25.12.2012 Views

View - IMM - RWTH Aachen University

View - IMM - RWTH Aachen University

View - IMM - RWTH Aachen University

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Abstract<br />

Nucleation and variant selection during the a–c–a phase<br />

transformation in microalloyed steel<br />

I. Lischewski, G. Gottstein ⇑<br />

Institute of Physical Metallurgy and Metal Physics, <strong>RWTH</strong> <strong>Aachen</strong> <strong>University</strong>, 52056 <strong>Aachen</strong>, Germany<br />

Received 19 August 2010; received in revised form 8 November 2010; accepted 8 November 2010<br />

Available online 4 December 2010<br />

This study addresses variant selection during the a–c–a phase transformation in recrystallized, low-carbon, microalloyed steel. Since<br />

variant selection was found to occur during incipient stages of transformation, the investigation focused mainly on c-grain nucleation<br />

from the a phase. To obtain comprehensive microstructure information, the sample was characterized by high-temperature electron<br />

backscatter diffraction in a scanning electron microscope and additional three-dimensional serial sectioning. It is shown that preferred<br />

nucleation at a grain boundary near a triple junction requires an orientation relationship close to the Kurdjumov–Sachs (K–S) correspondence<br />

to both adjacent a grains (nucleation rule I). Furthermore, a low disorientation of a {1 1 0} bcc crystallographic plane to<br />

the grain boundary plane favors nucleation (nucleation rule II). The results suggest that the nucleation of c grains is favored at triple<br />

junctions by a low surface energy of the nucleus. A corresponding model is proposed. On this basis, variant selection of different ferrite<br />

orientations and the product textures of the complete transformation cycle were successfully predicted for a random grain boundary<br />

plane distribution.<br />

Ó 2010 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.<br />

Keywords: Microalloyed steel; Phase transformation; Crystallographic texture; Variant selection<br />

1. Introduction<br />

Available online at www.sciencedirect.com<br />

Acta Materialia 59 (2011) 1530–1541<br />

The production of steel sheet involves many processing<br />

steps which are accompanied by a change in phase composition,<br />

texture and microstructure. These changes strongly<br />

influence the material properties. Low-carbon, microalloyed<br />

steel at high temperatures is usually hot-rolled in<br />

the austenite (c) regime. This is commonly accompanied<br />

by the formation of a pronounced crystallographic texture.<br />

During subsequent cooling, the austenite–ferrite phase<br />

transformation takes place. This reversible phase transformation<br />

causes a change in crystal structure, but is also<br />

associated with a dramatic change in microstructure and<br />

texture. A review of transformation textures was given by<br />

Ray et al. [1]. The crystallography of the phase transformation<br />

reveals a specific orientation relationship between aus-<br />

⇑ Corresponding author.<br />

E-mail address: Gottstein@imm.rwth-aachen.de (G. Gottstein).<br />

1359-6454/$36.00 Ó 2010 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.actamat.2010.11.017<br />

www.elsevier.com/locate/actamat<br />

tenite (parent phase) and ferrite (product phase). For<br />

carbon steels, this crystallographic correspondence typically<br />

follows the Kurdjumov–Sachs (K–S) orientation relationship,<br />

or close to it [2–5]. A specific orientation<br />

relationship can be used to predict the product texture.<br />

The same holds for the back transformation (a–c). However,<br />

experiments show that not all crystallographic variants<br />

are generated equally frequently, so that the<br />

computed and measured transformation textures do not<br />

coincide. This is attributed to variant selection, but the<br />

rules and reasons for variant selection in recrystallized<br />

materials are not known. As the material properties are<br />

strongly influenced by microstructure and texture, a basic<br />

understanding of the underlying principles of variant selection<br />

is indispensable for texture control during processing.<br />

Owing to experimental difficulties, variant selection has<br />

been investigated mostly at room temperature in the ferritic<br />

regime after previous deformation of the high-temperature<br />

austenite phase. Therefore, respective models had to


assume a copper-type rolling or cube-type recrystallization<br />

texture of austenite prior to transformation to ferrite. To<br />

explain the variant selection of deformed steels, the activity<br />

of slip systems [6,7], the Bain strain [2,8] or the elastic interaction<br />

of transformation events [9] was considered. However,<br />

variant selection is also observed for recrystallized<br />

austenite (Fig. 1), which is less pronounced than in the<br />

deformed state, but distinct and reproducible.<br />

Brückner and Gottstein proposed associating the variant<br />

selection in recrystallized steel with the Bain strain<br />

and residual stresses in recovered areas [3], but because<br />

of limitations in their experimental techniques, these<br />

assumptions could not be confirmed. The development of<br />

high-temperature orientation imaging, however, opened<br />

new avenues for observing phase transformations with spatial<br />

resolution and for obtaining information on local orientations.<br />

This allowed the mechanisms of phase<br />

transformations to be investigated and the underlying<br />

physics of variant selection to be revealed.<br />

Most previous investigations of variant selection phenomena<br />

pertained to the c–a phase transformation,<br />

whereas little attention was paid to forward transformation.<br />

However, since the transformation may occur in both<br />

directions during processing, the whole transformation<br />

cycle must be investigated [3,10].<br />

To acquire information on the location and orientation<br />

of the product phase with respect to the parent phase, local<br />

orientation measurements need to be conducted at the<br />

transformation temperature. Therefore, for this study a<br />

laser-powered heating stage for a scanning electron microscope<br />

was developed which would allow rapid heating and<br />

cooling as well as concurrent electron backscatter diffraction<br />

(EBSD) measurements. With such a device the whole<br />

transformation cycle a ? c and c ? a can be investigated<br />

in situ. This study focuses on the a ? c transformation<br />

which has been addressed only infrequently so far.<br />

I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541 1531<br />

Previous investigations on the same material revealed<br />

that the newly developed c-phase texture did not significantly<br />

change with progressing transformation [11].<br />

Accordingly, variant selection must already occur during<br />

the nucleation period. Furthermore, it was established that<br />

the K–S relationship which was originally developed for<br />

the martensitic transformation holds for the a–c–a phase<br />

transformation as well. Typically, a small deviation from<br />

the ideal K–S relationship is observed [12].<br />

The goal of this work was to reveal and to analyze all<br />

factors that cause a texture change and variant selection<br />

during the reversible a–c–a phase transformation in recrystallized<br />

microalloyed steel. The required local information<br />

of the parent and product orientation was obtained by<br />

high-temperature in situ EBSD and three-dimensional<br />

(3D) serial sectioning. These results were analyzed with<br />

respect to the underlying mechanisms of variant selection.<br />

The relations derived were used to predict the transformation<br />

texture. A model is presented to rationalize the<br />

observed preference of nucleus orientations.<br />

2. Experimental<br />

2.1. Material<br />

The material investigated was a commercial hot band of<br />

a microalloyed low-carbon steel. The chemical composition<br />

is shown in Table 1. Initially, the samples of this material<br />

were cold rolled in a reversing manner to 50% thickness<br />

reduction. The cold-rolled samples were annealed at<br />

700 °C for 5 h in a vacuum furnace, in order to obtain a<br />

stable microstructure for transformation annealing. Grain<br />

growth was suppressed during transformation annealing<br />

by the presence of stable micro carbo-nitrides. The sample<br />

size was 10 7mm 2 with thickness 1 mm and average<br />

grain size 7 lm.<br />

Fig. 1. Variant selection for the orientation u 1 = 75, U = 58, u 2 = 50 (allowed scattering of 5° disorientation): (a) corresponding 24 K–S variants in a<br />

{2 0 0}-pole figure; (b) experimental variant distribution in a {2 0 0}-pole figure.


