You are on page 1of 44

Cold rolling and annealing textures i~ low carbon and extra low carbon steels

R. K. Ray, J. J. Jonas, and R. E. Hook

The cold rolling texture of low and extra low carbon steels is primarily made up of a nearly perfect normal direction (NO) fibre and of two other components, centred at {001}<110) and {112}<110), which lie along the rolling direction (RO) fibre. This texture is influenced significantly by the hot band texture and not particularly by such metallurgical parameters as steel chemistry or the presence of precipitates. The annealing of heavily cold rolled materials strengthens the NO fibre and reduces the intensity of the RD fibre. The annealing texture, particularly the strength of the {111} components, depends significantly on the hot band texture and grain size, as well as on coiling temperature, cold reduction, and alloy chemistry. The {111} fibre is particularly beneficial for imparting good deep drawability (high rm value) to sheet steel, whereas the {001} has a detrimental effect. In conventional batch annealed AI-killed steels, cold reductions of around 70%, low coiling temperatures and slow heating rates induce the development of the most desired annealing textures and correspondingly the highest rm values. The optimum cold reduction increases to about 90% in Nb or Ti stabilised interstitial free (IF) steels. Grain growth after recrystallisation increases rm; thus longer annealing cycles and higher annealing temperatures are beneficial, if grain growth occurs. In the case of box annealing, the practical upper limit of annealing temperature is '" 720C. In the continuous annealing process, which involves higher heating and cooling rates, cold rolled steels can be annealed to advantage in the intercritical y + IX range. Carbon in solution and/or in the form of carbides is the single most deleterious element that impedes the development of sharp {111} annealing textures. Dissolved N, P, and Mn have effects similar to that of carbon. Niobium, Ti, or Si in solid solution enhances the intensity of the {111} or near {111} components. IMR/266
1994 The Institute of Materials and ASM International. At the time the work was carried out Professor Jonas and Professor Ray were in the Department of Metallurgical Engineering, McGill University, Montreal, PO, Canada. Professor Ray was on sabbatical leave from the Department of Metallurgical Engineering, Indian Institute of Technology, Kanpur, India, where he has returned. Dr Hook is with Armco Research and Technology, Middletown, OH, USA.

plastic flow in the plane of the sheet, while offering sufficient resistance to flow in the thickness direction. This property of a material is called the 'normal anisotropy' and is commonly evaluated in terms of the r rn or average r-value. This is defined by the relationship, r rn = (ro + 2r 45 + rgo)j4, where the subscripts 0, 45, and 90 refer to inclinations of the longitudinal axes of tensile testpieces to the rolling direction of the sheet, and each individual r-value is the ratio of width strain/thickness strain, as measured in a simple tensile test. It has been demonstrated that high normal anisotropies or rrn values are displayed by materials which have a high proportion of grains oriented with their {Ill} planes parallel to the sheet plane, i.e. by materials which possess a strong {Ill} type texture.' Other texture components, such as the {OOl}, have been found to be detrimental to the drawability and, in practice, the intensity ratio of the above two components, I {HI}/] {OOI}, is found to be approximately linearly related to rrn (see Fig. 1). The formation of texture in steel is influenced by its alloy chemistry, as well as by the processing parameters, which include the conditions of hot rolling, cold rolling, and annealing. The effect of these variables with respect to texture formation has been studied in great detail and some excellent reviews on the subject have been published in the recent past.i" However, there have been many advances in the past few years, especially in Japan, and a large volume of literature has been published. In the present paper, the existing knowledge in this area is updated and the subject as a whole is reviewed critically. This present work provides, together with a previous review," a comprehensive treatment of the development of textures in low C steels during the entire course of processing, starting from hot rolling, through cold rolling, up to the recrystallisation stage.
-28
.::. 24
E

Published by Maney Publishing (c) IOM Communications Ltd

o
a:::
<t

t= 2

z 1-6
~ "2

til

w ~ 08 ....
~ 04

./'~'<>/
"
...L--L. L-...J--I

Introduction
The major industrial application of low and extra low carbon sheet steels has been primarily for the purpose of deep drawing. The drawability of a sheet material is its capacity to achieve a high degree of

w O'-- """ >


<t

0.1

1-0

10 INTENSITY INTENSITY

100 (111) (001)

1000

Correlation between rm and 'in steels (after Ref. 1) Materials Reviews

II"'.! flOO1}
Vol. 39

texture ratio

International

1994

No.4

129

130

Ray et al.

Textures in low and extra low carbon steels RD

x
A

(001)[110]

----'q>2

I.
/
90

rr-----_.F-!---II---r
(110)[110] (111)[121] :
I I I I

ct>

I I I I I

I I I I I I I I I

(001)[010]

xO

Ox

: ND fibre

__

(001)[110]

{111} <112>

A {554} <225>

o {111} x {112}
Published by Maney Publishing (c) IOM Communications Ltd
2

(110)[001]

<110> <110>

(200) pole figure showing some important orientations in deep drawing steels (Ref. 4)

Three dimensional view of Euler space with locations of some important ideal orientations and fibres (Bunge notation)

Textures and mechanical properties


Representation of texture Textures in rolled sheet metals are generally represented as being of the type {hkl} <uvw), which signifies that the {hkl} planes of the grains lie parallel to the plane of the sheet, whereas their <uvw) directions lie parallel to the rolling direction. More complex textures can be described as consisting of a number of components of different severities. Conventionally, texture is described by means of pole figures. Detailed descriptions of the X-ray methodsv" that lead to the determination of pole figures can be found in several texts, e.g. by Cullity,? and also in the monograph on textures in metals by Hatherly and Hutchinson," The positions of the (200) poles of some useful texture 'components in deep drawing quality low and extra low C steels are presented in pole figure form in Fig. 2. It has been recognised that, while pole figures provide a useful description of texture, the information they contain is incomplete and at best semiquantitative.? A more complete description is provided by the crystallite orientation distribution function (CODF or ODF), which specifies the frequency of occurrence of particular orientations in three dimensional (Euler) orientation space. This space is defined by three Euler angles, which constitute a set of three consecutive rotations that must be given to each crystallite to bring its crystallographic 100) axes into coincidence with the specimen axes. Mathematical methods have been developed which allow an ODF to be calculated from the numerical data obtained from several pole figures. The most widely adopted notations employed for the description of ODFs are those proposed independently by Bunge'? and by Roe.ll They use generalised spherical harmonic functions to represent crystallite distributions. A detailed mathematical treatment of this subject can be found in the texts by Bunge.12,13 While the Euler angles proposed by Bunge

to describe the crystal rotations are b <1>, nd 2, a the set of angles employed by Roe are referred to as t/!, 0, and ~, respectively. These two sets of angles are related as follows':'
~1

= n/2 - t/!; <I> 8; =

~2

= n/2 -

(1)

<

In the Bunge notation, for cubic/orthorhombic crystal/specimen symmetry, a three dimensional orientation volume may be defined by using three orthogonal axes for b <1>, nd ~2 with each of the a Euler angles ranging from 0 to 90. This volume is divided into three basic ranges in which each orientation appears once. The value of the orientation density at each point in this volume is simply the strength or intensity of that orientation in multiples of random units. The information contained in a three dimensional ODF in Euler space can be expressed in terms of: (i) 'peak' type components that are indicated by pronounced maxima in the ODF (ii) 'fibre' type components in which a more or less constant intensity is found for a group of orientations related to one another by rotations around a particular crystallographic direction. A three dimensional view of the Euler space in Bunge notation is presented in Fig. 3. In this diagram, the locations of some technologically important fibres and a few ideal orientations have been plotted. A two dimensional view of the = 45 and 2= 45 sections, in the Roe and Bunge notations, respectively, is illustrated in Fig.4a and b, in which several ideal orientations are again identified. Figure 4c and d represent the nature of the ODF contour plots derived from the same {200}, {110}, and {211} pole figure data set, when plotted in the ~ = 45 (Roe notation) and 2 = 45 (Bunge notation) sections, respectively. These two diagrams clearly illustrate the relationship between the above two types of notation, and provide a basis for the comparison of texture data using the Roe and Bunge methods. The quantitative evaluation of textures has been made possible by the availability of ODFs. Many

International

Materials Reviews

1994

Vol. 39

No.4

Ray et ale Textures in low and extra low carbon steels

131

(110]
90(001)[110]

II RD
(001)[110] (111)[112]:
I

(i11)[110]

:
I I

I I

60 C)
Q)

~
_ (111)[121]: (001)[010]
I I I

"045 -

~
30-

_ (111)[011]: (001)[010] :

EP

~
(111)[011]
-

:
I I

I
(001}[110]

G)
(111)[121] :
I

(001)[110] (111)[112]

(111}[110]:

(110)[110]

0-

I (a)
Published by Maney Publishing (c) IOM Communications Ltd

I
30

I
90

I
0

(b)

45 160 S, deq
(111)I/ND

30 --<1>,

45
deg

$1'

60

I
[110]

90

II RD

(111)IIND 90 ....-. w

0--11

--

.90 S,deg

o
4

90 (d)
<1>, deg

(c)

Two dimensional views of: a t/J=45 section (Roe notation); b tP2=45 section (Bunge notation); c ODF contour plot in the t/J=45 section (Roe notation); and dthe same ODF plot depicted in the tP2=45 section (Bunge notation)

details of individual textures, which may be ambiguous or difficult to identify in a pole figure, can now be recognised more clearly, because of the much higher resolving power of ODFs. The ODF is designated as f(g) in the Bunge convention and as w(l/J, 8, t/J) in the Roe convention.* Texture and r-value The major texture components found in the cold rolling and annealing textures of low and extra low C (ELC) steels are listed in Table 1. Calculated values of the average strain ratio r m and the planar anisotropy ~rt pertaining to each texture component are * The ODF, f(g) or w(ljJ, (J, ), gives the probability density for the occurrence of crystals within an elementary volume of Euler space. The ODF has been defined so that its integral over the Euler volume is unity. Normalisation is with respect to a 'random texture', where f(g) or w(ljJ, (J, ) = 1 everywhere in the Euler space. Therefore, the ODF for a textured material is expressed numerically in ( x R), times random, units. t The parameter ~r is defined as follows:~r = (ro + rgO-2r4S)/2. This is a measure of the tendency for ear formation during deep drawing.

also listed. These values, estimated by Daniel and


Jonas'" by using the relaxed constraint method of

crystal plasticity, indicate the contribution of each texture component to the deep drawability as well as to the tendency to form 0 and 90 ears (~r>O) and 45 ears (~r<O). Some plots of r V. the angle 8 with respect to the rolling direction, as predicted by the relaxed constraint method.Pt'" are given in Fig. 5. An examination of Table 1 and Fig. 5 leads to some simple conclusions regarding the desirable or
Table 1 Major components observed in cold rolling and annealing textures of low C steels {Ref. 15)
rm tlr

Texture component

{001}(110) {112}(110) {111}(110) {111}(112) {554}(225) {110}(001)

04 21 2'6 26 2'6 51

-08 -27 0 0 11 89

International

Materials Reviews

1994

Vol. 39

No.4

132

Ray et al.

Textures in low and extra low carbon steels


4. 0 -r-r-..,.....-,--,-~..--..--r--T'r-rl--,-, {001}<110> 1.5
(l)

3.5
3.0
(l)

:J

2.5

:J

~ 1.0

J..

0.5

o I

o00'0 '0

cti 2.0 > J.. 1.5


1.0
0
\

\ 0

0.5
O.

o .0

.0 ~~+-+-f--t--'t-t-I'-t-i-t-t-t-'f-T-'f

0.0 +++++-+-+-+-+-+--t-+-+-t-+-t-t-t-1

30

60

90

30

60

90

8,deg
{lll}<llO> 3.5
3.0

8,deg
4.0 ....---r-...-.-..-.-..--.-.....-..--...,-.,....--r-lr-r-l

{111}<112>
'0'0 '0, 0'0'00.0 0"00'0'0 ,0"

~ 2.5
~ 2.0 J.. 1.5

0,
0'0.

Published by Maney Publishing (c) IOM Communications Ltd

30

60

30

60

8,deg
4.0 ......-r-r--r-r--r-r-r-r--r-r--.....-r-1.......,-,

8,deg
{110}<OOl>

3.5

{554}<225>

cti

s 2.5
!2.0
0.5 0.0 +-t-+-t--t-+-+-+-+-+-+-+-l-f-I--HI-H 1.5 1.0
0 000000000"

3.0

I 1
60 as predicted by relaxed

30

60

90

30

8,deg
5

8,deg

Plots of r v. angle 0 with respect to rolling direction for selected ideal orientations constraint method (after Ref. 16)

undesirable natures of individual ideal orientations. For good deep drawability (characterised by high rm and minimum l~rD, the components {111}(110) and {111}(112) are clearly beneficial, with the {554}(225) component coming in as a close third preference. This is why a strong and homogeneous {Ill} fibre texture is sought in low C steels designed for optimum deep drawability. Because the presence of the {111}(110) component promotes the appearance of sixfold ears of the 30, 90, 150,... type (with 'valleys' at 0, 60, 120,... ), whereas that of the {111}(112) component is responsible for the formation of 0, 60,120, ... ears (with 'valleys' at 30, 90, 150,... ), equal intensities of these two components (i.e. a 'balanced' {Ill} fibre) will lead to the relative absence of ears. The anisotropy of the yield strength, which is also related to deep drawability, is discussed below, together with that of the elastic modulus of textured steel sheet, which can be of importance in certain applications.
Texture and yield strength

direction is expected to vary for different texture components. These calculations were carried out by assuming that each texture component is the only one present (with no random background), and that there is a 15 gaussian spread about the respective ideal orientation.P'!" The yield strength varies somewhat with 8 for the components {111}(110), {111}(112), {554}(225), and {001}(110). By contrast, the 0'(8)/0'(0) ratio varies considerably more with 8 for the {110}(001) and {112}(110) components.
Texture and elastic modulus

The maximum value of the Young's modulus E of iron and steel is obtained along the (111) direction and the minimum along the (100). The value of E along a certain direction, whose direction cosines with respect to the crystal axes are x, y, and z, is given by the relationship 1 -=
E

1 ---3 (1 ----- 1) E100 E100 Ell!

(X2y2

+ y2z2 + Z2X2)
(2)

Figure 6 illustrates how the yield strength a measured along a direction inclined at an angle 8 to the rolling
International Materials Reviews 1994 Vol. 39 No.4

Generally speaking, the directional dependence of the

Ray et a/.

Textures in low and extra low carbon steels

133

1.25

~ {OOt}<II0>

1.25

: {112} <11 0>

0'

1.15 1.05 0.95 0.85 0.75 0.65

o
0000'0. 00000 000000

p'

1.15
0'

oP
0000 00 .0' 000000

~ CD b

CD b

J2

1.05 0.95 0.85 0.75

111111111

0.65

30

60

90
1.35

30

60

90

8,deg
1.35 1.25

8,deg
{111}<112>

{lll}<llO>

1.25

1.15 1.05 0.95 0.85 0.75 0.65


0000'0 0-0-0' '0'00000-0' 0.0

e
b

1.15 1.05
0'0 '0-00000-0' 00-0-0'0 '0-00

b Published by Maney Publishing (c) IOM Communications Ltd

"-" b 0.95

0.85 0.75 0.65

0
1.35 1.25

30

60

90
1.35

30

60

90

8,deg
{554}<225>

8,deg
{110}<OOl> 00'0'0

1.25

d
" ,0,0 ,0 0.0 0-00'

\
0

CD b

0' "-" J2

1.15 1.05 0.95 0.85 0.75 0.65


.00'0 0.0 '0. 0'000_00-0.0' 0'0-0

0'

1.15 1.05 0.95 0.85 0.75

'0,
0'0 -

b
~

0.65 +-+-+-+-+-+-+-+-+-+-+-+-+-+-++-+-+-1

30

60

90

30

60

90

8,deg
6 Relaxed constraint predictions of 0'(8)/0'(0)

8,deg
for selected texture components (after Ref. 16)

elastic moduli of polycrystalline metals can be predicted from that of the single crystal moduli. In practice, it is not easy to estimate, since the effect of grain boundaries as well as of the interaction between neighbouring grains must be taken into account. Three models have been proposed for the calculation of elastic properties in textured polycrystalline aggregates. These are: (a) the Voigt model, which assumes an identical strain state in all the crystallites; (b) the Reuss model, where the stress in each grain is supposed to be that applied to the bulk specimen; and (c) the Hill approximation.!? which employs an arithmetic average of the above upper (Voigt) and lower (Reuss) limits. For deep drawing quality metal sheets, the Hill formulation leads to the most accurate description of the elastic anisotropy.!? The values of the single crystal elastic constants that are selected have a significant effect on these predictions. Calculated values of Young's modulus.l'v'" using the Hill approximation, are plotted against 8, the angle to the rolling direction, for several important texture components in cold rolled and annealed low C steels (Fig. 7). The variation of E with angle 8 is least for the orientations {111}(110) and {111}(112), followed by {554}(225).

