V4 in Fig. 6(b), which indicates that within each packet in the lath martensite structures of car-
a combination of different variants (V1 and V4) is included bon steels, which is inconsistent with our results. This dif-
within this region. The variant number for each region is ference may be due to the coarseness of the austenite
shown in Fig. 6(a). Numerous regions include certain com- grains, since Morito et al. [11] used fairly coarse austenite
binations of two specific variants (V1 and V4, V2 and V5, grain sizes. It is plausible that the grain size of the prior
V3 and V6, V7 and V10, V8 and V11, V9 and V12, V14 and austenite grain affects the packet size and that the packet
V17, V19 and V22, V20 and V23, V21 and V24), but some size determines the appearance of the sub-blocks.
areas also have blocks with one variant. There is a specific Fig. 7(a) shows a skeleton image of Fig. 6(a), which is
rule for combining two variants, which is a rotation axis of composed of the prior austenite grain boundaries (black
Æ0 1 1æ and an angle of 10.53 (Table 4). Morito et al. [11] lines), packet boundaries (red lines), and variant notations.
reported the local regions composed of two specific vari- In Fig. 7(b), the boundaries are based on the misorienta-
ants in 0.0026–0.61 wt.% C steels, although their EBSD tion between adjacent points of the EBSD data. Three dif-
observations were limited to a small part of a prior austen- ferent criteria are used to draw the boundary: the green
ite grain with a coarse grain size of 200 lm. They called lines indicate a misorientation larger than 10, the blue
them ‘‘two specific K–S variant groups (sub-blocks)’’. lines that between 5 and 10, and the red lines that
The relationship that they observed in the sub-blocks is between 2 and 5. Fig. 7(a) and (b) clearly shows that most
identical to that in the present study described above. Thus, of the packet boundaries have misorientations larger than
it is concluded that there is a strict rule (Æ0 1 1æ – 10.53) for 10. However, as previously mentioned, only a blue bound-
the variant selection within the sub-blocks. Block bound- ary exists between V2 and the sub-block of (V20 and V23)
H. Kitahara et al. / Acta Materialia 54 (2006) 1279–1288 1285

Fig. 6. (a) Inverse pole figure (IPF) color map of the lath martensite structure in the 0.20 wt.% C steel. The white and red lines indicate the prior austenite
grain boundaries and packet boundaries, respectively. (b) {0 0 1} Pole figure shows the orientations of the lath martensite crystals within the prior austenite
that corresponds to the area surrounded by the white line in the IPF map (a). The symbols and numbers indicate the variant numbers.

in the right corner of the prior austenite grain, which is cer- within each sub-block. Even if the smallest misorientation
tainly a packet boundary. The misorientation between the criterion is used, there are no continuous red boundaries
ideal V2 and V20 or between the ideal V2 and V23 is 50.51 in Fig. 7(b).
or 47.11, respectively. This disagreement corresponds to a Fig. 8(a) shows the point-to-origin (point A) misorien-
significant orientation scattering around the ideal variant tation profile and Fig. 8(b) the point-to-point misorienta-
orientations in the real lath martensite crystals. Fig. 7 also tion profile between points A and B in Fig. 7(b) in order
clearly shows that there are no large-angle boundaries to understand the orientation change between and within
1286 H. Kitahara et al. / Acta Materialia 54 (2006) 1279–1288

Fig. 7. Boundary maps of the lath martensite structure in the same area as Fig. 6(a). (a) Prior austenite grain boundaries (bold black lines) and block
boundaries (red lines). (b) Boundaries using three different criteria.

the sub-blocks. Fig. 8(a) shows discrete changes in the ori- 60.0 for all the combinations between (V19 and V20),
entation between (V19 and V22) and (V20 and V23) sub- (V19 and V23), (V20 and V22), and (V22 and V23). The
blocks. Fig. 8(b) shows that the corresponding block orientation change is small within each sub-block
boundaries have a misorientation of approximately 60, (Fig. 8(a)). Fig. 8(b) shows some peaks close to 10 within
which corresponds to the calculated misorientation the sub-blocks. Although the ideal misorientation between
between the sub-blocks and the ideal misorientation of two variants within each block is 10.53, several bound-
H. Kitahara et al. / Acta Materialia 54 (2006) 1279–1288 1287