1532 I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541<br />

Table 1<br />

Chemical composition of the microalloyed low-carbon steel (wt.%).<br />

C Si Mn P S Al N Cu Cr Ni Sn Mo Ti Nb<br />

0.05 0.28 0.82 0.014 0.001 0.039 0.006 0.03 0.05 0.03 0.003 0.01 0.08 0.034<br />

After stabilization annealing, the metallographic preparation<br />

was carried out. The samples were mechanically polished<br />

and finally electropolished (A2 Struers, U =50V,<br />

t = 12 s). Details of sample preparation procedures are<br />

given elsewhere [13].<br />

2.2. High-temperature in situ EBSD<br />

To obtain all the necessary local information during the<br />

a–c–a phase transformation, EBSD measurements at elevated<br />

temperatures had to be carried out. For this purpose,<br />

a laser-powered heating stage was developed at the Institute<br />

of Physical Metallurgy and Metal Physics of <strong>RWTH</strong><br />

<strong>Aachen</strong> <strong>University</strong>. Details of its setup and operation are<br />

given elsewhere [14]. The basic features are summarized<br />

below.<br />

The hot stage was mounted on a JEOL JSM-6100 scanning<br />

electron microscope. Attached to the microscope was<br />

a standard EBSD-Detector System Nordlys by HKL with a<br />

70° tilted specimen stage for EBSD analysis. The EBSDdetector<br />

contained a four-quadrant forward scatter detector<br />

framing the phosphor-screen, which provided the orientation<br />

contrast (OC) images for enhancement of the<br />

microstructure information by EBSD.<br />

The hot stage consisted of three basic components<br />

(Fig. 2a and b). A SiC sample holder comprised the main<br />

body, which was illuminated by a commercial continuous<br />

diode IR laser with wavelength 810 nm and maximum<br />

power output 100 W. The incident IR laser light was emitted<br />

from an optical fiber. The specimen was attached to the<br />

SiC sample holder by a tungsten clamp (Fig. 2(4)).<br />

The temperature of various attachments (EBSD-detector,<br />

secondary electron-detector, hot stage) in the microscope<br />

chamber was recorded by thermocouples. Three<br />

thermocouples monitored the hot stage temperature (heating<br />

shield, sample holder) and controlled temperature and<br />

(a) rear outer heating (b)<br />

incoming<br />

4<br />

1<br />

3<br />

shield<br />

4<br />

cooling<br />

conduit<br />

5<br />

front outer heating shield<br />

(copper, passive cooled)<br />

4<br />

outgoing cooling<br />

conduit<br />

heating rate. The laser power output was controlled by an<br />

external computer linked via an RS232 interface using Labview<br />

8.0. The hot stage could be operated at any tilting<br />

angle and at a working distance as low as 15 mm. Typically,<br />

6–8 bands were used to determine the crystal orientation<br />

from the EBSD patterns. The orientation imaging of<br />

an area 250 180 lm 2 took 45 min.<br />

2.3. 3D analysis of grain boundary plane<br />

Exact information on the grain boundary plane (GBP)<br />

orientation can only be achieved from 3D images. Such<br />

measurements were conducted by serial sectioning with<br />

the following procedure. A ferrite sample was marked at<br />

several locations of the sample surface with a Vickers micro<br />

hardness indenter. The microstructure of the surface was<br />

measured by EBSD and OC imaging. After this first measurement,<br />

the sample surface was carefully mechanically<br />

polished. Three polishing steps were performed by using<br />

a Buehler/Phoenix 4000 polishing wheel at very low contact<br />

pressure. The embedded sample was polished for<br />

2 min with 1 lm diamond paste and 5 min with 0.25 lm<br />

diamond paste. The total thickness reduction after each<br />

polishing step was 0.6 lm, 1.3 lm and 2.4 lm. These values<br />

were easily obtained from the change in the indent shape<br />

with each polishing step. The use of several micro hardness<br />

markers allowed precise measurement of the thickness<br />

reduction and of the planarity of the sample surface.<br />

After each polishing step, the sample was placed back<br />

into the sample holder of the scanning electron microscope<br />

in the same position as before, and an OC image (step size<br />

0.1 lm) was generated. The different OC maps were superimposed<br />

(Fig. 3) using commercial image editing software<br />

(Paint Shop Pro7) to determine the position of the grain<br />

boundary traces. The different microhardness markers<br />

served as reference points and allowed matching of the<br />

Fig. 2. Design of the hot stage: (a) front view to reveal the basic outer parts (b) cross section of the stage interior: (1) copper shield (water cooled); (2)<br />

tantalum shield (not cooled); (3) SiC sample holder; (4) specimen mount–tungsten clamp; (5) heating stage copper base (active water cooled); (6) optical<br />

fiber guidance; (7) goniometer stage adapter. Overall dimensions 8.2 3.5 3.6 cm 3 .<br />