Characterisation of textures in low carbon sheet steels


Cold rolling and recrystallisation textures in ferrite based steels have commonly been described in terms of certain orientation fibres in Euler space. It is customary to plot the orientation density along these fibres and to use their relative intensities to distinguish the hot band, cold rolling, or recrystallisation textures of one steel from another. These fibres have been referred to as the a, y, 1'/, and 8.20,21 However, the terms a and y for the fibres can lead to confusion because of the a and y phases in steel. In order to avoid these difficulties, they will be referred to instead as the RD (rolling direction) and ND (normal direction) fibres, respectively, as defined in more detail below. In earlier publications, the existence of another fibre known as the {3was also reported; this was essentially on the basis of the theoretical prediction of cold rolling textures." The locations of all these fibres in Euler space are shown in Fig. 3 (except for the {3-fibre)and their specific orientation ranges are listed below: (i) RD or a-fibre running from {OOl}(110) to {111}(110) along (110)IIRD
International Materials Reviews 1994 Vol. 39 No.4

134

Ray et al.

Textures in low and extra low carbon steels

- {OO1}<t 10>
N

2.5

C) 2.0
N

'E

b orui '.5 ui 1.5

30

60

90

30

I "

I "

60

I I

e,deg
{111}<110>
~ 2.5 ~

e,deg
{111}<112>

:f
Published by Maney Publishing (c) IOM Communications Ltd

E
2.0

2.5

:f

b or-

b or-

2.0

w",.5

ui 1.5
1.0

1.0

0 3.0

30

60

90 3.0

30

60

90

e,deg
{554}<225>
N

e,deg
{110}<OOl>
N

'E

2.5
0_00-000-00-000-00_0-0-0-0

:f

z
2.0

b or-

:f

2.5

p
2.0
O/ 00' -0-

o0-0-0. 0, 0,
0'0'0_

b 0r-

dO

0 I

w",.5 , .0

ui' 1.5

+-+-t-t-+-t--l--+-+-+-+-+-+-+-+-~~

30

60

90

30

60

e,deg
7

e,deg

Plots of Young's modulus E v. angle (J with respect to rolling direction for selected texture components (after Ref. 16)

(ii) ND or y-fibre running from {111}(110) to {111}(112) along (111) liND (iii) RD' or 17-fibrerunning from {001}(100) to {011}(100) along (100)IIRD (iv) TD (transverse direction) or s-fibre running from {001}(110) to {111}(112) along (110)IITD (v) ND-RD or j3-fibre running from {112}(110) to {11 11 8}(4 4 11) along a (110) fibre axis inclined at "30 to ND in the ND-RD plane. Commercially produced sheet steels can be made with controlled compositions and processed to produce controlled crystallographic textures. The texture can be regarded as a controlled variable. The main types of commercial sheet steels, which are distinctive with respect to texture, can be classified based on steel composition as follows: 1. Rimmed steel, no longer widely produced because of the conversion from ingot pouring to strand casting. 2. Aluminium killed steel. 3. Aluminium killed steel, ELC (vacuum degassed/ vacuum decarburised to < 00100/0C).
0

4. 'Ordinary' Ti stabilised interstitial free (IF) steel, (typically ",0006-00100/oC (max.), ",0'07-0'12%Ti, i.e. highly alloyed). 5. 'Ordinary' Nb stabilised IF steel, (typically ",0'006-0'0100/0C (max.), ",0'08-0'12%Nb, i.e. highly alloyed). 6. Titanium stabilised IF steel, ELC type (typically ",0003AlC (max.), ",0'0020/0N, ",0,05-0,07% Ti, i.e. lightly alloyed). 7. Niobium stabilised IF steel, ELC type (",0030/0Nb), i.e. lightly alloyed). 8. Nb + Ti stabilised IF steel, ELC type (typically '" 0003%C (max.), '" 0002%N, '" 0'0100AlNb, '" 003%Ti, i.e. lightly alloyed). Details of the characteristics of the textures developed in these steels based on this classification follow.
Historical development of interstitial free steels

Ordinary (or conventional) 'interstitial free' (IF) steels* were developed commercially in the late 1960s * Technically, the term 'I-F Steel' should not be used, since it is a registered trade mark issued to the then Armco Steel Corp.; Armco was the first US steelmaker to make this type of steel starting in 1970.

International

Materials Reviews

1994

Vol. 39

No.4

Ray et a/.

Textures in low and extra low carbon steels

135

and early 1970s following the introduction of vacuum degassing technology in the steel industry. Since the current routine capability of achieving very low C and N contents (C = 30-50 ppm max., N = 20-30 ppm max.) was not realised at their inception, these original IF steels were relatively highly alloyed by present standards. When these IF steels were developed, the existing steelmaking technology resulted in higher interstitial element levels (C = 50-100 ppm, N = 40-80 ppm). The IF steels produced commercially were Al-killed (0'02-0'070/0AI) and alloyed with either Ti (generally 0'07-0'12%) or Nb (generally 0,080'120/0) or with binary additions of Nb + Ti (typically 0'05 Nb, 005% Ti). In these steels, if only Ti was added, the Ti scavenged both the C and N; if only Nb was added, the Nb scavenged only the C while the N was combined as AIN; and if both Ti and Nb were added, the Ti scavenged both the C and N (if the Ti/(C + N) ratio was <4: 1, the Nb would combine with the remaining C). It is likely that the first type ofTi stabilised ordinary IF steel was developed by Shimizu et al.22 of the Yawata Iron and Steel Co. Ltd (known today as the Nippon Steel Corp.). The introduction of this grade was followed by that of the Nb (Ref. 23) and (Nb + Tif4 stabilised types of IF steel by the Armco Steel Corp. (now AK Steel Corp.). A distinguishing feature of these ordinary IF steels was that they contained a considerable excess of the stabilising elements, which did not combine with either C or N. It is precisely this difference, compared with the more modern ELC (or ULC) IF steels, that leads to measurable differences in texture development. The latter contain only minor amounts of uncombined solute alloying elements. Extra low carbon (ELC) or ultra low carbon (ULC) IF steels came about as a result of the introduction to steelmaking technology of the bottom blown converter and a reformed RH vacuum degasser, which enabled significantly lower C and N levels (C = 30-50 ppm max., N =20-30 ppm max.) to be consistently attained. The attainment of lower interstitial levels permitted the use of substantially reduced solute additions (Nb and/or Ti) to achieve the interstitial free state. These events transpired in the 1980s. A principal driving force in this development was the widespread implementation of continuous annealing lines in Japan. The Kawasaki Steel Corp., in particular, played the major role in developing this type of IF steel. The article by Obara et al.25 dealing with ELC steels provides further details of the developments which occurred. While the ELC IF steels provide rm values which are relatively insensitive to coiling temperature and annealing method (continuous v. batch annealing), they require higher cold reductions to attain the levels of r m associated with ordinary IF steels. Strictly speaking, the ELC IF steels have somewhat lower rm values after continuous as opposed to batch annealing. The planar anisotropy is also different for the two annealing methods as a result of small, but significant, differences in the textures produced. The planar anisotropy of cold rolled (CR), batch annealed (BA) ordinary Nb stabilised IF steels (the first type introduced commercially) differs from that of the ELC
%

Nb stabilised IF steels produced today. This is partly because hot bands of the latter grade contain a less intense {112}(110) texture component than found in the hot bands of ordinary Nb stabilised IF steels. These characteristics of the various types of IF steel are described in detail in the sections that follow, where references are provided for each of the traits cited. As will be seen in more detail below, the relatively high level of Nb in the ordinary IF steels is responsible for more austenite pancaking, and therefore for the increased intensity of the rolling texture components in the austenite before transformation. The latter are in turn responsible for the more intense {112}(110) (and {332}(113) components in the ferrite after transformation. To compensate for the reduced Nb level and {112}(110) component intensity in the ELC IF steels, additional cold reductions are required. Conflicting reports plague the subject of textures in sheet steels and frequently make it difficult, and sometimes impossible, to draw clear cut conclusions. In many cases, discrepancies result from differences in the procedures used to obtain the results. The key word here is comparability. For this reason, the characterisation of the textures of the various types of steel in this section is based as far as possible on studies that employed comparable procedures. Table 2 lists the different grades of low and extra low carbon steels, the textures of which have been characterised. The compositions of the steels, their processing histories, and the types of texture sample used, sheet or composite.Pr" are also indicated in the table. The steels were commercially melted and cast and mill hot rolled (HR); an exception applies to the ELC IF steels, which were mill cast but laboratory processed thereafter. Cold rolling and annealing were conducted either in the mill or in the laboratory, as indicated in Table 2. When significant through thickness texture gradients were present." composite samples were employed to provide the average through thickness texture. Some of the results are taken from published articles; of necessity, however, the use of the comparative method has required the incorporation of some previously unpublished results as well. The development of cold rolling and annealing textures in low carbon and extra low carbon steels is affected by each aspect of their processing history. The operations of importance thus consist of hot rolling, cold rolling, and annealing. The characteristics of the textures formed during these three different stages of processing, and the factors that affect them, are discussed below for the different steels listed above.
Hot band textures in low carbon sheet steels

Published by Maney Publishing (c) IOM Communications Ltd

These steels are all ferritic, but are hot rolled in such a manner that rolling is completed while they are still in the austenite phase. An exception applies to some special ELC IF steels for which the final rolling reductions are intentionally carried out in the ferrite range (warm rolling). Results for this case are not presented here. The hot band ferrite texture is formed by transformation from the austenite and is not a deformation texture resulting directly from rolling
International Materials Reviews 1994 Vol. 39 No.4

136

Ray et al.

Textures in low and extra low carbon steels


RD .. --_ (001)[010]
CUBE (8)'\i

90 .. --.I.....,...~-'""'80 70
60 ~50
"'C

...._I.--&o..-.I_ ......

(001)[110]
GOSS (8)

(001)[Ho]

BRASS (4)/f

COPPER (8)

S (5)

(112)[110]

GOSS

(111)[1To]

GOSS

" ---------@- -------r(;l- ------~ " t- t(8) (112)[131]

-c--------~BRASS(~

0.5

BRASS

(8) (111)[H2]

~40 30 20 10

(8)

(111)[121]

(111)[011]
GaSS (8)

BRASS (8)

L
2 ...,....-r--"r----r'~ ...r,.-+-....,.- ....

(110)[110]

O~

Published by Maney Publishing (c) IOM Communications Ltd

Selected bcc product orientations in tP2 = 45 section (Bunge notation), showing fcc parent orientations from which they originate, (after Ref. 37)

o
9

10 20

30

40

SO 60

70 80

90

8,deg

tP = 45 section (Roe notation) of ODF of hot rolled AK steel (L-K composite sample) (after Ref. 28)

(except for the special case of ELC IF steels hot rolled in the ferrite range). Effect of y-to-ex transformation on ferrite textures During hot rolling, the parent austenite phase develops a crystallographic texture, which is later inherited by the ferrite on transformation. At relatively high temperatures of rolling, (i.e. above the Tnr or austenite no-recrystallisation temperature), a weak recrystallisation texture is formed. The austenite in this case displays the cube {001}(100) texture, which transforms primarily into the rotated cube {OOl}(110) in the ferrite." If, on the other hand, the austenite is not able to recrystallise during or after rolling, it develops a sharp texture containing the brass {110}(112), copper {112}(111), and S {123} 634) components, together with a weaker Goss {110} (001). During the austenite-to-ferrite transformation, the Kurdjumov-Sachs (KS) orientation relationship is generally followed.35,36 According to this model, one parent orientation transforms into 24 product orient-

ations or variants. Figure 8 represents the ~2 = 45 section (Bunge notation) in which selected bee product orientations, calculated according to the KS relationship, are identified, together with the fcc parent orientations from which they have formed.'? The number of variants corresponding to each product orientation is indicated within brackets after the name of the respective parent orientation. Hot rolled rimmed and drawing quality AI-killed steels (DQAK) Rimmed and AK steels have virtually identical and very weak, nearly random HR textures, with the {001}(110) (transformed austenite recrystallisation) component being approximately 2 times random (2 x R).26 The ODF of Fig. 9 for a mill produced AK HR steel'" shows the texture, which differs only negligibly from that of a HR rimmed steel." The recrystallisation of austenite during hot rolling is unimpeded in these two types of steel and is sufficiently rapid for recrystallisation to be essentially complete before the transformation to ferrite. This
textures
Annealing Type of PF %CR Temp., DC Time, h samples Ref.

<

Table 2

Steels used for characterisation


Composition, wt-% AI N

of crystallographic

Steel type Plain low carbon AK-HR (O'018%S) AK-CRA (0'018%S) AK-CRA (O'008%S) (low Mn, low S) Ordinary IF steels 0'24%Nb IF 0'30%Ti IF 0'095%Nb IF 0'13%Ti IF ELC IF steels 0'023%Nb IF (ELC) 0'037%Nb IF (ELC)

Mn

Process*

HR CT,t DC

0042 0042 0041

031 031 020

0'028 0028 0032

00093 0'0093 0'0060

A B
C

593 593 566

60 65

738 721

10 20

L-K L-K S(1/4T), RD

26 26 27

00063 0'0048 0'0033 00031 0'0028 00027

040 030 0'40 022 0'18 016

0044 0096 0023 0'040 0037 0059

0'0056 00044 0'0063 0'0056 0'0017 0'0038

C C D C

649 704 649 621 704 704

62 60 70 66 75 75

746 738 732 738 732 732

20 16 4 16 4 4

L-K L-K S(1/4T) S(1/4T), RD S(1/4T) S(1/4T)

28 27 29,30 27 31 31

E E

*A mill HR B mill HR, CR, and laboratory

SA C mill HR, CR, SA D mill HR and laboratory CR, SA E mill strand cast, laboratory HR, CR, SA.

tCT coiling temperature. PF pole figure L-K Lupata-Kula type composite sample (Refs. 32, 33) RD RD composite sample (Ref. 30) S(1/4T) sheet sample, 1/4 thickness position.

International

Materials Reviews

1994

Vol. 39

No.4

Ray et ale 90-1- ......