Fig. 8. Misorientations along the black line between A and B in Fig. 7. (a) Point-to-origin misorientation and (b) point-to-point misorientation.

aries within the sub-blocks have misorientations between 1. FE-SEM/EBSD is a useful tool for microstructure/ori-
5 and 10 (Fig. 8(b)). This characteristic orientation fea- entation analysis of lath martensite in low-carbon steels
ture is derived from the formation mechanism and the despite a high density of lattice defects in the martensite
nature of martensite. The strict variant selection rule crystal. Detailed crystallographic analysis of the lath
(10.53 around Æ0 1 1æ) and the change in the orientation martensite was achieved of the order of the prior austen-
within the sub-blocks will be discussed in the light of ite grain size or larger, and characteristic features of the
transformation strain in a future study. The present inves- lath martensite structure were determined.
tigation has clarified the various features of the lath mar- 2. EBSD analysis confirmed that the lath martensite in the
tensite structure of the order of the austenite grain size or steel investigated maintained the K–S orientation rela-
larger. tionship. However, the orientation of the martensite
crystals was broadly distributed around the ideal orien-
5. Conclusions tations of the expected variants. Procedures of the crys-
tallographic analysis of the lath martensite with the K–S
The lath martensite structure in a 0.20 wt.% C steel was orientation relationship have been detailed.
crystallographically analyzed on a relatively large scale 3. Variant analysis of the lath martensite was successfully
using EBSD with FE-SEM. The main points are summa- conducted from the EBSD data. All 24 variants did
rized below. not necessarily appear within an austenite grain. The
1288 H. Kitahara et al. / Acta Materialia 54 (2006) 1279–1288

packet/block structures within a prior austenite grain References


were accurately determined. In most cases, the packets
were clearly recognized from the orientation analysis. [1] Nishiyama Z. Martensite transformation. Tokyo: Maruzen; 1971.
Within each packet, all six possible variants did not nec- [2] Krauss G. Steels heat treatment and processing principles. Materials
Park (OH): ASM International; 1990.
essarily appear and the number of the variants seems to [3] Honeycombe RWK, Bhadeshia HKDH. Steels microstructure and
depend on the size of the packet. Sub-blocks composed properties. 2nd ed. London: Edward Arnold; 1995.
of two specific variants were observed in the lath mar- [4] Bhadeshia HKDH. Worked examples in the geometry of crys-
tensite structure. A strict rule for the variant selection tals. London: Institute of Metals; 1987.
in the sub-blocks, namely 10.53 around Æ1 1 0æ axis, [5] Inoue T, Matsuda S, Okamura Y, Aoki K. Trans JIM 1970;11:36.
[6] Swarr T, Krauss G. Metall Trans A 1976;7A:41.
was confirmed. It was determined that the orientation [7] Zaefferer S. J Appl Crystallogr 2000;33:10.
profile within each sub-block was not discrete but fairly [8] Adams BL, Wright SI, Kunze K. Metall Trans 1993;24A:819.
scattered. Thus, block boundaries with detectable mis- [9] Ueda M, Yasuda HY, Umakoshi Y. Acta Mater 2001;49:3421.
orientation were scarce within the sub-blocks. [10] Wilson AW, Madison JD, Spanos G. Scr Mater 2001;45:1335.
[11] Morito S, Tanaka H, Konishi R, Furuhara T, Maki T. Acta Mater
2003;51:1789.
[12] Kitahara H, Ueji R, Ueda M, Tsuji N, Minamino Y. Mater Charact
Acknowledgment 2005;54:378.
[13] Humphreys FJ, Hatherly M. Recrystallization and related annealing
phenomena. Oxford (NY): Pergamon; 1995.
The authors are grateful to the TORAY scientific foun- [14] Brandon DG. Acta Metall 1966;14:1479.
dation for financial support. [15] Ueji R, Tsuji N, Minamino Y, Koizumi Y. Acta Mater 2002;50:4177.

You might also like