1<br />

3<br />

6<br />

7<br />

2<br />

1<br />

5<br />

4


image position for each OC map. From this experimental<br />

procedure, the GBP orientation near the sample surface<br />

was obtained.<br />

3. Results<br />

Fig. 3. Correlation of two OC maps.<br />

3.1. Transformation texture development<br />

The texture development in the course of the a–c phase<br />

transformation was measured to determine the essential<br />

period where variant selection occurred.<br />

To obtain good statistics for texture calculation, especially<br />

for the low transformed volume fractions, many samples<br />

were investigated by high-temperature in situ EBSD.<br />

The c-phase orientations were binned with respect to their<br />

f(g)<br />

~ {112}<br />

<br />

8<br />

6<br />

4<br />

2<br />

0<br />

45 o<br />

~ {123}<br />

<br />

60 o<br />

75 o<br />

(a) -fibre (max. density)<br />

I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541 1533<br />

~ {011}<br />

<br />

90 o<br />

2<br />

90 o<br />

/<br />

1<br />

55 o<br />

45 o<br />

20 o<br />

60 o<br />

75 o<br />

(b) -fibre orientation<br />

volume fraction (1–10% = 5%, 15–25% = 20%, 35–<br />

45% = 40%), and the texture was calculated (Fig. 4).<br />

The results demonstrate that no significant texture<br />

change occurred with progressing a–c phase transformation.<br />

The small deviations were attributed to slight variations<br />

in the ferrite microstructure and texture of different<br />

specimens. Hence, the product texture and therefore also<br />

variant selection were essentially controlled by nucleation.<br />

Further investigations into variant selection phenomena<br />

therefore concentrated on the nucleation stage.<br />

3.2. Nucleation<br />

The ferrite (a) to austenite (c) phase transformation in<br />

steel proceeds by nucleation of the new phase and its<br />

growth. Partially transformed microstructures were<br />

recorded at 907 °C by orientation imaging via EBSD.<br />

The early stage of transformation allowed the nucleation<br />

sites of c nuclei to be located exactly in the a microstructure,<br />

and thus to establish the orientation relationship of<br />

a nucleus to its parent phase. An investigation of the early<br />

stages of transformation (17,000 austenite grains) showed<br />

that nucleation of the c phase occurred predominantly at<br />

grain boundary triple junctions (Table 2).<br />

Furthermore, it can be seen (Fig. 5) that the K–S orientation<br />

relationship was quite often met. Previous<br />

detailed investigations on a similar steel rendered the<br />

information that slight deviations from the exact K–S<br />

relationship were usually found [12,15]; this was also confirmed<br />

in the current study. A spread of the K–S toward<br />

Table 2<br />

Fraction of nucleation sites of c grains (%).<br />

Triple junction Grain boundary Inner grain<br />

90.3 9.6 0.1<br />

90 o<br />

2<br />

{011}<br />

<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 o<br />

{011} {011}<br />

<br />

Fig. 4. Texture development during progressing a–c phase transformation in terms of fcc fiber plots: austenite volume fraction 5% (s), 20% (h), 40% (D),<br />

100% ( ).<br />

f(g)<br />

30 o<br />

-fibre<br />

60 o<br />

{011}<br />

<br />

=45 o<br />

= 0<br />

2<br />

o<br />

90 o<br />

1<br />


1534 I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541<br />

907°C Austenite:4%<br />

=50 µm; Map1; Step=0.8 µm; Grid180x160<br />

Fig. 5. EBSD map of partially transformed microstructure during a–c<br />

phase transformation (4% austenite at 907 °C): ferrite, gray; austenite, red;<br />

high-angle grain boundaries, black lines, 60° h111i twins, blue lines; phase<br />

boundaries (colored according to deviation from ideal K–S relationship):<br />

0–5° (white), 5–10° (yellow), 10–15° (green), 15–20° (purple), >20° (black).<br />