80 70 60
"'Cl

Textures

in low and extra low carbon steels

137

~ 50

~40 30 20 10

O""'-;-;r-rTTTT-rl~r+-H-r-1~~~!"-'..,..,~.,J..L.J..!..~~~
30 40 50 60 70 80 90 10 20 30 40 50 60 70 80 90

e,deg Published by Maney Publishing (c) IOM Communications Ltd


10 i/J = 45 sections (Roe notation) of ODFs of hot rolled a 024%Nb IF steel (L-K composite samples) (after Refs. 27,28) ordinary IF steel and b 0300/0 ordinary Ti

interpretation is supported by the relative absence of other texture components in the ferrite, indicating that the austenite did not contain any rolling components before transformation."
Hot rolled ordinary interstitial free (IF) Nb and Ti stabilised steels

The ODFs of two highly alloyed mill HR IF steels are presented in Fig. lOa and b, one a 0'240/0Nb and the other a 030%Ti steeI.27,28,38 Compared with the HR rimmed and AK steels, the textures here are quite strong. The 0'24%Nb stabilised steel can be characterised as displaying a (110)IIRD partial fibre, with a strong peak near {112}(110) of 4 times random (4 x R). Weaker components of interest are the {001}(110) at 24 x R and the {554}(225) or, more accurately, {332}(113) at 22 x R. The {Ill} fibre texture ({lll}IIND) is not found. The 0300/0 stabilised steel also displays a strong Ti HR texture, but one which differs somewhat from that of the Nb steel. It is again characterised by a (110)IIRD partial fibre, but with the strongest component (6'3 x R) at {OOl}(110), while the {112}(110) IS weaker at 36 x R in this case. A further component of interest is the {332}(113) at 22 x R. Thus, the {001}(110) is more intense than the {lI2}(110) in the Ti steel, whereas the inverse relationship applies to the Nb steel. The results of Schlippenbach et al. 20 on the HR texture of a 022%Ti steel, measured at the midthickness of the sample, showed f(g) values equal to 33 and 47 at {112}(110) and {001}(110), respectively; this again shows that the latter orientation is favoured in the Ti steel. Conversely, in experiments performed on a 0'0950/0Nb steel, where the texture was again measured in the midthickness of the sample, reported f(g) values ?f 76 and 52 at {112}(110) and {OOl}(110), respectively, are in agreement with the observed trends in Nb steels, as described above.29,30 The strength of the {112}(110) component of the HR texture of Nb IF steel does not decrease much with reduction of the solute Nb content. For example, when the solute Nb content is reduced from 019 to

0'03%, the strength of the {112}(110) only declines from 38 x R to 25 x R (Ref. 28). The above results indicate that the dominant component of the HR texture is {112}(110) in the Nb steels and {OOl}(110) in the Ti steels. From Fig. 8, it can be seen that the {112}(110) ferrite component is derived principally from the Cu {112}(111) component of the unrecrystallised austenite, with a possible contribution from the weaker, Goss {110}(001) component. By contrast, the {OOl}(110) ferrite component is obtained partly from the cube {OOl}(100) component of the recrystallised volume fraction of the austenite, and partly from the brass {110}(112) component of the unrecrystallised volume fraction. Th~s it appears from the intensities that, during hot rolhng, the austenite in the Nb steel is essentially pancaked (unrecrystallised), while the austenite is partially recrystallised in the Ti steel. The retardation of austenite recrystallisation in the Nb steel during hot rolling is attributable to two complementary factors: (a) the presence of solute Nb in the austenite, and (b) the precipitation of Nb carbonitrides in the matrix. Hook and Ny038 have, for example, shown that ferrite recrystallisation is retarded in Nb IF steels by both Nb in solid solution and by the presence of fine dispersions of NbC; Nb in solid solution having the greater effect. This view is supported by the low interstitial content of these steels. By contrast, Ti in solution in the austenite does not appear to be as effective as Nb in retarding austenite recrystallisation. Furthermore, Ti carbonitrides are less likely to form than Nb carbonitrides at hot rolling temperatures in the austenitic range, and are therefore less effective in preventing austenite recrystallisation.
Hot rolled extra low carbon Nb and Ti stabilised steels

Examples of hot rolled textures evaluated by ODF analysis for ELC IF steels falling in the composition ranges defined previously were difficult to find. The only HR texture results reported for steels finish rolled in the austenite range are in the form of pole
International Materials Reviews 1994 Vol. 39 No.4

138

Ray et al.

Textures in low and extra low carbon steels


0r""'T"-r-""I""'"""r-----,---r---_~!"""I

figures. Hutchinson et a1.39 published a (200) pole figure for a HR steel containing 0'004%C, 0'04%AI, O'003%N, 0'028%Nb, and 00330/0Ti.They described the texture as weak and concluded that the alloy content was insufficient to prevent recrystallisation between passes. Kwon etal/" presented a (110) pole figure for a HR steel containing O'00350/0C, '049%AI, 0 0'00390/0N, and 00770/0 Ti. They interpreted the texture as weak and considered that it did not result from transformation of a substantially unrecrystallised austenite. Figure 11 shows the ~2 =45 section (Bunge notation) of the ODF from a hot rolled ELC IF steel containing 00018% C, 014% Mn, 0004%P 0003% S 0'042%AI, 0'00230/0N, 0'010%Nb, and 0'0790/0Ti, (Ref. 41). The sample employed for texture measurement was prepared from the midthickness of the hot band. The hot band texture of this ELC Nb + Ti steel is qualitatively similar to that of the AK steel L-K composite sample (see Fig. 9). A quantitative comparison of Fig. 11 with Fig. 9 is of doubtful validity because of the different sample types used to evaluate the textures. Both textures are characterised by {001}(110) as the major component. The hot rolling texture of the ELC Nb + Ti steel can also be seen to resemble that of the ordinary IF steel containing Ti (Fig. lOb, L-K composite sample), which was obtained by transformation from the texture of a partly recrystallised and partly pancaked parent phase. Samuel et al.42 studied the recrystallisation behaviour of a Ti-containing ELC IF steel under simulated strip rolling conditions. The steel composition was 0004%C, 0'2%Mn, 0008%S, 0'0040/0N, and 00660/0 They observed that very fine Ti(C,N) and Ti. Ti4C2S2 particles are precipitated during hot rolling. The relatively high temperatures of roughing, combined with the small precipitate volume fraction (because of the low C and N concentrations) rendered these precipitate particles ineffective for the retardation of recrystallisation during the early stages of rolling. In the finishing stages, there is little effect on the recrystallisation kinetics because of the short interpass times involved. When finishing temperatures just above the Ar3 were employed, some strain accumulation took place during the later stages of finishing, which led to the initiation of dynamic recrystallisation. These results indicate that the austenite of hot rolled E;LC IF steels is essentially in a recrystallised state before transformation, which explains why the ferrite transformation texture is weak. As discussed above, the cube texture in the recrystallised austenite transforms primarily into the rotated cube in the ferrite (Fig. 8). Similar behaviour is expected from Nb-containing ELC IF steels. Najafi-Zadeh et al.43 investigated the influence of hot strip rolling parameters on austenite recrystallisation in three ELC IF steels with ",000350/0C and containing 0'056%Nb, 0'0650/0Ti, or 0'0280/0Nb+ 0035%Ti. They found that the increase in flow stress in the finishing passes (due to retained work hardening) is more marked in the Nb than in the Ti or the Nb + Ti steel. This indicates that the Nb in solution was more effective in retarding recrystallisation than the Ti. (Here the interpass times were too short
0

30

90 0

30

60 <i>l,deg

90

11

tP2 = 45

Published by Maney Publishing (c) IOM Communications Ltd

section (Bunge notation) of ODF from 001 %Nb + 0080/0 hot rolled ELC IF steel (sheet Ti sample) (after Ref. 41)

to permit significant precipitation). Such solute retardation is responsible for the higher intensities of the transformed rolling texture components (i.e. the transformed Cu, brass, and S conlponents, see Fig. 8) in the hot band textures of Nb as compared with Ti or Nb + Ti steels.
Cold rolled textures in low carbon sheet steels Rimmed and drawing quality AI-killed steels

Cold rolled textures are virtually identical for rimmed and AK steels when they are evaluated on a comparable basis, as shown by Heckler and Granzow." It has been noted that the HR textures of the two types of steels are similar, displaying only a weak {001}(110) component. With increasing cold reduction, these low carbon steels develop both a partial (110)/IRD fibre and a {111}IIND fibre. The transformed austenite recrystallisation component, {001}(110), also sharpens perceptibly. The strongest component of the texture is displaced from {111}(110) at 60% cold reduction towards {112}(110) at 80%. All these features can be seen in the ODF plots (Fig. 12a and b) of a rimmed steel, cold rolled 60 and 800/0(Ref. 26).
Cold rolled ordinary Nb and Ti stabilised IF steels

Ordinary Nb stabilised IF steels have CR textures characterised by a sharp (110)IIRD partial fibre with maximum intensities at the {112}(110) orientation." When the cold reduction is held constant at 60%, the strength of the {112}(110) (transformed Cu) orientation increases with the concentration of Nb in solid solution; this reflects the increase in the strength of this orientation in the HR state attributable to austenite pancaking. The ratio of the strength of {112}(110) to {554}(225) (transformed brass) in the CR state is about 2: 1 (Ref. 28). The {001}(110) component is approximately as strong as the {554}(225) component. As shown above, the {112}(110) orientation is the strongest component in HR Nb stabilised IF steels because of austenite pancaking (Fig. lOa), while in

International

Materials Reviews

1994

Vol. 39

No.4

Ray et ale
90 -t-...... ""'80 70 60
"t:S

Textures in low and extra low carbon steels

139

-.I_~~-+-~----!~-t--~ .. _~~-- ............. ~ -1~ -

~3 ~2

if

50

~40 30 20 10

(b) O ......... -...-..-..- _ ... --l ~ ;",,;,;. - ....-

o
Published by Maney Publishing (c) IOM Communications Ltd
12

10

20

30 40

70 80

90

10

20

30 40

50

60 70

80 90

8,deg
l/J = 45 sections (Roe notation) samples) (after Ref. 26) of ODFs of rimmed steel, cold rolled a 600/0and b 80% (L-K composite

HR ordinary Ti stabilised IF steels, the strongest orientation is {001} (110) as a result of austenite recrystallisation (Fig. lOb). In a similar manner, the texture of CR ordinary Ti stabilised IF steels differs quantitatively from that of the equivalent Nb steels. The former can be characterised as having major orientations whose strengths decrease in the order {112}<110), {554}(225), and {001}(110). However, the intensities of these components are appreciably lower than in the comparable Nb steels. ODFs are presented in Fig.13a and b for the highly alloyed 024%Nb and 030% Ti cold rolled IF steels discussed above. The listing of key texture components in Table 3 shows how the strengths of the {112}(110) (transformed Cu) orientations increase with the concentration of Nb in solution, which varied from 0002%, to 003%, and to 019% in the order given. While cold rolling results in significant strengthening of the {Ill} (110) in rimmed and AK steels, it leads to greater enhancement of the {112}(110) in
90 80 70 60
"t:S

Nb and Ti stabilised IF steels at comparable cold reductions of 60%. These results are explicable in terms of the relative strengths of the {112}(110) component in the hot bands of these four types of steel, as shown in the section 'Effect of hot rolled grain size on cold rolled and cold rolled and annealed textures' below. Cold rolled extra low carbon Nb and Ti stabilised steels Figure 14a and b shows the ~2 =45 sections (Bunge notation) of the ODFs of an ELC IF steel containing both Nb and Ti, after cold reductions of 70 and 850/0 (Ref. 44). The HR texture of this steel has been described above (see Fig. 11). Cold rolling produces sharp RDI/(110) and NDII(lll) fibres in this steel, the intensities of which increase with the amount of cold reduction. The main component, {001}(110), of the hot rolled texture also sharpens significantly with increasing cold reduction (see Fig. 14). Furthermore,
0

~.-.1!~~~~~~~""""''''~''IIIIIIII!II'''''''''''''

....,.III!IIII

_'"'"

if

50

~40 30 20 10

o ....... . .......... ~ ..,. ~~~_ o 10 .20 30 40 50 60 70 80 90 10 20 30 40 50 60


(a)
13 .t/J = 45 sections

.....
70 80 90

8, deg
ordinary IF steels containing

(b)
a 0240/0Nb and

(Roe notation) of ODFs of cold rolled b 030%Ti, (L-K composite samples) (after Refs. 27,28)

International

Materials Reviews

1994

Vol. 39

No.4

140

Ray et al.

Textures in low and extra low carbon steels

o
~

{001}<110> {111}<110>

{112}<110> {111}<123>

0 {223}<110>
{111}<112>

30

tn
Q)
"'C

e
60

Published by Maney Publishing (c) IOM Communications Ltd

11 (a)

10
(b)

90

90
30 CP" deg

60

90

30 CP" deg

60

14

tP2 = 45 sections (Bunge notation) of ODFs of ELC IF steel of Fig. 11, after cold reductions b 850/0, (sheet samples) (after Ref. 44)

of a 700/0 and

the strongest component of the cold rolling texture, which is located at {112}(110) after 700/0 cold reduction, undergoes a perceptible shift to the {223}(110) position after 850/0rolling. The development of cold rolling textures in low and extra low carbon steels has recently been dealt with in some detail by T6th et al.45 They predicted the stability of the main rolling texture components analytically, using a rate dependent theory for mixed {112} 111> and {110} 111> slip. Both full constraint (Taylor) and relaxed constraint (lath and pancake) grain interaction models46-49 were employed for this purpose. An orientation stability map, derived by them and plotted in the ~2 = 45 section (Bunge notation), is presented in Fig. 15. In this figure, the numbers associated with the contour lines refer to an orientation stability parameter S. This displays high values in regions of Euler space where the orientations are relatively stable and low values (these can even be negative) where the degree of stability is low. A study of the above diagram indicates that the principal component inherited from the recrystallised hot band, i.e. {001}(110), is a fairly stable orientation, and that is why its intensity increases during subsequent cold rolling. Of the two major texture components observed in hot bands processed in the y no recrystallisa tion range, the {113} 112} 110> was
I"V

<

<

found to be highly stable during further deformation, whereas the {332}(113) component shifts towards {554}(225) and then to {111}(112) during subsequent rolling. At large strains, there is a net rotation from the {111}(112) to the {111}(110) position. The orientations possessing the highest stability are the {112}(110) and {445}(110), followed closelyby the {223}(110). The experimental cold rolling textures of low and extra low carbon steels, described above, agree remarkably well with the above predictions. In the case of the ELC IF steels, however, the {223}(110) appears to be slightly more stable than the {112}(110) at larger strains (see Fig. 14).
Annealing textures in low carbon sheet steels

Valid comparisons of the cold rolled and annealed (CRA) textures of various steels should be made for the same amount of reduction and annealing method (batch v. continuous). For some steels, heating rate also has an important effect on the nature of the recrystallisation texture developed.
Annealed rimmed steel and drawing quality AI-killed steels

I"V

<

The ODFs of the batch annealed rimmed and AK steels published by Heckler and Granzow" (see Table 2), indicate the following trends. During the

Table 3
Steel 0'073%Nb 0092%Nb 0'24%Nb 030%Ti

(J)

(IjI, 8, tP) intensities Condition


CR CR CR CR

of components
{112}(110) 72 78 90 50

( x Random) (Refs. 27, 28)


{001}(110) 41 41 41 39 {5541(225) 46 44 42 41 {111}(112) 46 44 41 38 {111}(110) 40 41 43 36

International

Materials Reviews

1994

Vol. 39

No.4

Ray et al.
(001)[110] ~r-----:=--~~"T"---========t(001 )[110]

Textures in low and extra low carbon steels

(114)[110] (113)[110] (113)[332] -1 (112)['111]

eo

(112)[1~O] (223)[110] (445)[110] 3 (111)[011] 0 (111)[110] 4 --

--------"i.:....-------5~~:.:..:...::..:...:..:~(111)[TI2] (111)[121] 1 (554)[225]

(332)['113]

with the implementation of somewhat higher cold reductions, resulted in stronger CRBA textures in the AK steels, as illustrated in Fig. 16b; this led in turn to higher rm values, 18 in this case. For continuous annealing, the textures are the same for both types of steel; they essentially correspond to that of a BA rimmed steel, and are perhaps even somewhat weaker. Because of the reduced intensity of the {Ill} fibre, the "- values for both types of steel generally fall in the range 1,1-1,2. The development of strong CRA textures and high rm values in conventional AK steels requires, among other factors, the slow heating rate associated with the tight coil box annealing process.
Annealed ordinary interstitial free Nb and Ti stabilised steels

cp,-

15

Orientation stability map for rolling presented in tP2= 45 section (Bunge notation) (after Ref. 45)

Published by Maney Publishing (c) IOM Communications Ltd

early stages of recrystallisation, the RDII(110) and ND II 111) fibres decrease in intensity in both steels. The <Ill) fibre decreases to a greater degree in the rimmed than in the AK steel. On further recrystallisation and grain growth, the (Ill) fibre increases in strength in both steels, though to a greater degree in the AK steel. As a result, the strongest component of the texture is the {111}(110) at ",5,5 x R in the AK steel (see Fig. 16a); because of the trends described above, the {111}(110) is only ",3,0 x R in the rimmed steel. The CRA rm reported for the AK steel was 1'5, while it was reduced to 12 for the rimmed steel because of the lower intensity of the {Ill} fibre. The ODF for the CRBA AK steel (0'310/0Mn, 0'018%S) of Fig. 16a should be compared with that pertaining to a current low Mn, low S, AK steel (0'20%Mn, 0'008%S), shown in Fig. 16b.27 The AK steel investigated by Heckler and Granzow was of a composition and processing history typical of its day (1969). Subsequent advances in steelmaking technology, which provided for desulphurisation, permitted reductions in the Mn content. This, combined

<

The textures of annealed IF steels are of particular interest because the plastic properties, especially the r-values, depend so sensitively on the texture. The CRBA textures and inplane r-value distributions differ significantly from those for the rimmed and AK steels. In general, both the Nb and Ti IF steels display CRBA r m values of about 2 or higher. At equivalent cold reductions of about 60%, the Nb steels usually have an r-value distribution 1'0 < 1'45 < 1'90' while the Ti steels usually have 1'0 > 1'90' Increasing the amount of reduction results in moderate increases in the CRBA 1'0 and 1'90 values, and in a more substantial increase in the 1'45 values for both steel types. The 1'45 values seem to increase more rapidly with cold reduction in the Nb steels, so that occasionally 1'0 < 1'45 > 1'90' These changes in r-value are intimately associated with changes in the volume fractions of the individual texture components, and in their relative contributions to the r-value distributions shown in Fig. 5. ODFs for the two highly alloyed CRBA ordinary IF steels, discussed above, 0'24%Nb and 0300/0 Ti, are presented in Fig. 17a and b. At first glance, the textures appear to be equivalent, but the planar distribution of r-values in the two steels27,28 is quite

90 80 70 60 ~ 50
"C

""~""""-~~"""""""""'-""'--I-~~~~~&..