the Nishiyama–Wassermann relationship was usually<br />

observed. Reasons for this deviation have been discussed<br />

elsewhere [16,17]. For the sake of simplicity, the results<br />

were associated with a K–S relationship. If a deviation<br />

from the ideal K–S relationship of up to 15° was allowed,<br />

many c grains had a K–S relationship to more than one a<br />

mother grain. Usually, two a mother grains were<br />

observed. To find out whether this was a systematic effect<br />

or just incidental, the experimental results were compared<br />

with a simulated nucleus distribution. For this, the experimental<br />

data were evaluated from the EBSD measurements<br />

with an in-house developed computer code. The<br />

maximum allowed deviation from the K–S relationship<br />

was set to 15°. A statistical nucleus distribution was<br />

generated by distributing 2000 experimental c grains randomly<br />

in a representative measured ferrite microstructure.<br />

This rendered the information that the actual measured<br />

nucleation frequency with two K–S mother grains was<br />

twice as high as that in the statistical distribution<br />

(Fig. 6). Also, the measured frequency of three mother<br />

grains was significantly higher than predicted by statistics.<br />

It is noted that the real number of nuclei with two or<br />

three a mother grains might have been even higher than<br />

measured, because potential ferrite mother grains might<br />

have been located below the sample surface. The results<br />

illustrate that the nucleation of c grains during the a–c<br />

phase transformation preferentially occurs with two a<br />

mother grains, each of which has a K–S orientation relationship<br />

to the nucleus. This will be called nucleation rule<br />

I in the following. These experimental results actually<br />

confirm the hypothesis of Tomida et al. [18], who were<br />

the first to propose that a product grain needs more than<br />

one parent grain. On this basis, they were able to compute<br />

the transformation texture adequately.<br />

frequency [%]<br />

75<br />

70<br />

65<br />

60<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

40.8<br />

71.7<br />

3.3. Grain boundary plane<br />

The experimental procedure to determine the GBP orientation<br />

was described in Section 2.3. The main problem<br />

was to obtain the combined information of the GBP orientation<br />

and the nucleation site of the c grains during the<br />

reversible a–c phase transformation. Because in situ 3D<br />

measurements at elevated temperatures were not possible,<br />

a different strategy was used. A ferrite sample was annealed<br />

for a long time (35 h/780 °C vacuum) in order to establish a<br />

stable microstructure. This sample was then annealed to a<br />

partly transformed state and, to avoid microstructural<br />

changes, a fast (10 min map time) EBSD measurement<br />

was conducted at constant temperature (907 °C). After<br />

the measurement, the sample was rapidly cooled to room<br />

temperature to conserve the general features of the hightemperature<br />

microstructure. At room temperature, a second<br />

EBSD measurement was conducted. This EBSD map<br />

served as reference for the 3D measurements and contained<br />

information on how the microstructure had changed from<br />

the two-phase high-temperature microstructure (Fig. 7).<br />

Only those triple junctions and grain boundaries were<br />

investigated where c-phase nucleation was found during<br />

the high-temperature anneal, and the surrounding a grains<br />

did not show a significant change after cooling to room<br />

temperature. This allowed the orientation of the observed<br />

c nuclei to be associated with the orientation of the GBP.<br />

The influence of the GBP on nucleation was investigated<br />

by considering three grain boundaries and nine triple junctions<br />

where nucleation of a c grain at high temperatures<br />

was found. Fig. 8 shows a typical example of an OC<br />

map. It is a part of the EBSD map in Fig. 7b, which was<br />

measured at RT. In this example, a c grain had nucleated<br />

at a triple junction and revealed a K–S relationship to both<br />

a grains A and B. It can be seen that all necessary information,<br />

such as grain orientations (from EBSD), nucleation<br />

site, inclination and position of the GBP is available. From<br />

this information, it was derived that all GBPs at the triple<br />

junction were associated with close-packed planes with<br />

48.9<br />

25.8<br />

Experimental values<br />

Calculated values<br />

1<br />

2<br />

3<br />

number of α-grains at a triple junction with a K-S relationship<br />

Fig. 6. Frequency of statistically expected and experimentally measured<br />

K–S relationships of c grains at triple junctions.<br />

10.3<br />

2.5


50 µm<br />

RD<br />

respect to the K–S relationship {1 1 0}a//{1 1 1}c of each<br />

a–c grain combination. Because of the typical deviation<br />

from the ideal K–S relationship, the crystallographic<br />

{1 1 0} planes of the ferrite were taken as reference for<br />

the disorientation to the GBP. Presumably, this plane<br />

was the location from where the c grain nucleated.<br />

The results of the analysis of grain boundaries and triple<br />

junctions are summarized in Table 3, where the orientations<br />

of a grains with a K–S relationship to the c grain<br />

are shown. The c-grain orientation and the related crystallographic<br />

variant according to K–S are listed. The crystallographic<br />

{1 1 0} plane for this variant and the calculated<br />

disorientation to the GBP is also indicated. As a result,<br />

I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541 1535<br />

(a) (b)<br />

Fig. 7. EBSD maps of (a) a partially transformed ferrite sample at 907 °C and (b) the ferrite sample (same position) at room temperature (coloring as in<br />

Fig. 5); white circle indicates the triple junction of the example in Fig. 8.<br />

grain C<br />

10µm<br />

~70°<br />

~60°<br />

~50°<br />

14° from TD<br />

25° from TD<br />

grain A<br />

grain B<br />

6° from RD<br />

Fig. 8. OC map of an investigated triple junction with A–C as adjacent grains (the position of this example is marked in Fig. 7b): angles 50°, 60° and 70°<br />

define the inclination of the three GBPs (arrows indicate direction of inclination); angles 25°, 6° and 14° define the position of each grain boundary trace<br />

related to TD and RD.<br />

50 µm<br />

RD<br />

RD<br />

in general one of the two K–S mother grains had a small<br />

disorientation of its {1 1 0} plane to the GBP. This<br />

{1 1 0} plane was defined by the calculated variant. Hence,<br />

it is concluded that the preferred nucleation at a triple junction<br />

proceeds by nucleation at a GBP with a K–S relationship<br />

to two adjoining grains. This hypothesis is supported<br />

by the measurements. The observed disorientation of the<br />

close-packed planes and the GBP was in the range 0–15°<br />

(GB1–GB3 in Table 3); 15° was used as the cut-off because,<br />

in this range, crystallographic deviations from the ideal lattice<br />

can be compensated by dislocation arrangements such<br />

as low-angle grain boundaries. The required alignment of<br />

the GBP with a {1 1 0} plane is termed nucleation rule II.


1536 I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541<br />

Table 3<br />

Results of the investigated triple junctions and grain boundaries.<br />

TJ/GB a mother grain orientation Euler<br />

angles (°)<br />

The disorientations determined between the f110g a planes<br />

and the GBP may contain some uncertainty because, especially<br />

at high inclination of the GBP, only a rough definition<br />

of the disorientation angle was possible.<br />

Nevertheless, the results obtained confirmed the expected<br />

tendency. At TJ8 in Table 3, three mother grains were<br />

observed with a K–S relation to the c nucleus. Also for this<br />

case, the aforementioned correlation held true. TJ9 represents<br />

a c-grain nucleation only with one mother grain. This<br />

case, as an exception, violates nucleation rule I, but may be<br />

specially promoted by only a small disorientation of the<br />

f110g a plane to the GBP.<br />

It is noted that the condition of a specific GBP orientation<br />

does not define a variant selection rule by itself. However,<br />

in combination with the requirement that this {1 1 0}<br />

plane is selected as the coherent phase boundary, nucleation<br />

rule II imposes another restriction on potential crystallographic<br />

variants and thus also constitutes a variant<br />

selection rule.<br />

3.4. Prediction of variant selection in ferrite<br />

c-grain orientation Euler angles<br />

(°)<br />

The information obtained, outlined in Section 3.3, can be<br />

used to predict the preferred K–S variants during transformation,<br />

and therefore also the product textures during a–c<br />

phase transformation. The procedure is demonstrated here,<br />

K–S var Crystallographic {1 1 0}<br />

plane<br />

Disorientation to<br />

GBP<br />

GB1 A: 184.1 8.8 33.2 290.9 43.6 9.3 3 ( 110) 60°<br />

B: 2.0 44.8 67.3 8 ( 1 10) 15°<br />

GB2 A: 228.4 33.4 45.2 253.6 18.3 65.2 13 (1 0 1) 16°<br />

B: 165.2 45.4 79.3 5 (1 10) 62°<br />

GB3 A: 331.2 44.6 50.0 36.0 36.9 80.1 5 (1 10) 9°<br />

B: 318.5 53.0 42.8 20 (0 11) 69°<br />

TJ1 A: 192.3 41.6 38.3 106.6 38.0 28.8 3 ( 110) 18°<br />

B: 11.6 36.9 29.9 17 (0 1 1) 53°<br />

TJ2 A: 185.9 8.6 31.4 56.3 44.0 68.7 12 ( 101) 20°<br />

B: 289.6 49.9 49.2 12 ( 101) 90°<br />

TJ3 A: 312.8 42.0 51.5 43.0 44.8 74.2 20 (0 11) 42°<br />

B: 117.5 25.3 82.8 9 (1 0 1) 80°<br />

TJ4 A: 157.0 37.4 50.3 139.5 32.8 17.4 21 (0 1 1) 7°<br />

B: 357.8 17.8 54.9 2 (1 1 0) 5°<br />

TJ5 A: 193.9 43.8 51.4 280.8 21.6 56.8 14 (1 0 1) 73°<br />

B: 20.5 46.4 59 21 (0 1 1) 8°<br />

TJ6 A: 55.7 49.4 45.4 174.6 38.1 3.6 12 ( 101) 16°<br />

B: 8.1 32.1 40.2 16 ( 10 1) 63°<br />

TJ7 A: 8.3 31.7 40.1 239.4 23.1 63.9 1 (1 1 0) 10°<br />

B: 157.6 22.7 57.3 5 (1 10) 73°<br />

TJ8 A: 180.3 23.3 40.9 266.4 24.8 62.8 19 (0 11) 44° (A–B)/18° (A–C)<br />

B: 30.5 36.3 32.7 6 (1 10) 40° (B–A)/43° (B–C)<br />

C: 183.2 35.6 50.6 18 (0 1 1) 60° (C–A)/52° (C–B)<br />

TJ9 A: 346.1 20.0 43.5 113.1 24.2 28.8 2 (1 1 0) 6<br />

as an example, for the variant selection of two a-phase orientations.<br />

To obtain the transformation texture, this procedure<br />

is applied to the entire ferrite texture. From the abcc<br />

fiber, orientations with Euler angles u1 =0°, U =30°,<br />

u2 =45° and u1 =60°, U =55°, u2 =45° from the cbcc<br />

fiber were taken. Nucleation rule I requires that, for nucleation<br />

near a triple junction, two a grains with a K–S relationship<br />

to the new c grain are necessary, in general. For<br />

the calculation, a typical ferrite microstructure measured<br />

by EBSD was taken. This microstructure contained<br />

3000 triple junctions.<br />

For each of the three a grains at a triple junction, the 24<br />

variants (Table 4) of the K–S relationship were calculated.<br />

Thereafter, the misorientation of all 24 variants to the three<br />

a grains was calculated. If the misorientation between two<br />

variants of two a grains was


educed to one variant in the majority of cases. Hence, for<br />

selection of the actual variant, the influence of the GBP<br />

between two a grains is also important. For the computation<br />

of the transformation texture, it was assumed in the<br />

first approximation that the GBP distribution in the investigated<br />

steel was random, and therefore did not influence<br />

the product texture. The calculation and prediction of variant<br />

selection of different a orientations is then reduced to<br />

nucleation rule I. Since nucleation rule II was neglected, at<br />

some triple junctions several possible c orientations would<br />

meet the required correlation. In these cases, all such variants<br />

were taken into account.<br />

The product was filtered with regard to any a orientation<br />

of interest. For better statistics of the variant distribution<br />

for the two selected orientations, a small deviation (7°)<br />

about the ideal position was allowed. The experimental<br />

data were obtained in a similar way. Several hundred partly<br />

transformed microstructures were imaged using high-temperature<br />

in situ EBSD, and the newly developed c grains<br />

were assigned (K–S 10° dev.) to their a-mother orientations.<br />

In this way, a product list was generated which contained<br />

information on the distribution of the selected<br />

variants to all a orientations and also the texture information<br />

on the austenite phase. This list was then filtered with<br />

respect to the above specified a orientations (7° dev.).<br />

The results (Fig. 9) demonstrate overall good agreement<br />

of the measured and predicted variant distributions.<br />

Apparently, the variant selection is already well predicted<br />

using only nucleation rule I. However, some of the predicted<br />

variants show some conspicuous deviations from<br />

the experimental data. This may be caused by a slightly<br />

non-random GBP distribution. Also, other factors, such<br />

I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541 1537<br />

Table 4<br />

Correlation of the crystallographic planes and directions of the 24 K–S variants.<br />

No. Ferrite plane/direction Austenite plane/direction Rotation axis<br />

1 (1 1 0) [ 1 1 1] (1 1 1) [1 1 0] [1 12]<br />

2 (1 1 0) [1 11] ( 11 1) [1 1 0] [1 1 2]<br />

3 ( 110)[ 1 11] (11 1) [ 1 1 0] [1 1 2]<br />

4 ( 110)[111] ( 1 1 1) [ 1 1 0] [1 1 2]<br />

5 (1 1 0) [1 1 1] ( 1 1 1) [1 10] [ 1 12]<br />

6 (1 10)[ 1 11] (11 1) [1 10] [ 1 1 2]<br />

7 ( 1 10)[ 1 1 1] (1 1 1) [ 1 10] [ 11 2]<br />

8 ( 1 10) [1 11] ( 11 1) [ 1 10] [ 112]<br />

9 (1 0 1) [ 1 11] (11 1) [1 0 1] [1 2 1]<br />

10 (1 0 1) [ 1 1 1] (1 1 1) [1 0 1] [ 1 21]<br />

11 ( 101)[1 11] ( 11 1) [ 1 0 1] [1 2 1]<br />

12 ( 101)[111] ( 1 1 1) [ 1 0 1] [ 12 1]<br />

13 (1 0 1) [1 1 1] ( 1 1 1) [1 0 1] [1 21]<br />

14 (1 0 1) [1 11] ( 11 1) [1 0 1] [ 1 2 1]<br />

15 ( 10 1) [ 1 1 1] (1 1 1) [ 10 1] [1 2 1]<br />

16 ( 10 1) [ 1 11] (11 1) [ 10 1] [ 121]<br />

17 (0 1 1) [1 11] ( 11 1) [0 1 1] [2 1 1]<br />

18 (0 1 1) [1 11] (11 1) [0 1 1] [2 11]<br />

19 (0 11)[ 1 1 1] (1 1 1) [0 11] [ 2 1 1]<br />

20 (0 1 1) [1 1 1] ( 1 1 1) [0 11] [ 211]<br />

21 (0 1 1) [1 1 1] ( 1 1 1) [0 1 1] [2 1 1]<br />

22 (0 1 1) [ 1 1 1] (1 1 1) [0 1 1] [2 1 1]<br />

23 (0 1 1) [1 11] ( 11 1) [0 1 1] [ 2 11]<br />

24 (0 1 1) [ 1 11] (11 1) [0 1 1] [ 21 1]<br />

as inhomogeneities or the different statistical base of the<br />

experimental and predicted data, could be important. This<br />

exercise was also extended to other orientations. Again, the<br />

predicted variant distributions showed good agreement<br />

with experiments.<br />

3.5. Texture prediction of the a–c–a phase transformation<br />

3.5.1. General procedure<br />

With the analysis presented above, one can scan the ferrite<br />

microstructure for triple junctions that comply with<br />

nucleation rule I to determine the orientation spectrum of<br />

the c nuclei. Since nucleation rule II was neglected, several<br />

triple junctions allowed for more than one possible c orientation.<br />

For calculation of the transformation texture, all<br />

possible variants at a triple junction were taken into<br />

account. For simplicity, the orientations of the different<br />

variants were averaged. This reflects the assumption that<br />

each possible product orientation will actually be created<br />

in a real sample, because triple junctions with a defined<br />

combination of three a orientations (allowing for some<br />

scatter) were found to appear quite often in a real microstructure<br />

of a sample with several hundreds of thousands<br />

of grains. For a random GBP distribution, any possible<br />

product orientation according to nucleation rule I will be<br />

preferred at some triple junction.<br />

The c- and a-phase textures were measured on completely<br />

transformed samples by EBSD. The predicted and<br />

measured c-phase textures are compared in Fig. 10 in terms<br />

of the face-centered cubic (fcc) a and b fibers. The texture<br />

prediction was accomplished under the assumption that<br />

each product grain had the same volume. This assumption


1538 I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541<br />

(a)<br />

norm. frequency<br />

(b)<br />

norm. frequency<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0-30-45<br />