....- -1.-.-'-

~40 30 20 10

O;-~---r_.,.... ...,r--...,... ..._t~+_-r-~-.p.. .... ~~~~__..~~ (b)


o
16 10 20 30 40 50 60 70 80 90 10 20 30 40 SO 60 70 80.90

8,deg
tP=45 sections (Roe notation) of ODFs of a AK steel, CR 60% and laboratory BA at 738C for 10 h (L-K composite sample) (after Ref. 26) and b AK steel (low Mn, low S), commercially produced 650/0CRBA, (RD composite sample) (after Ref. 27) International Materials Reviews 1994 Vol. 39 No.4

142

Ray et at.

Textures

in low and extra low carbon steels

90 80 70 60

-I-~~~""',....""~~~~""~""~~-~~~~~~~"""'"

if
"'C

50

~ 40 30 20 10
O-Pi -- --p.;.;........,.;.....,."
... -;. .... ...... .:....

o
17

10

20

30 40

50

60

70 80

90

10

20

30 40

50

60 70

80 90

9,deg Published by Maney Publishing (c) IOM Communications Ltd


iP=45 sections
(L-K composite (Roe notation) of ODFs of CRBA ordinary samples) (after Refs. 27,28) IF steels containing a O24%Nb and b O30%Ti

different, even though the r m values are the same; this is indicated in Table 4. In contrast to the CRBA AK steels, in neither of these materials does the {111}(110) component dominate the texture. Rather, {111}(112) is more important: than {111}(110), as are the off {111} axis components such as the {554} and {667} orientations, whichare 5 off the {111} axis, as well as the {557}, which is 10 off the {111} axis. The intensities of selected texture components in the two steels27,28 are given in "Table5. It :is evident from Fig. 17a and b that both steels contain weak {110}(113) component (8=90, tf/= 25),';'which is stronger in the Nb (2'3 x R) than in the Ti (V4 x R) steel. This orientation is a characteristic of .Nb IF steels even at much reduced Nb contents; but is not a notable feature of Ti IF steels at reduced Ti contents. In fact, the textures differ significantly in Ti IF steels when the Ti content is reduced to much below the 030% Ti level discussed here. In many respects, they gravitate towards those of AK steels, while the textures of Nb IF steels do not change much as the Nb content is reduced. This is because Nb is an effective pancaking agent even at solute levels as low as 003%. An ODF difference map27 for the two steels (0'24%Nb-0'30%Ti) is presented in Fig. 18. Positive regions of the map correspond to locations where the orientation distribution (OD) is stronger in the Nb steel. It should be noted that orientations near {112}(110) are stronger in the Nb steel, while those near {111}< 110> are stronger in the Ti steel. As indicated in Fig. 5, the {112}(110) component contributes high values of r45, while the {111}(110) is associated with ro < r90 Thus, the differences in the

planar distribution of r-values shown above are a result of these texture differences. ODFs for the two more lightly alloyed CRBA ordinary IF steels, 0'095%Nb and 0'13%Ti, are presented in Fig. 19a and b (Refs. 27, 29, 30). The texture of the 0'095%Nb steel (Fig. 19a), looks very much like that of the 0'24%Nb steel (Fig. 17a), in spite of the substantial" reduction in alloy content. However, the texture of the 0'13%Ti steel (Fig. 19b), differs substantially from that of the 030%Ti steel (Fig. 17b). Its {111} fibre is very strong, and there is less perturbation with respect to the relative strengths of its components. The difference ODp27 for the latter two steels (0'095%Nb-0'13%Ti) is reproduced here as Fig. 20. As before, orientations near {112}(110) in the 0'095%Nb steel are stronger by up to 4 x R units in this case. Similarly, orientations near {111}(110) are stronger in the Ti steel (4 x R units). The r-values of
90 80 70 60

~ 50 ~40 30 20 10 0

Table 4
Steel 0'24%Nb O'30%Ti

Planar . distribution (Refs. 27, 28)


Condition CRBA CRBA

of

r-values,

IF steels

10 20

30

40

50 60

70 80

90

9,deg

r-.
198 195

'0
125 203

'45

'gO

18

230 1'73

205 229

t/J = 45 section (Roe notation) of difference ODF for the two CRBA ordinary IF steels of Fig. 17 (O24%Nb-O30%Ti) (after Ref. 27)

International

Materials

Reviews

1994

Vol. 39

No.4

Ray et al. 90 ~"'-:-~""""''''''''''~~~'''''~''''--l'-.a--e~~~~~~~~1!!'480 70 60 ~50


"Cl

Textures in low and extra low carbon steels

143

(b)

~40 30 20 10

O .

o
Published by Maney Publishing (c) IOM Communications Ltd
19

~--r---r............ .....,.
10 20 30 40 50 60

~:-,ro:-:-~ 70 80 90

--...-. ~ 10 20 30 40 50

*--I~l.f60 70 80 90

8,deg
tP=4So sections (Roe notation) of ODFs of CRBA ordinary IF steels containing (sheet samples) (after Refs. 27,29,30) a 0.09S%Nb and b 013%Ti,

these two steels and of the low Mn, low S, AK steeI27,29,3o mentioned above are listed in Table 6. As explained above, the high r 45 value in the Nb steel is associated with the increased {112}(110) intensity and the high r90 value in the Ti steel with that of the {111}(110). For the sake of completeness, difference ODFs are presented below for the (Nb IF - AK) and (Ti IF - AK) steels. In order to make this possible, a 0'13%Ti steel ODF (RD composite sample) is first introduced in Fig. 21 and an AK steel ODF (sheet sample) in Fig. 22. The resulting difference ODFs are illustrated in Fig. 23. From Fig. 23a, it is evident that the Nb IF steel contains higher intensities of the {554}(225) (2 x R), {110}(113) (2 x R), {112}(110) (3'1 x R), and near 8 = 40, t/J= 75 orientations (4 x R). Conversely, the AK steel contains more of the {111}(112) (4'2 x R). The difference ODF for (0'13%Ti - AK, low Mn, low S) is illustrated in Fig. 23b. The Ti IF steel contains higher intensities of {554}(225) (2 x R), {111}(112) (1'8 x R), and the {557} and {667} components at t/J=70, which are stronger by 22 x Rand 23 x R, respectively. The differences in texture between these BA steels and those processed by continuous annealing (CA) are probably small as the r-value differences are quite small." It has also been shown27,29,3o,38 that for ordinary Nb stabilised IF steels, annealing conditions and heating rates have virtually no effect on r-value.
Annealed extra low carbon Nb and Ti stabilised steels

Discussions of ELC IF steels almost always focus on the factors affecting r-value, while texture data are sparse. There is a need for a unifying investigation of
Table 5
Steel 0'24%Nb 030%Ti

the HR, CR, and CRA textures of this class of steel, carried out in such a way that the results are comparable. The apparent paradox pointed out by Hutchinson et al.39 between the modest dependence of r m on Nb content in ELC steels'" and the sharp dependence observed in ordinary Nb IF steels'" (see Fig. 24), probably arises because of a lack of comparability. The paradox involves 'the conclusion' that excess solute Nb is necessary to ensure a high rm value, which is not borne out by the ELC (ULC) steel data. It appears instead that no paradox exists, and that it arose because of an attempt to compare two sets of non-comparable data. First, it must be recognised that the Mn contents of sheet steels produced in the 1960s and 1970s were typically in the 0,3-0,4% Mn range. The materials investigated by Hook et al.28 all had Mn contents in this range. The "o values for these steels lay in the range 1,2-1,5. The exception was a 01%Mn steel that had a substantially higher ro value of 18. Now current ELC IF steels have Mn contents of less than 0200/0,typically 0,10-0,15%. The Mn level is known to have important effects on r-value and texture, as will be shown in the section 'Effect of substitutional elements' below. It can be reasonably inferred that had the steels studied by Hook et al. contained Mn levels as low as those of the ELC steels referred to by Hutchinson et al., their rm values would have been higher, especially at solute Nb contents below 0030/0. In addition, the ordinary IF steels had been cold reduced by 60%, whereas the ELC steels received reductions in the 75-790/0 range. Thus the 'apparent' paradox can be resolved on the basis of the combined effect of both a reduced Mn content and the use of substantially higher cold reductions on increasing the r-values of ELC Nb IF steels compared with the ordinary type of Nb IF steel.

Strength

of selected components,
{554}(225) 79 85

CRBA IF steels ( x Random) (Refs. 27, 28)


{557}(472) 83 69 {667}(121) 84 79 {111}(112) 80 76 {111}(110) 38 54

{112}(110) 46 22

International

Materials Reviews

1994

Vol. 39

No.4

144

Ray et al.

Textures in low and extra low carbon steels

90 80 70 60

90 80 70 60 ~50
"'C

~ 50

~ 40 30 20 10 0 0 10 20 30 40 50 60 70 80

~40 30

(
0 10 20 30 40 50 60 70 80 90

/
90

20 10 0

\
8,deg
21

8,deg tP=45
section (Roe notation) of ODF of 013%Ti CRBA IF steel (RO composite sample) (after Ref. 27)

Published by Maney Publishing (c) IOM Communications Ltd

20

tP = 45 section (Roe notation) of difference ODF for the two CRBA ordinary IF steels of Fig. 19 (0095%Nb-013%Ti) (after Ref. 27)

Figure 25a and b represents the annealing textures of two ELC Nb IF steels containing 00230/0 and 00370/0 Nb, rcspectively.P These steels were cold rolled 75% and then batch annealed. The two textures are similar and are characterised by nearly perfect {111} fibres with intensity maxima located at {554}(225) (8=60, t/J=OO) and {667}(121) (8=50, t/J = 60). There is a perceptible difference between the annealing textures of ELC and ordinary (see Figs. 17a and 19a) Nb IF steels. In the latter case, the general intensities are lower and the {111}(112) component is appreciably less intense than the {111}(110). Furthermore, in the ELC grades, off {Ill} axis orientations such as the {554} and {667}, which are 5 off the {Ill} axis, predominate. The greater balance between the two {111} components leads to lower Ar-values (see Fig. 5) and the higher concentration of orientations close to the {Ill} fibre axis raises the r m value. Figure 26a and b illustrates the recrystallisation textures obtained after laboratory salt bath annealing of the Nb + Ti ELC IF stcel'" for which the HR and CR textures were presented above (see Figs. 11 and 14). Evidently, recrystallisation of this material leads to general weakening of the RDII(110-) fibre and appreciable strengthening of the ND II (111) fibre. These effects become more prominent when the cold reduction preceding recrystallisation is increased. Overall, the intensities of the {223}(110), {112}(110), and {001}(110) components decrease drastically while there is a corresponding sharpening of the {111}(123) and {111}(112) components. However, the off {Ill} axis orientations {554}, {667}, and {557} dominate the texture.
Table 6
Steel 0'095%Nb 0'13%Ti AK steel (low Mn, low S)

Factors controlling the textures of cold rolled and annealed low carbon sheet steels The development of textures in cold rolled and annealed low carbon steels is controlled by both steel chemistry and by processing parameters such as the hot band texture and grain size, coiling temperature, amount of cold reduction, and heating rate during annealing. Of these, the effects of hot band texture and grain size will be dealt with in the next section, while those of the interstitial and substitutional solutes will be discussed in detail in the sections 'Effect of interstitial elements' and 'Effect of substitutional elements', respectively, below. Since the subject has already been reviewed by Hutchinson," only brief reference will be made here to the effects of the processing variables. In a recent conference, Hutchinson" described how the various material and process variables are to be controlled so that the texture/anisotropy of cold rolled 9 0 -I--.L...-.,...a..-...I....r-~~+-~~+T-r-+--+80

70 60 ~50

~40 30 20 10

(
'1

Planar distribution of r-values for IF steels and AK steel (Refs. 27, 28)
Condition CRBA CRBA CRBA
(m

1
10 20 30 40 50 60 70 80 90

O-t-........ -...-. ...I..l.~...,.""-I.Io.,...~ .......,.

(0

(45

(so 205 254 235

o
22

193 202 177

1'49 207 1'83

208 1'73 144

8,deg tP= 45 section


(Roe notation) of OOF of low Mn, low 5, CRBA AK steel (sheet sample) (after Ref. 27)

International

Materials Reviews

1994

Vol. 39

No.4

Ray et ale Textures in low and extra low carbon steels

145

90 80 70
0 0

60 ~ ;;40 30 20 10 0 0 10 20 30 40 50 60 70 80 90 10 20 30 40
(a) SO ~ SO

0
0

60 70 80 90
(b)

8,deg

Published by Maney Publishing (c) IOM Communications Ltd

23

Difference ODFs for CRBA steels a O095%Nb IF (Fig. 19a) -AK (Fig. 22), both sheet samples, and b O13!cJTi IF (Fig. 21) - AK (Fig. 1Gb), both RD composite samples (after Ref. 27)

and annealed sheet steels can be optimised; he also ranked these variables according to their significance. Table 7, reproduced. in slightly modified form from the above work, shows how different and conflicting requirements can be satisfied to produce the best possible results. One important such case concerns the interaction between coiling temperature and heating rate during annealing, as can be seen in Fig. 27. In Table 7, the term 'IF steel' is used to include both ordinary as well as ELC IF steels. It will, however, be useful to bring out the differences between these two grades of steel by comparing their responses to several important material and process variables for the purpose of attaining high r-value. Three factors significantly influence the formability of ELC IF steel sheet: the composition, annealing method, and degree of cold reduction. Of these factors, the effect of cold reduction is the most substantial. For example, Fig. 28 shows the relationship between degree of cold reduction and r m value for two (Nb + Ti) IF steels produced by laboratory cold rolling of rolled hot band 28 mm thick." The steels had the compositions given in Table 8. The rm values at 600/0 reduction and below are poor relative to those at 75% reduction and above for these two batch annealed steels.
Table 7 Control of parameters (Ref. 50) for optimising

Table 9 lists r-values for six ELC IF steels: two Ti alloyed, two Nb alloyed, and two Nb + Ti alloyed. These results were obtained from mill produced strand cast slabs that were laboratory hot rolled, cold rolled, and annealed. Batch annealing was conducted at 732C using a 4 h soak. Continuous annealing was simulated using resistance heating to 843C for 20 s. Figure 29 depicts the relationships between rm value, cold reduction, and annealing method. It is evident that these ELC IF steels require higher cold reductions to achieve equivalent rm values, of about 20, than the ordinary much more highly alloyed IF steels. Clearly, the rm values of the ELC IF steels are somewhat inferior for continuous compared with batch annealing. However, it has been showrr'" that heating rate has no measurable effect on the r-value of the more highly alloyed Nb IF steels.

Effect of hot rolled grain size on cold rolled and on cold rolled and annealed textures
Hot rolled grain size has been shown to have a very significant effect on both CR and the subsequent CRA textures. Hot band texture also has a significant effect on CR and CRA textures. When evaluating
of cold rolled and annealed steel sheets

texture/anisotropy

Low carbon steels Parameter Carbon content Manganese content Microalloying (AI, Nb, or Ti) Soaking temperature for hot rolling Hot rolling schedule Finish rolling temperature Coiling temperature after hot rolling Cold rolling reduction Heating rate of anneal SA Low (*) Low (*) AI (***) High (***) CA Low (**) Low (**)
(#)

IF steels SA or CA Low (***) Low (***)

Low (*)

(#) > A3 (**)


Low, <600 C ,...,70%
D D

(***)

(#) > A3 (**) High, > 700 C (***)


D

NbfTi (***) Low (*) (**) > A3 or <A1 (**) High (*) ,...,90%
(#)

,...,85% 5-20 K S-1 (**) ,...,850 C


D

Maximum temperature of anneal


(#)

20-50 K h -1 (***) ,...,720 C (***) vital.

,...,900 C
D

not critical; (*) significant;

(**) important;

International

Materials Reviews

1994

Vol. 39

No.4

146

Ray

et

a/.

Textures in low and extra low carbon steels

Table 8

Compositions of ELC (Nb + Tl) IF steels used to determine effect of % cold reduction on rm value, Fig. 28 (Ref. 31)
Composition, wt-% Mn AI N Nb Ti

2.2

ULC

2.0
0045 0038 00011 00012 0009 0'007 0030 0026

Steel

3 4

00036 0'0033

011 0'10

IF results concerning the effect of hot band grain size on CR and CRA textures, care must therefore be taken to ensure that the methods used to vary the grain size do not also vary the hot band texture. The effect of initial grain size on the cold rolled texture in a rimming steel was studied by Jones, Hudd, and Dasarathy, and reported by Dasarathy." No significant difference in the cold rolled texture was observed when the hot band grain size was varied over the range from ASTM 5-6 to ASTM 12-13. However, on recrystallisation by means of a rapid anneal at 700C, the annealing textures differed widely. The {Ill} ND fibre was strengthened appreciably in the fine grained material; by contrast, the {llO}(OOl) orientation was intensified in the coarse grained steel. It was shown by Hook29 for a Nb stabilised ordinary IF steel that the hot band grain size has a
Table 9 Extra low carbon IF steels: compositions
Composition, wt-% Nb (solute) Ti (solute)

1.6

1.4 0

0.02 0.04 0.06 o.~ Nb IN SOLID SOLUTION, wt-%

Q1

24

Effect of dissolved Nb on rm values of ELC and ordinary IF steels (after Ref. 39)

Published by Maney Publishing (c) IOM Communications Ltd

profound effect on the CR texture and thus on the subsequent CRBA texture. In steels of this type, it is not possible to increase the hot band grain size by a subcritical (below the AC3 temperature) anneal. A hot band sample was therefore austenitised at 982C, air cooled to 694C, and then held for 1 h to ensure complete restabilisation of the steel. The anneal produced an increase in the hot band grain size from an

and r-values

Steel

Mn

AI

Nb

Ti

A - ELC (Ti)
B- ELC (Ti)

C- ELC (Nb) 0- ELC (Nb)


E - ELC (Nb + Ti) F- ELC (Nb+ Ti)

00024 00043 00028 00027 00033 0'0033

0'10 0'16 0'18 0'16 0'08 0'15

00052 00043 00070 00064 00051 00080

0071 0067 0037 0059 0059 0043


ratio

0'0017 0'0047 00017 00038 00013 00041

<0002 <0002 0023 0037 0028 0'008

0001 0016

0068 0080 <0'002 <0002 0006 0055

0045 0040

0003 0'016

Plastic strain Batch anneal Steel HR coiling temp., DC %CR

Continuous

anneal

'0

'45

'90

r-.
1'97 2'13 217 2'19 202 211 211 2'23 191 190 205 220 209 2'16 220 226 176 2'14 215 211 188 210 207 211

'm
164 217 214 222 1'77 185 1'96 201 179 182 203 209 176 180 202 198 150 197 210 204 157 198 197 199

'0

'45

'90

A - ELC (Tl)

B- ELC (Ti)

C- ELC (Nb)

0- ELC (Nb)

E - ELC (Nb + Ti)

F - ELC (Nb + Ti)

566 566 705 566 566 566 705 566 705 566 705 705 705 566 705 705 566 566 705 566 566 566 705 566

65 75 75 80 65 75 75 80 65 75 75 80 65 75 75 80 65 75 75 80 65 75 75 80

186 201 109 2'16 1'80 188 186 192 183 166 198 196 184 194 196 197 169 180 190 1'77 164 180 187 180

181 197 210 200 192 200 202 222 172 186 189 2'14 194 204 208 220 164 2'12 206 206 179 202 202 208

240 256 254 262 2'46 252 256 255 236 224 2'44 2'58 264 261 267 268 209 2'55 258 252 230 255 238 248

172 220 203 234 165 180 181 198 1'90 164 2'12 210 178 166 202 190 1'56 183 216 2'04 150 180 190 186

135 186 196 193 163 170 182 182 154 1'70 176 186 1'48 188 1'75 184 132 184 180 184 1'48 188 175 184

212 274 262 268 217 222 238 240 220 226 2'50 2'54 2'19 2'16 244 245 180 236 263 243 184 234 2'48 2'40

International

Materials Reviews

1994

Vol. 39

No.4

Ray et al.

Textures in low and extra low carbon steels

147

90 80 70 60
"t:S

-li-~-""~~~~~~~~~~--a._~~~t-:-:'~~~~,.-.~~
1

(b)

~ SO

~40 30 20 10 0 . - .. o 10 20 30 40 SO 60 70 80 90 10 20 30 40 SO 60 70 80 90

9,deg Published by Maney Publishing (c) IOM Communications Ltd


25
l/J = 45

sections (Roe notation) of ODFs of 75% CRBA ELC IF steels containing (sheet samples) (after Ref. 31)

a 00230/0Nb and b 00370/0Nb,

equiaxed 7 urn to 50 urn, and substantial coarsening of the NbC precipitates. It did not, on the other hand, lead to significant changes in the hot band texture; compare Fig.30a and b.27,29 In the same study.i" it was shown that the NbC precipitates in the hot band are substantially coarsened by subcritical annealing at 817C, and that this had no measurable effect on the CR and CRA textures and r-values, verifying the earlier results." The subsequent CR textures produced by processing the two hot bands of different grain size (but of virtually identical texture) differed remarkably, as shown in Fig. 31a and b.27,29 The difference ODF in Fig. 32 shows that the steel with the coarser hot band

grain size had, following cold rolling, a much weaker partial (110) fibre.F In particular, the intensity of the {112}(110) component in the CR sheet decreased from 113 x R to 72 x R as a result of the increase in hot band grain size. The resulting annealed texture was also significantly influenced by the initial hot band grain size, as shown in Fig. 33a and b.27 As demonstrated by the difference ODF (Fig. 34), the CRA texture associated with the increased hot band grain size displayedr? (a) a significant reduction in the strength of orientations near {223}(110); (b) reduced {554}(225) and near {110}(113); and (c) enhanced near {001}(110) (()= 0, 1{I=10, 80) and {221}(110) ((}=70, 1{I=90
0 ).

<>
~

{001}<110> {111}<110>

{112}<110> {111}<123>

0 {223}<110>
{111}<112>

\l

30

CJ)
Q.)
"'0

e
60

(a)

(b)

90

90
30 <P" deg 60

90

o + Ti

30 <p" deg

60

90

26

l/J2 = 45

sections (Bunge notation) of ODFs of the Nb cold reductions of a 700/0and b 850ft (after Ref. 44)

ELC IF steel of Figs. 11 and 14 annealed after

International

Materials Reviews

1994

Vol. 39

No.4

148

Ray et al.

Textures in low and extra low carbon steels 2.2 -t--------Ioo..- ...... CT 2.0

20

---

..

-----'---.a...---1

"" ."

" ~igh

----~
1.8 1.6
Q)

IF steels "

'" lower' .... _

~ 1.8

co
>

:J ~E 1.4

I'AI-killed
and rimming

"

---

~
f-l5 1.6

Batch Annealed at 704C

1.4

I
1.2
1.0 0.1

steel. high
coiling

-- ---8A

Rimmin~/

'"

....... temperature
CA 28

1.2 -t----...T..------,r----.,.---r----r--.,.--or-----1 50 60 70 80
% COLD REDUCTION

Steel3 Steel4

90

steel low CT...,. /

10

100

1 000

Effect of amount of cold reduction on rm values of two CRBA Nb + Ti ELC IF steels (see Table 8) (after Ref. 31)

HEATING RATE, K rnin"

Published by Maney Publishing (c) IOM Communications Ltd

27

Effect of heating rate during annealing on rm value of different steel types with varying coiling temperatures after hot rolling (after Ref. 50)

The changes in the eRA texture resulting from coarsening the hot band grain size also changed both the magnitude and distribution of r-value to ro = 1'4, r45=1'2, r90=1'7 from ro=1'5, r45=2'1, r90=21. Hutchinson et al.39 also carried out a systematic study of the influence of hot band grain size on subsequent annealing textures in ELe steels. For this purpose, samples from a commercially processed low carbon steel hot band (0'240/0Mn, 0'0070/0P, 0'016%8, and O'll%AI) were heat treated and decarburised to yield a final carbon content of 0001 % in four different grain sizes', 13, 18, 39, and 94 um. They observed some. differences in the weak starting textures in the four hot rolled materials, which gave rise to detectable differences in the cold rolled textures. These relatively small differences were, however, considerably amplified by recrystallisation in the final annealed textures. The ODFs of the 70% cold rolled and annealed specimens with the four different initial grain sizes are illustrated in Fig. 35. Annealing was carried out using a simulated continuous annealing cycle, which involved 2 min of holding at a maximum temperature of 800oe. Figure 35 shows that an intense ND fibre is developed in the finest initial grain size material, with only a weak spread along the RD fibre. Even a modest increase in the initial grain size produces a significant reduction in the intensity of the ND fibre, along with some strengthening of the minor components. With further increases in initial grain size, a continued weakening of the ND fibre is observed, especially in the vicinity of {111}<110), together with a notable increase in the intensity of the Goss component {110} <001). Hutchinson et al.39 also investigated the effect of cold reduction on the subsequent annealing texture. They studied two hot bands of a commercially processed lowe steel, with grain sizes of 134 and 39 urn, respectively. They found that, for the finer grained starting material, the ND fibre becomes sharper and more intense but does not change in character as the
International Materials Reviews 1994 Vol. 39 No.4

amount of cold deformation is increased. By contrast, in the coarser grained steel, though some strengthening of the ND fibre takes place with increasing cold work, this is accompanied by a change in the position of the maximum from {111}<110) to {111}<112). All these features are shown in the texture plots of Fig. 36. The final annealing texture is thus a function of the hot band texture, hot band grain size, and amount of cold rolling before annealing. The combined effects of the last two factors on the r m and I1r values of very low C steels are depicted in Fig. 37. It is clear from these figures that in order to optimise the final texture, the hot band grain size should be held below 15 urn and preferably around 10 urn. Thus there seems to be a very powerful effect of grain size, even within the limited range (~10 to ~ 30 urn) that is represented by normal industrial practice.

Effect of interstitial elements


Carbon - dissolved and precipitated Although high levels of C have been known to be undesirable in deep drawing steels, Fukuda'f was perhaps the first to demonstrate convincingly that e has a deleterious effect on plastic anisotropy. Figure 38, taken from his work, shows the interrelationship between total carbon, amount of cold reduction before annealing, and mean plastic strain ratio r m in some vacuum melted steels. These results were obtained on steels that were box annealed and in which the slow heating rates ensured that the C distribution throughout the material was close to the equilbrium level. Three different effects of carbon have been identified, which all reduce the strength of the {Ill} texture components in steels. These are caused by: (a) e present in solid solution during cold rolling, 53-56 (b) e present in solid solution during annealing, 56-59 and (c) e present in combined form as cementite or pearlite before cold rolling.54,55,6o Although there has been substantial progress in understanding the effect of C on annealing textures, controversy still exists over the respective contributions of the above factors, since they are sometimes difficult to separate experimentally.

Ray et al.

Textures in low and extra low carbon steels

149

2.3 42.2 2.1

__

...a...__

.L-_-lL.--_--I__

---I...,....~7"'_t

2.3
ELC Nb IF

ELC Nb IF

2.2 2.1

"a

= ~
f-45

QJ

2.0
"a

1.9 1.8 1.7 1.6


(a)

= ~
f-4

QJ

2.0 1.9 1.8 1.7 1.6

Steel C-BA Steel C-CA

Steel D-BA Steel D-CA

1.54----r----.,.---.,,....------,----,.----t 70 80 60 % COLD REDUCTION 2.3 -+-__ .1-... 1..-.._----10 __ ---&..__

90

1.5 60 2.3 2.2 2.1


QJ "a

70

80

90

% COLD REDUCTION

-1...,...._-1

Published by Maney Publishing (c) IOM Communications Ltd

o
2.2 2.1 2.0
QJ "a

Steel B-BA Steel B-CA

ELC (Ti+Nb)

IF

= ~
f-4

1.9

= ~
f-4

2.0 1.9 1.8 1.7

5 1.8

Steel E-BA Steel E-CA


CT= 566C

1.7
ELC Ti IF

1.6
(c)

CT= 566C

1.6 90 1.5 60 70

1.5-+---.,.---.,r------,.----r----.----I 60 70 80 % COLD REDUCTION


29

(d)

80

90

% COLD REDUCTION

Relationships between rm value, cold reduction, and annealing method (BA and CAl for a steel C, b steel D, c steel 8, and d steel E of Table 9 (after Ref. 31)

Carbon and cold rolling texture Lavigne et al.54 have shown that the amount of solute

carbon as well as the size and morphology of cementite particles have only a marginal effect on the cold rolling texture. They obtained similar (200) pole figures of cold rolling textures in two samples of an Al-killed steel in which soluble C concentrations as well as the size and distribution of cementite particles differed widely. Although solute C does not produce any visible change in the cold rolling texture, it is nevertheless likely to influence crystal rotations during cold deformation." C atoms normally occupy octahedral sites in the bee lattice. The probability that a C atom lies on a {lID} plane is 173 times higher than of it lying on a {112} plane, both of which are active slip planes in bee crystals. 56 During deformation, dislocations can pile up in the vicinity of grain boundaries. The interaction between dislocations and solute C atoms can then produce more work hardening on the {ll0}(111) slip systems, so that further deformation is transferred to the {112}(111) systems. Although crystal rotations within the grains during cold rolling do not seem to be heavily influenced by C in solution, the rotations in the vicinity of grain boundaries may

be significantly affected." The local textures produced close to grain boundaries can therefore be affected by the solute C level, though these differences may not be readily apparent in the overall macroscopic texture. In view of the important role played by grain boundaries in the nucleation of new grains during recrystallisation, it would therefore be of interest to investigate the microtextures produced near grain boundaries during cold rolling and in particular to establish how these are affected by the solute C level.
Carbon and annealing texture