60-55-45<br />

Prediction<br />

Experiment<br />

2 4 6 8 10 12 14 16 18 20 22 24<br />

K-S variants<br />

Prediction<br />

Experiment<br />

2 4 6 8 10 12 14 16 18 20 22 24<br />

K-S variants<br />

Fig. 9. (a, b) Measured and predicted variant distribution of two ferrite<br />

orientations normalized to a maximum frequency of 1.<br />

was substantiated by the fact that the transformation texture<br />

did not change much with progressing transformation<br />

so that growth selection could be ruled out. For comparison,<br />

the product texture without variant selection was also<br />

calculated.<br />

3.5.2. a–c phase transformation<br />

The ferrite texture of a 50% cold-rolled and recrystallization<br />

annealed sample is shown in Fig. 10a. The c fiber<br />

is weakly developed with a maximum intensity close to<br />

the inverse brass component {1 1 2} h1 10i. The measured<br />

c-phase texture is composed of a uniformly developed b<br />

fiber with slight preference of the copper component<br />

{1 1 2} h 111i (Fig. 10b). The maximum intensity on the<br />

a fiber of austenite was found at the Goss {0 1 1} h1 00i<br />

and brass {0 1 1} h 211i components. Altogether, the austenite<br />

transformation texture was weaker than the ferrite<br />

starting texture. The predicted c-phase texture with variant<br />

selection showed excellent agreement with the experimental<br />

results. The texture prediction without variant selection led<br />

to a much weaker product texture.<br />

3.5.3. c–a phase transformation<br />

The measured c-phase texture (Fig. 10) was used as the<br />

starting texture for a computation of the back transformation<br />

texture. Again, the product texture revealed typical<br />

body-centered cubic (bcc) fibers (Fig. 11). The experimental<br />

ferrite product texture after a complete transformation<br />

cycle was relatively weak. But, good prediction of the<br />

experimental texture was also possible in this case. Because<br />

of the weak product texture, prediction without variant<br />

selection also yielded a reasonable fit. The slight deviation<br />

from experimental texture was probably due to insufficient<br />

grain statistics of the back transformed sample.<br />

4. Discussion<br />

4.1. Nucleation<br />

From previous investigations, it is known [11] that variant<br />

selection takes place already during nucleation. The<br />

product texture is therefore essentially determined by the<br />

nucleation texture. As detailed in this study, the c grains<br />

typically comprised a K–S relationship (with some scatter)<br />

to both a grains at a grain boundary next to a triple junction.<br />

The allowed deviation from the ideal K–S relationship<br />

was commonly in the range 10–15°. Furthermore, in<br />

one of the two a grains, a close-packed {1 1 0}bcc plane<br />

was parallel to the GBP. This suggests that each c grain<br />

had only one real mother grain from which nucleation<br />

started. It is proposed to associate the preference of such<br />

nucleation sites with minimization of the nucleus surface<br />

energy, which reduces the critical nucleus size and thus<br />

strongly increases its nucleation rate. An exact K–S orientation<br />

relationship is accompanied by a coherent interface<br />

between the two phases and thus a low interfacial energy<br />

[20–23]. If a grain boundary is parallel to the close-packed<br />

crystallographic planes {1 1 0}bcc//{1 1 1}fcc, it already<br />

resembles a configuration of low interfacial energy [24].<br />

Based on these facts, the following model is proposed for<br />

the observed nucleation geometry (Fig. 12). A nucleus with<br />

semi-spherical shape is considered for geometrical simplicity.<br />

Various studies report an interfacial energy of K–S<br />

phase boundaries in the range 0.27-0.56 J m 2 [25,26].<br />

The interfacial energy of a random ferrite grain boundary<br />

is 0.8 J m 2 [25].<br />

The energy DG GB<br />

het of a nucleus of a new phase is given by<br />

[27].<br />

DG ¼ V ðDGl eÞþrS ð1Þ<br />

where DGl is the chemical driving force, V is the volume of<br />

the nucleus, r is the specific interfacial energy, e is the<br />

strain energy density, and S is the interfacial area.<br />

For heterogeneous nucleation<br />

DG GB<br />

het ¼<br />

4<br />

3 pr3 ðDGl eÞþ4pr 2 r SK<br />

a c SðhÞ ð2Þ<br />

where SðhÞ is the so-called shape factor, which is defined<br />

for a spherical cap geometry by


(a)<br />

(b)<br />

f(g)<br />

{001}<br />

<br />

0 o<br />

-fibre<br />

30 o<br />

ð2 þ cos hÞ ð1 cos hÞ2<br />

SðhÞ ¼<br />

4<br />

8<br />

6<br />

4<br />

2<br />

0<br />

f(g)<br />

~{112}<br />

<br />

8<br />

6<br />

4<br />

2<br />

0<br />

(a)<br />

45 o<br />

{112}<br />

<br />

~{123}<br />

<br />

60 o<br />

{111}<br />

<br />

60 o<br />

75 o<br />

-fibre (max. density)<br />

{110}<br />

<br />

90 o<br />

~{011}<br />

<br />

90 o<br />

2<br />

f(g)<br />

{111}<br />

<br />

8<br />

6<br />

4<br />

2<br />

60 o 75 o 90 o<br />

0<br />

90 o<br />

/<br />

1<br />

55 o<br />

45 o<br />

20 o<br />

(b)<br />

-fibre<br />

ð3Þ<br />

{111}<br />

<br />

1<br />

60 o<br />

-fibre orientation<br />

Fig. 10. (a) a bcc fiber and c bcc fiber of the ferrite starting texture; (b) b fcc fiber and a fcc fiber presentation of austenite-textures: s, experimental; h, with<br />

variant selection predicted; D, without variant selection predicted.<br />

{001}<br />

<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 o<br />

{112}<br />

<br />

30 o<br />

{111}<br />

<br />

60 o<br />

{110}<br />

<br />

f(g) f(g)<br />

90 o<br />

{111}<br />

<br />

60 o 75 o 90 o<br />

0<br />

-fibre -fibre<br />

8<br />

6<br />

4<br />

2<br />

{111}<br />

<br />

Fig. 11. abcc fiber and cbcc fiber presentation of a-textures: s, experimental;<br />

h, with variant selection predicted; D, without variant selection<br />

predicted.<br />

I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541 1539<br />

1<br />

α 3<br />

75 o<br />

90 o<br />

The wetting angle h results from the surface energy<br />

balance<br />

h ¼ arccos ðra a r HC<br />

a c Þ<br />

r LC<br />

a c<br />

2<br />

f(g)<br />

{011}<br />

<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 o<br />

-fibre<br />

γ<br />

{011}<br />

<br />

30 o<br />

{011}<br />

<br />

60 o<br />

{011}<br />

<br />

90 o<br />

Less coherent phase<br />

boundary<br />

(only K-S)<br />

Better coherent phase boundary (K-<br />

S+GBP)<br />

Fig. 12. Schematic illustration of nucleation of an austenite (c) grain at a<br />