Annealing of cold rolled low C steel sheet involves: (a) partial dissolution of cementite, during heating, liberating some carbon which diffuses into the matrix, (b) recovery, and (c) recrystallisation of the deformed ferrite. The level of solute C in the matrix during annealing depends on the amount of C in solution in the matrix at the end of cold rolling and on how much C is liberated from the cementite particles. Abe et a1.55 contend that C is released from cementite particles into the matrix even during cold rolling by a kind of mechanical dissolution process. Ushioda et a1.61 proposed a quantitative analysis of texture development during the continuous
International Materials Reviews 1994 Vol. 39 No.4

150

Ray et ale

Textures in low and extra low carbon steels

90 "-1 80 70 60
"CS

~50

~40 30 20 10 0 0
(a)

10

20

30 40

50

60

70 80

90 10

20

30 40

50

60 70

8,deg

80 90 (b)

Published by Maney Publishing (c) IOM Communications Ltd

30

t/J = 45 sections

conditions

(Roe notation) of ODFs of 0095%Nb (sheet samples) (after Refs. 27,29)

ordinary IF steel in a HR and b HR + austenitised

annealing of cold rolled low C steel sheet. They suggested that the final texture is determined primarily by a competition between the dissolution of cementite and recrystallisation of the deformed ferrite matrix. The amount of C that is liberated by the dissolution of cementite during annealing is a function of hot band coiling temperature, annealing temperature, and heating' rate during annealing. When high coiling temperatures are employed, the eutectoid transformation occurs during slow cooling, so that the carbide constituents become coarse and widely dispersed (leading to a large interparticle spacing, A). In contrast, under low temperature coiling conditions, rapid cooling through the transformation produces a more uniform dispersion of fine carbides (leading to a smaller interparticle spacing). Coarse carbide dispersions are known to be favourable to the development of {111} texture during rapid annealing.v' This is
11.2

because there will be much less C dissolved into the ferrite matrix from the widely spaced coarse carbides during rapid annealing. This effect appears to be strong enough to counterbalance the detrimental effect of the nucleation of grains of random orientation around the carbide particles during recrystallisation. Figure 39 shows how the mean plastic strain ratio rm of fully processed steel sheet varies with coiling temperature. It is now evident that the presence of coarse hot band carbides favours the development of desirable annealing textures. This arises because fine carbides are able to dissolve more rapidly and to diffuse C into the ferrite matrix more effectively than coarse carbides. Such a mechanism has the potential to play a role during annealing, since the activation energy for the dissolution of cementite, 23 kcal mol-1 (Ref. 54) is substantially less than that for recrystallisation, 82 kcal mol-1 (Ref. 61). Ushioda et al.61,63

90 80 70 60
"CS

~t'l~
0 4

~50

~40 30 20 10

o 10 20 30 40 50 60 70 80 90 10 20 30 40 50 60 70 80 90
8,deg
31
t/J

= 45 sections (Roe notation) of ODFs of 0095%Nb ordinary austenitised + CR conditions (sheet samples) (after Refs. 27,29)
Materials Reviews 1994 Vol. 39 No.4

IF steel in a HR + CR 700/0and b HR +

International

Ray et ale

Textures

in low and extra low carbon steels

151

90 80 70 60 ~50
"'0 "'0

90 80 70 60 ~SO ;:;: 40

;:;:40 30 20 10 0 0 10 20 30 40 50 60 70 80 90

-0.5

-0.5

30 20 10 0 0 30 40 50 60 70 80 90

-1

e,deg Published by Maney Publishing (c) IOM Communications Ltd


32
t/J = 45 section (Roe notation) of difference ODF [(HR + CR) - (HR+austenitised + CR)] for O095%Nb ordinary IF steel of Fig. 31 (after Ref. 27)

e,deg
34 t/J = 45 section (Roe notation) of difference ODF
[(HR + CRBA) - (HR+ austenitised + CRBA)] for O095%Nb ordinary IF steel of Fig. 33 (after Ref. 27)

made a detailed investigation of the kinetics of cementite dissolution. Figure 40 shows calculated C concentration profiles in ferrite as a function of distance from the ajFe3C interface for different annealing temperatures, heating rates during annealing, and interparticle spacings. It is evident from the above figure that the dissolved carbon content of the ferrite can be minimised by having widely spaced cementite particles in the hot band. The calculated average C concentration as a function of heating rate and annealing temperature is presented in Fig.41 for two different interparticle spacings A. For lower heating rates C'" 10-2 to 10-1 Ks-1), the level of solute C is essentially at equilibrium at any temperature, irrespective of the fineness or coarseness of the cementite dispersion. Figure 42, also taken from the work of U shioda et al.,61 shows a plot of calculated average C concentration for
90 80 70 60
"'0

different heating rates and three different interparticle spacings. For batch annealing ("" 10-2Ks-1), the dissolved C content of the ferrite is practically independent of the interparticle spacing and corresponds to the equilibrium value. This behaviour changes as the heating rate is increased beyond about 10-1 K S-l, with the result that coarse, widely spaced carbide particles lead to a reduction in the amount of solute C, because of the long distances and short times available to the C atoms for diffusion. The above research has established the importance of keeping the solute C level low during continuous annealing. If the hot band is coiled at a high enough temperature so that coarse and widely spaced carbides can form, and if the steel is heated rapidly after cold rolling, the recrystallisation of ferrite will take place before any significant re-solution of C can occur. The absence of solute C seems to be instrumental in the

>(
-0) 1
___
rm

~ 50

0 1
rm

~40 30 20 10

= 1.93

= 1.41

o
(a)

10

20

30 40

50

60

70 80

90

10

20

30 40

50

60 70

80 90

e,deg

(b)

33 t/J = 45 sections (Roe notation)

of ODFs of O095%Nb ordinary IF steel in a HR + CRBA and b HR+ austenitised + CRBA conditions (sheet samples) (after Refs. 27,29) International Materials Reviews 1994 Vol. 39 No.4

152

Ray et al.

Textures in low and extra low carbon steels A {111}<112> Y{111}<110>

{O0 1} < 110 >

{11 O}< 00 1 >, Gass

2 1 7

(b)

Published by Maney Publishing (c) IOM Communications Ltd

(d)

35

tP2 = 45

sections (Bunge notation) of ODFs of sheets cold rolled 700/0 and annealed, for different initial grain sizes do (after Ref. 39)

production of strong {Ill} textures, though the precise mechanism of this effect is less clear. Research has been carried out to ascertain the stage of the annealing process at which C is influential in changing the texture. Kubotera et al.57 suggested that the critical stage occurs rather early, probably before recrystallisation is optically visible. Later work by Hutchinson and Ushioda64 demonstrated convincingly that the dissolved C is most effective during

recovery and the early stages of recrystallisation. As mentioned above, during and after cold rolling, solute C as well as C atoms liberated from cementite are expected to segregate to dislocations; this will inhibit recovery in the early stages of annealing. Since the mobility of C atoms is so high at the annealing temperatures involved, they are expected to have only a negligible direct effect on the movement of dislocations, subboundaries, and grain boundaries. It therefore seems likely that the C atoms interact with substitutional atoms such as Mn to form relatively immobile complexes, which are much more effective in retarding recovery. There is increasing evidence that such interactions, not only between C and Mn, but also between other interstitial-substitutional pairs, are indeed important during annealing. The formation and characteristics of such complexes are discussed in detail in the next section. The action of the C-Mn complexes in inhibiting recovery can produce a distinct change in the annealing texture. Dillamore et ale 65 have shown that,
"#. 80
z. (a) (b)

o
i=

~
~70

~ ::::i ~
u

: n1 U.O
,n~
o.~
INITIAL GRAIN SIZE,JLfl\llag scale) 60\-I1)'----....L----L--~---'-' 20 30 50 100 INITIAL GRAIN SIZE: p,m

~ 60 '-:I10:---~20:---~30:---~S~0 --100""'"

J'

a do= 134 urn, CR 60%; b do=39 urn, CR 60%; c do= 134 urn, CR 80%; d do=39 urn, CR 80%

a rm values;

b Sr values

36

Effect of cold rolling reduction on annealing texture for two different initial grain sizes do (after Ref. 39) Materials Reviews 1994 \.101. 39 No.4

37

Effects of initial grain size and cold rolling reduction on calculated anisotropy of very low carbon steels (after Ref. 39)

International

Ray
2.0

et

al.

Textures in low and extra low carbon steels

153

~---r-----.------r----r-----"'"
1.8

0.001%C 1.5
Q)

1.6
0.02%C ::J

~ 1.4
L-

1.2
1.0 0.09%C

1.0 550 600 650 700


C

750

COILING TEMPERATURE, 0.5


L...- __ -L.- __ ---'.L-- __ ~ __ ___'

50

60

70
COLD ROLLING

80
REDUCTION,

Published by Maney Publishing (c) IOM Communications Ltd

90 %

39

Dependence of mean plastic strain ratio r m on coiling temperature (after Ref. 62)

38

Effects of carbon content and cold rolling reduction on mean plastic strain ratio rm (after Ref. 52)

after cold rolling, the stored energy in a sample depends on the local orientation, and varies from one orientation to another. These differences in stored energy provide the driving force for strain induced boundary migration (SIBM), which can lead to the preferential formation of recrystallisation nuclei in low stored energy regions such as {001}(110). Prolonged recovery will reduce these differences in stored energy, thereby minimising the driving force for SIBM. The reduction of the intensity of the {001} component can bring about a strengthening of the {Ill} texture. Steels with high solute C contents are prone to dynamic strain aging and shear banding during cold rolling.t" Matrix grains with {111}(112) orientations are the most amenable sites for the formation of shear bands. During subsequent annealing, grains of the Goss orientation generally nucleate in these heavily deformed regions. The nuclei then grow rapidly at the expense of the {111}(112), with which the Goss has a 35 (lID) orientation relationship. This does not differ much from the 27 (110) relationship, which is responsible for the fastest growth rate in bee metals and alloys."? Thus the nucleation and growth of Goss oriented grains in shear bands leads to the weakening of the {Ill} texture.
Extra low carbon steels

The foregoing shows that carbon plays a dominant role in the development of textures in cold rolled and annealed low carbon steels. An idea of the importance of this role can be gained from Fig. 43, taken from the work of Hutchinson et al.39 This figure was prepared on the basis of the results obtained by numerous investigators. It demonstrates that a very low level of carbon ("-'10 ppm) is sufficient in itself for the attainment of good deep drawability, which can be achieved even without the addition of elements such as Ti or Nb. These, by their 'gettering' effects, lead to further reductions in the solute carbon level.

Unlike traditional low C steels, the ELC varieties are relatively insensitive to soaking temperature before hot rolling, coiling temperature, and rate of heating during annealing." Above all, whereas it is important to finish roll conventional low C steels within the austenite temperature range to avoid a drastic decrease in r m by inadvertently hot rolling in the ferrite field, ELC steels may be successfully warm rolled. Hashimoto et al.68 have found that in an ELC steel, warm rolling (deforming in the \J. no recrystallisation region) resulted in satisfactory textures in hot rolled sheet. Subsequent cold rolling and annealing of this material led to the development of a strong {Ill} fibre texture, even after low cold rolling reductions. Recently, there has been increased interest in ELC steels which, when warm rolled with or without a subsequent recrystallisation anneal, can yield materials with high rm values. Hashimoto et al.69 made a detailed study of the effects of solute C content on texture development in an ELC AI-killed grade and in a few IF steels; these were warm rolled and then annealed. These authors found that the most important factor which determines the recrystallisation texture is the solute C content of the steel during warm rolling. If the solute C content is brought down to nearly zero, by the addition of Ti and/or Nb and suitable heat treatment, it is possible to attain r m values of "-'20. The predominant annealing texture component in these steels is "-'{554}(225), which is also present in the cold rolled and annealed grades of IF steel. From their results, Hashimoto et al. 69 concluded that there is no essential difference in texture and deep drawability between warm rolled and annealed low C steels and cold rolled and annealed sheets. In fact, desirable annealing textures and high rm values can be obtained in warm rolled and annealed materials, provided rolling is carried out in a solute C free condition and a good lubricant is used. Finally, it should be mentioned here that there has been an attempt to develop deep drawing quality cold rolled and continuously annealed sheet by using vacuum degassed extra low carbon and nitrogen
International Materials Reviews 1994 Vol. 39 No.4

154

Ray et al.

Textures in low and extra low carbon steels

200

I
o

I S-'

ala.
150

A. = 20 urn, HR = 0'1 K 7Do e

(a)

eto:

A. = 20 J.1m, R = 10 K s-' H 700 e


0

(b)

650

100

650

600
50
E c::a.
c::a. :z
C)

550

600
550

500
~lOO
I

ee

CD

450
I

~ 200
::-

_--,--....----,-"""T""'-~__r-..,....-_r___r- ....
A=80J.1m, HR=0'1

9
(d)

10

C)

--'

en

Published by Maney Publishing (c) IOM Communications Ltd

ala
150

Ks-'

700 DC

a/a.

A. = 80 J.1m, R = 10K s - 1 H

100

6 0

600 50

550
500 400-- 450
0

12

16

20

24

21

32

36

40 0

12

16

20

24

28

32

36 4O

o ISTAHtE ,v-m

40

Results of calculations showing dissolved carbon profiles at various annealing temperatures in steels subjected to two different heating rates (HR) and containing two different interparticle spacings ). (after Ref. 61)

Al-killed steel without any micro alloying addition.?" It has been shown that these steels develop a sharp {111}(112) annealing texture and exhibit an rm value of about 1'8, coupled with very high ductility ( 540/0 total elongation) and a work hardening exponent of around 0'27.70
rov

Nitrogen - dissolved and precipitated

It has been reported that solute nitrogen can influence the development of annealing texture in rimming as well as in extra low carbon steels.71-73 These authors investigated the effect of N on the recrystallisation textures of these steels by changing the N content after cold rolling. When uncombined with other elements, N appears to have an effect similar to that of C. Figure 44, taken from the work of Takahashi and Okamoto," shows that the best results, from the viewpoint of texture formation, are obtained when the amount of N present is 10 ppm. The resemblance between Figs. 44 and 43 (which shows the influence of C in steel), is striking and points to the strong possibility that similar mechanisms are responsible for the effects of these two interstitial elements.
rov

In order to determine the influence of solute N in almost pure iron, Okamoto and Mizui?" vacuum melted an Fe-0'5C alloy and then fully decarburised and denitrided it to produce a high purity Fe with 00005%C. The material was cold rolled 75% and then controlled amounts of N were introduced, after which it was recrystallisation annealed at 650C for 1 h. Figure 45 illustrates the effect of solute N, present during annealing, on the intensities of the various texture components. Here the intensity of the {111} texture increases with N content, while the {200} intensity decreases concurrently. The {110} and {112} components do not seem to be influenced by the amount of N present. These results are consistent with those of Tagashira et al.,75 who showed that the strength of the {lll} component increases with N content in iron alloys not containing any Mn. The discrepancy between the results of Refs. 73 and 74 is apparently due to the much higher alloying element and impurity level (especially that of Mn) present in the steel in the former case. In commercial low and extra low carbon steels, some Mn is almost invariably present. The effect of N on the recrystallisation texture, like that of C, varies with the Mn

International

Materials Reviews

1994

Vol. 39

No.4

Ray et a/.

Textures in low and extra low carbon steels

155

(a)

(b)
700C 150

700C 150,

E 0.
0.

Z
cd: 0

0 ~ 100

650C

100

u w

en en Ci

600C 50 550C

w 50

o cd:
a: w

Published by Maney Publishing (c) IOM Communications Ltd

O .... --.....,;;;:...----~ ~-... .....