ferrite (a) grain boundary near a triple junction.<br />

1<br />

α 1<br />

α 2<br />

ð4Þ


1540 I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541<br />

ΔG [J]<br />

ΔG *<br />

ΔG<br />

TP<br />

*<br />

GB<br />

where ra a is the interfacial energy of the a grain boundary,<br />

rHC a c is the interfacial energy of the coherent a–c phase<br />

boundary (K–S + GBP), and rLC a c is the interfacial energy<br />

of the less coherent a–c phase boundary (only K–S).<br />

The dependency of DG GB<br />

het is plotted in Fig. 13. Evidently,<br />

the work of nucleation DGGB and therefore the nucleation<br />

rate strongly depend on the interfacial energy of the phase<br />

boundaries. Thus, for preferential nucleation, a low interfacial<br />

energy, i.e., high coherency, is required.<br />

While this nucleation model can account for preferred<br />

nucleation of c-phase grains at ferrite grain boundaries,<br />

experimental results demonstrate that nucleation actually<br />

starts at triple junctions. This was repeatedly confirmed<br />

by high-resolution, high-temperature in situ EBSD measurements.<br />

It is proposed that this behavior is associated<br />

with the line tension rl a of the triple junction. Nucleation<br />

at a triple junction can also use the triple line energy as a<br />

driving force, which further promotes nucleation at such<br />

locations. This modifies Eq. (2) to<br />

DG TP<br />

het ¼<br />

r *<br />

1<br />

4<br />

3 pr3 ðDGl eÞþ4pr 2 r SK<br />

a c 2r r l<br />

a<br />

The functional dependences DG TP<br />

het<br />

given in Fig. 13.<br />

Function DG TP<br />

het<br />

~r 2<br />

-r 3 +r 2 -r<br />

~r 3<br />

SðhÞ<br />

ð5Þ<br />

ðrÞ and DGGB<br />

het ðrÞ are<br />

ðrÞ reveals, besides the maximum pre-<br />

dicted by Eq. (2), also a relative minimum at a small<br />

nucleus size r 1 . Although the corresponding free energy<br />

gain is very small (10 20 J), it proves that the triple line tension<br />

favors the new phase, but of course the nucleus is too<br />

small to grow. A viable nucleus has to attain size r2 .<br />

If one assumes that rl a is independent of h the critical<br />

nucleus size follows from<br />

@DG TP<br />

het<br />

@r ¼ 4pr2 ðDGl eÞþ8pr r SK<br />

a c 2r l<br />

a ¼ 0 ð6Þ<br />

or<br />

ð8p rSK<br />

r c<br />

r1;2 ¼ Þ<br />

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

ð8p rSK r cÞ2 16p ðDGl eÞ 2rl q<br />

8p ðDGl eÞ<br />

a<br />

ð7Þ<br />

r *<br />

2<br />

~r<br />

r *<br />

3<br />

-r 3 +r 2<br />

r [nm]<br />

Fig. 13. Dependency of free energy DG on nucleus size r for nucleation at<br />

a grain boundary (DGGB) and at a triple junction (DGTP ).<br />

For an evaluation of r1=2 , the various thermodynamic<br />

quantities of Eq. (7) have to be known. According to the<br />

literature, ra a = 0.8 J m 2 , rHC a c =0.27 J m 2 , rLC a c = 0.56<br />

Jm 2 [25,26] and DGl ¼ 1:4 10 7 J=m3 [28].<br />

The only existing value for a triple line energy was measured<br />

by Zhao et al. on copper [29]. In lieu of any such data<br />

on Fe, it is assumed that the triple line energy of ferrite is of<br />

the same order of magnitude, i.e.,<br />

r l<br />

a ¼ 6 10 9 Jm 1<br />

With these values, if the elastic energy contribution is neglected,<br />

one obtains from Eq. (7) r 1 ¼ 1 nm and r 2 ¼<br />

79 nm, with the corresponding free energies DG TP ðr 1 Þ¼<br />

10 20 J and DG TP ðr 2 Þ¼2:95 10 17 J. This compares with<br />

DG GB<br />

het ðr 3 Þ¼3:15 10 17 J (r 3 ¼ 80 nm) and DGhomðr 0 Þ¼<br />

1:5 10 14 J.<br />

These results are interpreted as follows. Above the<br />

a ! c transformation temperature, the new phase spontaneously<br />

forms an embryo at the triple line, but this<br />

embryo is unable to grow. By thermal fluctuations, the critical<br />

nucleus size r 2 is obtained with a rate<br />

_N exp<br />

DG TP<br />

het ðr 2 Þ<br />

kT<br />

If nucleation of the new phase is supported by the triple<br />

line tension, the nucleation barrier DG TP<br />

hetðr2Þ is smaller than<br />

the nucleation barrier at grain boundaries. This favors nucleation<br />

of the new phase at triple junctions, while variant selection<br />

remains controlled by grain orientation and GBP.<br />

It is noted that triple junctions are also often preferred<br />

in diffusion-controlled phase transformations because they<br />

constitute high-diffusivity pathways. In the current case,<br />

this aspect is of no importance, because the composition<br />

does not change during the phase transformation in this<br />

microalloyed steel.<br />

4.2. Origin of variant selection<br />

The variant selection during the a ! c ! a phase<br />

transformation of recrystallized low-carbon steel is not very<br />

pronounced, but noticeable. As substantiated in this study, it<br />

is due to preferential nucleation of particular variants<br />

because of low interfacial energies at triple junctions, which<br />

essentially requires that the viable variant has a reasonable<br />

K–S relationship to two grains at a triple junction and that<br />

the GBP matches a {1 1 0} plane of one of the grains. The<br />

frequency of occurrence of such configurations with a high<br />

propensity for nucleation depends, of course, on the distribution<br />

of orientations and grain boundary character. For<br />

a completely random crystallographic texture, variant selection<br />

during the phase transformation is unlikely to occur,<br />

since all misorientations across grain boundaries would be<br />

equally likely and, therefore, equally favorable for all<br />

24 K–S variants. As a conclusion, the variant selection during<br />

the phase transformation in recrystallized microalloyed<br />

steel is actually a consequence of the crystallographic texture<br />

prior to transformation.<br />

ð8Þ


Nucleation rule II, i.e., a specific GBP, serves as an<br />

amplification of the texture-induced preference of a particular<br />

variant. It favors a specific among several possible<br />

variants, and thus conveys a larger weight to that particular<br />

variant in the orientation distribution. Hence, the grain<br />

boundary character distribution as well as the crystallographic<br />

texture will affect the selection of variants. In<br />

essence, both nucleation rules are important for a prediction<br />

of the transformation texture.<br />

It is finally noted that this study was conducted on<br />

recrystallized material, where variant selection is obvious,<br />

but not as conspicuous as in deformed material or during<br />

the martensitic phase transformation in steels. In the latter<br />

cases, other energetic contributions and physical principles<br />

may also play a role in the selection and preference of particular<br />

variants, as outlined in the pertinent literature<br />

[6,30–35].<br />

5. Summary<br />

Variant selection during a ! c ! a phase transformation<br />

was investigated in recrystallized low-carbon<br />

microalloyed steel. High-temperature orientation imaging<br />

by EBSD was applied to determine the location of transformation<br />

nuclei and their orientation relationship to the environment.<br />

The following results were obtained.<br />

1. The preferred selection of variants occurs during the<br />

nucleation stage of the phase transformation.<br />

2. Nucleation of the new phase occurs predominantly at<br />

triple junctions.<br />

3. The new phase was always observed to be related to the<br />

mother phase by an approximate K–S relationship.<br />

4. Among the 24 crystallographically equivalent variants,<br />

only that orientation was observed that complied with<br />

two basic requirements:<br />

a. the selected variant must have an approximate K–S<br />

relationship to two adjacent grains (nucleation rule I)<br />

b. a variant is particularly preferred over others in addition<br />

to nucleation rule I, if its {1 1 1} c plane was parallel<br />

to a {1 1 0} a grain boundary (nucleation rule II).<br />

5. The computed variant selection based on these principles<br />

showed excellent agreement with experimental<br />

results and allowed an adequate prediction of the transformation<br />

texture.<br />

6. A model is proposed to account for the observed preferences.<br />

It is based on a low surface energy of the nucleus.<br />

It was demonstrated that the triple line energy promotes<br />

nucleation of the new phase at triple junctions. The<br />

observed nucleation rules can be associated with a higher<br />

nucleation frequency of variants with low surface energy.<br />

7. It is stressed that these results pertain to the a !<br />

c ! a phase transformation of recrystallized microalloyed<br />

steel, where no compositional changes accompany<br />

the transformation. In deformed materials, complex<br />

steels or for martensitic transformations, other principles<br />

may predominate.<br />

I. Lischewski, G. Gottstein / Acta Materialia 59 (2011) 1530–1541 1541<br />

Acknowledgements<br />

The authors wish to thank Professors Lasar S. Shvindlerman<br />

and John J. Jonas for stimulating discussions.<br />

Financial support by the Deutsche Forschungsgemeinschaft<br />

(DFG) through grant Go 335/33 is gratefully<br />

acknowledged.<br />

References<br />

[1] Ray RK, Jonas JJ, Butrón-Guillén MP, Savoie J. ISIJ Int<br />

1994;34(12):927–42.<br />

[2] Kurdjumov G, Sachs G. Z Phys 1930;64:225.<br />

[3] Brückner G, Gottstein G. ISIJ Int 2001;41:468–77.<br />

[4] Nolze G. Z Metallkunde 2004;95(9):744–55.<br />

[5] Nolze G. Cryst Res Technol 2006;41(1):72–7.<br />

[6] Butron Guillen MP, da Costa Viana CS, Jonas JJ. Metall Mater<br />

Trans A 1997;28A:1755–68.<br />

[7] Wittridge NJ, Jonas JJ, Root JH. Metall Mater Trans A<br />

2001;32A:889–901.<br />

[8] Humbert M, Gardiola B, Esling C, Flemming G, Hensger KE. Acta<br />

Mater 2002;50:1741–7.<br />

[9] Bate P, Hutchinson B. Acta Mater 2000;48:3183–92.<br />

[10] Wenk H-R, Huensche I, Kestens Leo. Metall Mater Trans A<br />

2007;38(2):261–7.<br />

[11] Lischewski I, Kirch DM, Ziemons A, Gottstein G. Texture, stress,<br />

and microstructure. Cairo: Hindawi Publ. Corp.; 2008 [ID 294508].<br />

[12] Bunge HJ, Weiss W, Klein H, Wcislak L, Garbe U, Schneider JR. J<br />

Appl Crystallogr 2003;36:137–40.<br />

[13] Lischewski I. Doctoral dissertation, <strong>RWTH</strong> <strong>Aachen</strong> <strong>University</strong>; 2010.<br />

[14] Kirch DM, Ziemons A, Burlet T, Lischewski I, Molodova X,<br />

Molodov DA. Rev Sci Instrum 2008;79:043902.<br />

[15] Lischewski I, Gottstein G. Mater Sci Forum 2005;495–497:447–52.<br />

[16] Kundu S, Bhadeshia HKDH. Scripta Mater 2006;55:779–81.<br />

[17] Kundu S, Bhadeshia HKDH. Scripta Mater 2007;57:869–72.<br />

[18] Tomida T, Wakita M, Yoshida M, Imai N. Ceramic transactions. In:<br />

Proceedings of the ICOTOM, 2008, vol. 200. Pittsburgh, Pennsylvania;<br />

2008. p. 325–32.<br />

[19] Ameyama K, Maki T, Tamura I. J Jpn Inst Met 1986;50:602.<br />

[20] Furuhara T, Aaronson HI. Acta Metall Mater 1991;39:2873.<br />

[21] Furuhara T, Maki T. Mater Trans, JIM 1992;33:734.<br />

[22] Nie JF, Muddle BC, Furuhara T, Aaronson HI. Scripta Mater<br />

1998;39:637.<br />

[23] Oishi K, Furuhara T, Maki T. In: Koiwa M, Ohtsuka K, Miyazaki T,<br />

editors. Proceedings of the international conference on solid–solid<br />

phase transformations, JIM; 1999. p. 573.<br />

[24] Furuhara T, Maki T. Mater Sci Eng 2001;A312:145–54.<br />

[25] Nagano T, Enomoto M. Metall Mater Trans 2006;37A(3A):929–37.<br />

[26] Tong M, Li D, Li Y. Acta Mater 2004;52(5):1155–62.<br />

[27] Gottstein G. Physical foundations of materials science. New<br />

York: Springer; 2004. p. 400.<br />

[28] Kempen ATW, Sommer F, Mittemeijer EJ. Acta Mater<br />

2002;50:3545–55.<br />

[29] Zhao B, Verhasselt JCh, Shvindlerman LS, Gottstein G. Acta Mater<br />

2010;58:5646–53.<br />

[30] Bokros JC, Parker ER. Acta Metall 1963;11:1291–301.<br />

[31] Higo Y, Lecroisey F, Mori T. Acta Metall 1974;22:313.<br />

[32] Furubayashi E, Miyaji H, Nobuki M. Trans ISIJ 1987;27:513–9.<br />

[33] Humbert M, Wagner F, Liu WP, Esling C, Bunge HJ. In: Proceedings<br />

of the ICOTOM 8, Santa Fe, USA; 1988. p. 743.<br />

[34] Miyaji H, Furubayashi E. Texture Microstruct 1993;22:43–51.<br />

[35] Sum M, Jonas JJ. Texture Microstruct 1998;31:187–215.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!