10 102 10-2 10-1 10 102 HEATING RATE, KS-1 a A=20 urn: b A=80 urn

41

Average dissolved carbon as function of heating rate and temperature for two different interparticle spacings). (after Ref. 61)

content of the steel." It is well known that the profile of the Snoek peak is influenced by the presence of Mn in iron, which indicates that there is an interaction between Nand Mn atoms." This phenomenon is discussed in more detail in the next section. Research has been carried out to understand and control the development of annealing textures in AI-killed steels during box annealing. Leslie et a1.77 confirmed that in these steels AIN precipitation significantly modifies the recrystallisation texture. After successful processing, these steels exhibit strong

Co Co

fr = 0.01

o zn
a:

() 50

(J)

o w ~

en 5
w a: w

~
10-1 1 10 HEATING RATE, Ks-1

42

Calculated levels of dissolved carbon expected to be present at start of recrystallisation as function of heating rate for steels with three different interparticle spacings). (after Ref. 61)

{Ill} textures, together with relatively weak {001} components. The AIN reaction and analogous phenomena have been discussed in detail by Hutchinson" and therefore only the salient features of this reaction are dealt with here. It is now established that in order to produce strong {Ill} textures by utilising the AIN reaction, processing should consist of the following steps: 1. To begin with, the Al and N must be in solid solution before cold rolling. The required soaking temperature for a steel with known Al and N contents can be calculated from the solubility product of the AIN reaction."? For a typical steel containing 0'0050/0N and 0030/0 AI, the soaking temperature should be '" 1230C to ensure full decomposition of the AIN. 2. The precipitation of AIN in the hot band should be prevented during and immediately after hot rolling. This is achieved by the use of water sprays after hot rolling and coiling at temperatures below '" 600C; the latter leads to rapid cooling in the temperature range for AIN precipitation, i.e. from 900 down to 600C. The Al and N held in solution in this way do not seem to play any special roles during subsequent cold rolling (typically to '" 70% cold reduction). 3. Finally, during box annealing, slow heating provides adequate time for the Al to diffuse to form clusters or precipitates before the commencement of recrystallisation. Sometimes a two stage heating cycle is employed, where a low temperature anneal at '" 550C is followed by a second at '" 700C. The precipitation of AIN takes place at the lower temperature and this is followed by recrystallisation of the steel at the higher temperature." The AIN precipitates
International Materials Reviews 1994 Vol. 39 No.4

156

Ray et al. 2.5

Textures in low and extra low carbon steels

r-..,...-----,~---__r"----_r_ ...
2.0

1,8 2.0
Q)

co

> 16 ~E '

1.5
1,2

1,00 1.0

10

30

40

50 .ppm

60

10

f()

NI TROGEN

44

Published by Maney Publishing (c) IOM Communications Ltd

Effect of nitrogen content on steel (after Ref. 73)

rm

value of unkilled

10-4

10-3

10-2

10-1

CARBON CONTENT, wt-%

43

Effect of carbon content on rm value as estimated from various sources in literature (after Ref. 39)

(or clusters) are known to exert considerable influence on recovery and recrystallisation. They seem to retard the nucleation of grains with orientations such as {001}, while not affecting significantly the growth of the {Ill} grains." Thus the {Ill} eventually becomes the most dominant component of the annealing texture. The interaction between the precipitation of AIN and recrystallisation in controlling the {Ill} texture will be discussed in detail in the section 'Mechanisms of annealing texture formation' below. When Al-killed steels are to be processed via continuous annealing, the opposite strategy must be employed. That is, high coiling temperatures are used instead of low ones. Under these conditions, it is important for all the nitrogen to be fully precipitated in the form of AIN particles. When the AIN is present in the form of a relatively coarse dispersion, its presence does not interfere with grain growth and the associated texture changes." Grain growth is beneficial for enhancing the intensity of the {Ill} texture, (see the section 'Nucleation sites for recrystallised grains and their growth' below). The addition of boron to such steels can be advantageous, since it forms BN with the N present, which is a more stable compound than AIN; moreover, the BN particles are sufficiently coarse ('" 025 urn in diameter) for them to be almost ineffective in pinning grain boundaries." However, in order to achieve satisfactory normal anisotropy, the C content of the steel should be extremely low.

elements in the steel, such as C and N (Refs. 62, 75, 82-86). According to Hu and Goodman." an rm value of '" 25 is obtained at about 0'02%Mn, and this decreases drastically with increase in Mn content, attaining an approximate value of unity when more than ",0'4%Mn is present. Again, below a Mn level of 0,005%, the behaviour has been found to be erratic and there is a drastic decrease in rm ' It has also been found that, except for the very lowest Mn concentrations, an increase in Mn content retards the kinetics

14 13

E a
"'C

12 11 1.4
1.2
0_0 0
{211}

~ ~

e
0

a:

~ en z
f-

0.8 0.6 0.4 0.2 0 0.10 0.05


-0-0
{110}

~
0

~ cr;
CJ

f-

Effect of substitutional elements


Manganese and its interactions with C and N

10
NITROGEN,

20
ppm

30

Manganese in low C steels is known to be detrimental to the development of deep drawing textures. It can affect the annealing texture either as an element in solid solution'" or by way of interacting with other
International Materials Reviews 1994 Vol. 39 No.4

45

Effect of solute nitrogen on intensities of annealing texture components in high purity iron (after Ref. 74)

Ray et al.

Textures in low and extra low carbon steels

157

0,6
1.8 Annealed 830-(

0,5

1,6

~ -,
::J ~ 1.4 ~E
Q)

0,4

~ ~ < ~
< :I:

rm="O~
0,3

Annealed

700 - I:

0,2

0,1

1.0

0,02

004
CARBON, wt-e4

0,06

0.08

Published by Maney Publishing (c) IOM Communications Ltd

- 0,1

0.1

0,2

K-value (uncombined

Mn), %

47

46

Dependence of rm on amount of uncombined (K-value) (after Ref. 62)

Mn

Combined effect of carbon and manganese in solution on rm value of box annealed steel sheets (after Ref. 82)

of recovery and recrystallisation. It should be mentioned here that the steels studied by Hu and Goodman were laboratory vacuum melted ones and did not contain AI. An explanation for the effect of Mn in solution has been given8788 in terms of solute drag. According to this mechanism, though the earliest grains that nucleate during annealing are predominantly of the {Ill} type, the solute drag effect retards their growth, thereby allowing for the additional nucleation of less favourably oriented grains. For processing under continuous annealing conditions, the effective uncombined Mn level is commonly defined by the K-value,62 where K = [%Mn] -55/32[%S] -55/16[0/00]. The Kvalue, (when positive) is assumed to represent the amount of Mn in solid solution. However, Tagashira et ai.75 have pointed out that since mixed oxides and sulphides are formed with increasing Fe/Mn ratios, the dissolved Mn content should necessarily be greater than the K-value. Figure 46, taken from the work of Toda et ai.,62 shows that the highest rm values are obtained when K lies in the range 0-0'1 %, i.e. when only small amounts of Mn are present in solid solution. (~s mentioned above, the degradation of annealing textures in low C steels has been associated with the simultaneous presence of solute C and Mn in the ferrite. Hughes and Page,82 who varied both the Mn (0'001-0'70/0) and C (0'001-0'08%) levels in their steels, found that an optimum steel composition with ",,0030/0C and ",,015 Mn led to the highest fm value. These steels, which were vacuum melted in the laboratory and did not contain any Al addition, were cold rolled 85% and annealed using a slow heating cycle (box annealing). The salient features of their findings are summarised in Fig. 47. Hutchinson and Ushioda 59 and Osa wa et ai.89 have independently shown that the powerful effect of C on annealing textures is, in fact, due to an inter%

action between dissolved C and Mn. The combined effect of these two elements on rm value is presented in Fig. 48. This clearly shows that the detrimental effect of Mn can be largely eliminated when the C content is sufficiently reduced. There can be a further kind of interaction between Mn and C in low C steel. In Fe-Mn-C alloys, the equilibrium partitioning of Mn, between the ferrite and cementite phases, is a function of temperature." By annealing at lower temperatures, cementite can become further enriched with Mn (Ref. 91). From their study of a low C Al-killed steel, Suzuki and Abe92 observed that, when the cementite is enriched with Mn, the intensities of the {Ill} components are increased while the {110} intensities are decreased. It is known that during cold rolling, C atoms are liberated from the cementite and condense on dislocations." This occurs because the binding energy of a carbon atom to a dislocation (0'5 eV)94 is higher than that of a carbon atom to the cementite lattice (0'4 eV).95Again, the binding energy of a C atom to a Mn atom is higher than that of a C atom to an Fe atom.?" As a result, the amount of C which can be liberated from cementite during cold rolling is decreased when the Mn concentration in the cementite is increased. The enrichment of cementite with Mn also leads to the depletion of Mn in the ferrite matrix, and this will further act to improve the annealing texture. As mentioned previously, there are indications that N like C, can also interact with Mn in steel. Tagashira et ai.75 made a detailed study of the effects of Mn and N on the development of annealing textures. Their results, for both box and continuous annealing conditions, are summarised in Fig. 49. In this figure contour maps of the 1(111)/1(001) intensity ratio have been plotted as functions of the Mn and N concentrations. In general terms, box annealing produces rather sharper textures than does rapid annealing; the difference may be partly due to more extensive grain
International Materials Reviews 1994 Vol. 39 No.4

158

Ray et al.

Textures in low and extra low carbon steels


2.5-t------'----....I..------L--....I..------L--...L---+

2.0 r

-nr- . __

1.8

~ ;:1 ea ~

2.0

Q)

1.6

co

:J

>
E

\.
~

'"'

e
1.5

'- 1.4

0\
0 0.4

0.002%C
1.0 0.00

+---.-----r---,...---r----.-----.--.---r--_.._---.--.---r--_..____+_
0.05 0.10 0.15 0.20
wt- %

0.25

0.30

0.35

Nb (SOLUTE),

1.2 O.007%C 50 1.0

60% CR, SA 746C, Ref. 28;

%CR =

r,

SA 700C, Ref. 109

<, '~.04%C
0.10%C~. ---0.6 0.8 0.2
MANGANESE

Effect of Nb (in solid solution) on Nb IF steel (after Refs. 28,109)

rm

value of a

Published by Maney Publishing (c) IOM Communications Ltd

o
48

CONTENT, wt-%

Dependence of rm on carbon and manganese concentrations of different grades of steel (after Ref. 39)

growth?" during the former process. Optimum annealing textures for deep draw ability were obtained when the steel contained a rather low level of Mn (",,0'080/0) in solid solution in combination with intermediate N concentrations ("" 10-20 ppm). Titanium and niobium Titanium and Nb bearing ELC steels have assumed greater importance recently as the new generation of deep drawing steels suitable for processing by continuous annealing. Extensive work has confirmed the considerable increase in average plastic strain ratio r m that can be achieved by the addition of sufficient quantities of Ti and/or Nb to low carbon steel to fix the C and N.28,98-108 Messien and Greday'?" have shown that the CRBA r m values increase dramatically with an increase in either Ti/( C + N) or Nb/( C + N) ratio, and pass through a maximum in both cases. High "- values were achieved over a restricted range for these ratios

0.3 rft. ~ 0.2

~
I

0.3 rft. ~ 0.2

+.. ~
c 0.1
0 +"

<a.6
2

+..

8 120 N,ppm
and box annealed;

8 120 N,ppm
and

a 49

70% cold rolled rapidly annealed

b 70% cold rolled

Contour maps of 1(111//(001) intensity ratio as function of Mn and N concentration (after Ref. 75) Materials Reviews 1994 Vol. 39 No.4

in each case. The rm values exceeded 20 for Ti contents such that 12 < Ti/(C + N)< 15, and for Nb contents such that 9 < Nb/(C + N) < 12. The result concerning Nb is totally at odds with the results of Hook et al.28 recast and shown in Fig. 50, where the r m values increase dramatically with increasing Nb (solute) content (increasing NbjC ratio) and reach a plateau, but show no maximum. The results of Messien and Gredayl06,109 are replotted in the same diagram in order to compare their results with those of Hook et al.28 In Ref. 106, Messien and Greday did not disclose either the annealing conditions used or the specific compositions of the steels. An obvious reason why the rm values in Ref. 106 decreased beyond some optimum ratio, could be that the laboratory steels were not fully annealed, i.e. fully recrystallised, for the higher Nb contents because of the 'conventional anneals' that were used. This is quite possible in view of the demonstrated extremely strong effect that solute Nb has in retarding the recrystallisation of cold worked ferrite.38,110 Thirteen steel compositions were investigated by Messien and Greday.l'" for which the Nb contents ranged from 0034 to 04450/0. The C + N contents were such that 15 ~Nbj(C + N)~286. If one assumes that the highest Nb/(C + N) ratio corresponds to the highest Nb content, then C + N =0,0156 and Nb (solute) = ",,0,32 wt-%. This level of Nb would have a very significant effect in retarding recrystallisation. Support for this proposition is found not only in Refs. 38 and 110 but also in Gillanders et al.los for simulated batch annealing conditions. In a later report,'?" Messien and Greday listed the specific compositions of the steels and noted that the annealing temperature used was 700C (heating rate 20 K h -1, no soak time specified). They disclosed that the steels with the higher Nb contents were not fully annealed. The mill processed, batch annealed (746C) steels reported by Hook et al.28 were fully recrystallised. The significant decline in r m values at higher solute Nb (or Ti) levels'?" is then in fact a result of the incomplete annealing disclosed by Messien and Greday later.'?" This fact is not well recognised because of the limited distribution of the disclosure.l'" In Fig. 50 the more rapid increase in r m value initially may be due to the low Mn content (0'140/0Mn)

International

Ray et ale
2.5-t--I----.L.------L...----L--...1----J-

Textures in low and extra low carbon steels

159

= -;
;;.
J".

2.0

&

1.5

52 1.0-t--.....----.~-y---.----r-~--..,.-~-.---.....-~ 0.00 0.05 0.10 0.15 0.20 0.25 0.30


Ti (SOLUTE), wt-%

l/J2

= 45 sections for two fully processed ELC steels containing a Ti and b Ti + Nb (after Ref. 39)

51

Published by Maney Publishing (c) IOM Communications Ltd

o 66% CR, SA 738C, Ref. 27; 0 %CR = 7, SA 700C, Ref. 109 Effect of Ti (in solid solution) on rm value of a Ti IF steel (after Refs. 27,109)

of their steel compared with that of Hook et ale (O3-04%Mn). Figure 51 shows the relation between rm value and increasing solute Ti content (increasing Ti/(C + N) ratio), which was obtained by replotting the data from the reports by Messien and Greday.'?? Superposed on these plots are the data for two Ti IF steels containing 013 and 030%Ti, respectively'? Thi~ comparison suggests strongly that the maximum in the curves, which results from a decline in rm values for solute Ti>015%, or Ti/(C + N 17, is merely a consequence of inadequate annealing under laboratory conditions (at 700C using a heating rate of 20 K h-1). It is suggested that the steel with the highest Ti content, i.e. 0187% Ti (solute), was not completely annealed and therefore had a low r m value. The two steels (013 and 030%Ti) investigated by Hook?7 were fully recrystallised and showed similar rm values. It is well known that very sharp near {Ill} fibre textures develop during the annealing of heavily cold rolled (75% and above) Nb and/or Ti bearing low carbon steels. Figure 52, taken from the work of Hutchinson et al.,39 shows that two ELC steels (with ~40 ppm C, ~ 30 ppm N, and ~004%AI), containing either 0089%Ti or 0033%Ti+0028%Nb display almost identical textures after processing by continuous annealing, with the appearance of strong <lll>IIND fibres of nearly constant intensity. The measured r m values in the two cases were ~ 25. Hutchinson et al.39 further observed that removal of the elements Ti and/or Nb from the steel causes a general weakening of the ND fibre, which reduces the r-value fairly uniformly. In addition, several new texture components appear, which further decrease the r 45 value. They concluded that the superior text,!res of the Ti and/or Nb bearing steels are principally the result of the finer hot band grain sizes brought about by precipitation of the Ti and/or Nb carbonitride particles. In contrast to the behaviour of box annealed Al-killed steels, in continuously annealed IF steels precipitation has to occur before cold rolling and not afterwards. Whereas the annealing texture of AK

steels can be characterised as dominated by {111}<110) type components, textures in the Ti and/or Nb IF steels are usually centred on near {554}(225) and on {Ill} off axis components in the range of near {667}(121) to {557}(5 12 5). For the same processing history, and for both over and under stoichiometric compositions in relation to the C + N content, the Nb steels develop sharper annealing textures than the Ti steels.'!' The processing conditions are, however, less critical in the case of Ti steels because of the higher affinity of Ti than Nb for C and N. The development of these characteristic preferred orientations in Ti-bearing steels has been attributed to: (a) the effects of Ti(C,N) precipitates,112,113 (b) Ti in solution,102,114 and (c) Ti as a scavenger of the interstitial solute atoms.'!" Similar effects of Nb in Nb-bearing low carbon steels have also been suggested.Pr" The scavenging effect of both Nb and Ti in removing C and N renders the ferrite matrix almost interstitial free, leading to the formation of sharper ~ {Ill} annealing textures. Sufficient recovery before recrystallisation and the retardation of high angle boundary migration by the precipitated particles further helps in the development of this texture.28,38,111,116When the amounts of Ti and Nb are insufficient to fix completely the interstitial solutes, {110} <001) oriented grains nucleate along the deformation bands, thereby decreasing the intensity of the {Ill} texture. In the case of Nb steels, Nb in solid solution has a greater effect in retarding recrystallisation compared with the effect of the Nb(C,N) precipitates. In fact, an excess of solute Nb has been found to ensure a high rm value, see Figs. 24 and 50. Phosphorus Low carbon Al-killed steels with up to 01 %P, commonly known as the 'rephosphorised steels', are very popular for deep drawing purposes. In addition to its ~ow cost, P provides effective solid solution hardening In steel. Phosphorus additions have been used in steels with both coarse and fine cementite particles and also in steels that were decarburised before cold rolling. Matsudo et al.117 made a detailed study of the effect of P in a few rimmed and Al-killed steels which were cold rolled 75% before recrystallisatio~ annealing. Both slow and rapid heating rates were used to simulate box and continuous annealing, respectively. Their findings are enumerated below and are also shown in Fig. 53:
International Materials Reviews 1994 Vol. 39 No.4

160
2,0

Ray et al.

Textures in low and extra low carbon steels

r------~----~----.,.__--___,
" ' . , Slow heating Rapid heatl ng

ence in the cold rolling and subsequent annealing stages. The texture of the annealed sheet changes with Si content from being centred mainly on the

{111}(110) orientation for O%Si to the {554}(225)


for 1 and 20/0Si steels. A slow heating rate during annealing produces a sharper texture in the 2 % Si steel. The differences in the textural features of the 0 and 20/0Si steels, at different stages of processing, are illustrated in Fig. 54. The r m value has been found to vary in a complicated way with the Si and Mn levels when both are present. The rm values of high Mn-containing steels increase as the Si content is increased from 0 to 2%. In O%Si steels, a significant decrease in rm is encountered beyond a solute Mn content of 01 %; by contrast, in 2%Si steels, rm is maintained at a reasonably high value ( ~ 1'5) for Mn solute concentrations of up to 1.00/0.120 These trends are indicative of some kind of interaction taking place between Si and Mn atoms. Silicon seems to have a beneficial or offsetting effect when the presence of dissolved Mn causes a degradation of the texture. Recently, a demand has arisen in the automotive industry for high strength sheet steels with excellent deep drawability and bake hardenability. This interest led to the study of the effect of Si on the development of recrystallisation textures in a Ti-stabilised sheet steel containing a small amount of C in solution.P' The addition of Si was found to be effective in producing a strong {111} annealing texture, even though solute C was present, for both slow and rapid heating rates. The main component of the annealing texture here is again {554}(225). It has been suggested that Si has a beneficial effect in decreasing the solute C content at the cold rolling and/or annealing stages. Silicon is known to lower the limit of solid solubility of C in et-iron.122 Sulphur Any study of the effect of S on the deep drawability of steel must consider the conditions of sulphide precipitation in hot bands and clarify the roles of soluble S and of MnS during both box and continuous annealing. In box annealing, deep draw ability improves somewhat with decreasing Mn contentr" however, at very low Mn levels, deep drawability deteriorates. This latter effect is considered to be due to the presence of dissolved S in the hot bands.P It should be mentioned here that the steels investigated in Ref. 81 were laboratory vacuum melted and did not contain AI. In the continuous annealing of low C rimmed steel, deep drawability is known to improve when a high coiling temperature is used. The reason is that S is able to precipitate out as MnS under these conditions and thus purify the matrix, which then exhibits improved deep drawability.F" In a recent study of the effect of S on the deep drawability of ULC rimmed steel, Kobayashi et ai.8S have come to the following conclusions: 1. During slow annealing, the presence of large amounts of dissolved S before cold rolling improves the deep draw ability, because the S precipitates out as fine sulphides during recovery and the early stages of recrystallisation. These precipitates act somewhat

<,

....
1,8
CD :J

']'

............

decarburised /Y A/

"'-", ---_

ro

> ~E
1,4

1,2

1,0

--_-----r---------" --------..:._-~------ fine carbides

t'

Published by Maney Publishing (c) IOM Communications Ltd

0,02

0,04 PHOSPHORUS,

0$)6

wt-'.

53

Dependence of on P content for steels with different carbide spacings and annealed at two different heating rates (after Ref. 117)

'm

(i) P, like C, is detrimental to the formation of the {111} annealing texture (ii) when P and C coexist, higher P contents result in higher r m values, though ones which are lower than those of decarburised samples. This suggests a possible interaction between C and P atoms (iii) regardless of the P content and carbide morphology, slow heating during annealing improves the r m values; this indicates that in order to be effective, P must be able to segregate to grain boundaries or sub boundaries before recrystallisa tion. HU86,118 has shown that the r m values of rephosphorised steels are not adversely affected by increasing the Mn content to 0250/0 and the C content to 0050/0. Thus, P seems to be beneficial in situations where the presence of dissolved Mn and/or C may prove deleterious to the formation of satisfactory annealing textures. Ohashi et al.119 have observed that the effect of P on annealing texture formation in Ti and Nb stabilised steels is similar to that in plain carbon steels. Phosphorus has been found to raise the intensity of the {111} component only in insufficiently stabilised steels. Contrary to the suggestions of Matsudo et ai.,117 they have not found that the segregation of P at grain boundaries leads to any beneficial effect. Silicon Silicon is generally used in high strength hot and cold rolled steel sheet as a solid solution strengthener and ferrite stabiliser. A maximum r m value of 17 was attained in a high strength low C steel sheet containing about 2%Si and the steel showed excellent deep drawability.F" Silicon affects the texture of steel right from the hot rolling stage, producing a marked differInternational Materials Reviews 1994 Vol. 39 No.4

Ray et al.

Textures in low and extra low carbon steels

161

0% Si Steel 90 90

2%SI

Steel

3 60
C)
Q)

60
C)
Q)

"C

"C

~
30

30

0 0
(a)

30 a,deg

60 Hot

90 Rolled Sheet

30 a,deg

60

90

O%Si

Steel 90

2%Si

Steel

Published by Maney Publishing (c) IOM Communications Ltd

90

60
C)
Q) Q)

60
C) "C

"C

30

30

30 (b) a, deg

60 Cold

90 Rolled Sheet

30 a,deg

60

90

O%Si Steel 90 90

2%Si

Steel

60
C)
Q) Q)

60
C)
"C

"C

~
30

30

7. 0 30 (c) a,deg Annealed 54 t/J = 45 sections of ODFs of two steels containing conditions (after Ref. 120) Sheet 60 90 0 30 a,deg 60 90

0 and 20/05iin: a hot rolled, b cold rolled, and c annealed

like the AIN clusters in Al-killed steels. By contrast, small amounts of soluble S degrade the deep drawability, since the S in this case precipitates out at a later stage of recrystallisation and inhibits growth of the desirable {Ill} component. 2. During rapid annealing, soluble S deteriorates deep drawability as the dissolved S precipitates out only at the later stages of recrystallisation and retards grain growth. 3. Among the sulphides, (X- MnS particles (100200 nm) have little influence on deep draw ability,

whereas precipitates of fJ-MnS (20-40 nm) in the hot band are detrimental to deep draw ability during both slow and rapid annealing.
Substitutional-interstitial dipoles

The interaction between substitutional and interstitial atoms in bee alloys has been a subject of speculation for many investigators concerned with the interpretation of internal friction data.P" Osawa and Kurihara'F' prepared two steels with markedly different C levels, i.e. <0002 and 004 wt-%, and
International Materials Reviews 1994 Vol. 39 No.4

162

Ray et al.

Textures in low and extra low carbon steels

2.0r---------------C, wt-% < 0.002 0.04 1.8 -0----.0 Cr

-()-- -e-.---

Mn Cr SI

1.6
Q)

-0..... --0_ ... --b-.... - .


...... 6...
.........

--0---0--

-../\- -.t.-

co 1.4 > ~E
1.2
1.0
N

::l

(/)

~ 10' t------t-----t-+-----,..-F-----I--I o a... is u


I
C

(a)

~
LL

E E2.2
'+C>

p-------

o Z o

10-5t----y---t-----+-t---....;;;;:.-t

b o
L!)

~ 2.0 u
1. 8

cV
~p
SI Si

Mn

i=.~ u ~ a: LL a:

~ -I

Published by Maney Publishing (c) IOM Communications Ltd

~
~ 1.6

~ 1061------...."'t----_-+--t-.;:a. / OOer ----0 _ -Mn


(b)

c--+-----+--4

z ~1.
I-

en ~ 1. 2 "--..a.--'""'---'--_''_--'---'---""----'L-..a.--..r.-...a-.-&--L.-.L.--J en 0 0.2 0.4 0.6 0.8 1.0 1.2

.,.-0:,'"-..-.,"

z
~

, I , , I " I ,'---..a.--'""'--....I.-_"_ I -""---""----'''--''''"---..r.--'--.-&--L.-.L.--J o 0.02 0.04 0.06 0.08 0.10 0.12 p, wt-%

Mn Cr, or Si, wt-%

55

Effects of dissolved Mn, Cr, Si, and P in two series of steels containing < 0002 and 004%C on a rm value and b tensile strength at 500C: 1 kgf mm-2==g8 MN m-2 (after Ref. 125)

56

Equilibrium Mn-C dipole concentration as function of temperature for steels containing 002 and 030%Mn and various C concentrations (after Ref. 127)

added different amounts of the substitutional solute elements, Mn, Cr, Si, and P. The tensile strengths of these alloys at 500aC and the rm values of annealed sheets of these materials are presented in Fig. 55. It is of interest that there is a clear cut difference in the effect of the substitutional elements at the two different C levels. The difference is large for the addition of Mn and Cr, whereas it is marginal for that of Si and P. This indicates that Mn and Cr interact more strongly with C than Si or P. This interaction is likely to involve the formation of complexes or dipoles between substitutional and interstitial atoms. These dipoles, in turn, can interact more strongly with dislocations than the individual atoms. It was mentioned above that there is a possibility of a strong interaction between Mn and C atoms in steel, which can play an important role in the formation of annealing textures. Abe et al.126 have suggested that complexes or atomic dipoles of Mn and C can exist in ferrite and that the interaction energy of such dipole formation is 04 eV. Hutchinson and Ushioda64 have proposed a model to describe how such dipoles can pin dislocations at elevated temperatures, thus affecting recovery and the nucleation of recrystallised grains, They assumed that C atoms are bound 'to dislocations through an elastic interaction (interaction energy ",,0,5 eV) and to neighbouring Mn atoms by an electronic interaction (substitutionalinterstitial dipole). In order for the dislocation to
International Materials Reviews 1994 Vol. 39 No.4

move, it is necessary to break the weakest link of the complex or to drag the complex along with it. Since the latter process involves vacancy migration, it can occur only slowly. On the basis of calculations using their model, Hutchinson and Ushioda showed that Mn-C dipoles in steels can interact effectively with subboundary dislocations, thereby modifying the nucleation processes that give rise to recrystallised grains during annealing. Abe127 also calculated the equilibrium density of Mn-C dipoles as a function of temperature. The calculated values were plotted for two steels containing 002 and 0300/0 Mn with various C levels, as shown in Fig. 56. On the basis of a study by Okamoto and Takahashi 128 of a steel containing 018 % Mn, Abe127 concluded that the maximum intensity of the {Ill} component in the annealing texture was achieved when an optimum Mn-C dipole density was present at the onset of recrystallisation. This worked out to be. 45 x 10-6 mole fraction under the experimental conditions of Ref. 128. Using this value in conjunction with Fig. 56, it should be possible to calculate, at least approximately, the optimum amounts of Mn and C which will maximise the intensity of the {Ill} component. This concept indicates that if the C concentration is greater than 16 ppm by weight, the optimum Mn content is about 00240/0. If, on the other hand, the Mn content is 0300/0, then the steel should be decarburised down to a C level of 2 ppm by weight.P?

Ray

et

al.

Textures in low and extra low carbon steels

163

Published by Maney Publishing (c) IOM Communications Ltd

Abe127 also considered the theoretical possibility of AI-N and Mn-N dipole formation in view of the observed interaction between these two pairs of substitutional-interstitial elements. He concluded that the energy of formation of an AI-N dipole is of the same order as that of a Mn-C dipole. The interaction coefficient of N with Mn is much higher than with AI; hence Mn-N rather than AI-N dipoles are expected to be formed in solid solutions of Fe containing Mn, when these are supersaturated with respect to Nand Al.127 The pinning of dislocations by Mn-N and AI-N dipoles will retard recovery in the manner of the Mn-C dipoles. In the former case, however, the dipoles will decompose into their constituent atoms when heated to high temperatures. Thus the AI-N dipoles break down and become free to form AIN clusters (or precipitates), which are known to lead to the formation of a pronounced {Ill} texture in Al-killed low carbon steels during the batch annealing process.

Mechanisms of annealing texture formation


Driving force for recrystallisation Recrystallisation textures of low and extra low carbon steels differ considerably from the corresponding cold rolling textures. In general, with recrystallisation, there is an increase in the intensities of the {Ill} components, while those of the {001} orientations decrease drastically. This difference can be explained in part on the basis of the orientation dependence of the stored energy of cold work. Quantitative electron microscopy107,129-132 and X-ray line breadth measurements107,131-135 have shown that the distribution of stored energy depends on orientation in the cold worked metal. The stored energy in the deformed regions of various grains represents the driving force for recrystallisation on annealing. The Taylor factors calculated by Urabe and Jonas41,44 indicate that the lowest possible value is located near the Goss {110}(001) and the second lowest at the rotated cube {OOl}(llO) position (see Fig. 57). This suggests that these two specific orientations have, in fact, the lowest stored energies. On the basis of the information now available, the stored energy of deformation for various orientations can be written in the following sequence:
E{llO}(OOD

of high stored energy such as those of the {Ill} fibre, certain nuclei grow while others do not or grow more slowly. The more successful nuclei appear to be those that have orientation relations with respect to the matrix represented by preferred (110) axis rotations. This leads to the replacement of the RD (rolling or deformation) fibre by the {Ill} fibre and related (e.g. {554} (225) orientations." Using back reflection Kossel patterns to determine the crystallographic orientations of the new grains appearing in a 200/0 recrystallised steel sheet, Benoit et al.136 found that these are mainly of {Ill} orientation. The predominance of {Ill} oriented grains during the early stages of recrystallisation has been confirmed by several other workers.137-139 The nucleation rate can be plotted against annealing time on the basis of stored energy considerations, as illustrated in Fig. 58 for the principal orientations. According to this view, the {110} and {Ill} texture components should nucleate first and have the longest times available for growth. The {110} component is unlikely to become strong, however, because of its relatively low density in the cold worked matrix. The {001} will be the least favoured orientation to nucleate, and even if present at the beginning of recrystallisation, will decrease in volume fraction because of the more rapid selective growth of favourable orientations, such as the {Ill}. Nucleation sites for recrystallised grains and their growth During the recrystallisation of cold rolled steel, {Ill} oriented grains appear to nucleate adjacent to grain boundaries.l'P l'" This observation is consistent with the view that the {Ill} fibre forms by selective growth and explains why a sharper {Ill} fibre texture is achieved when the hot band grain size is finer." The Goss {ll0}(001) oriented grains nucleate preferentially along deformation bands.l'" As the density of deformation bands increases with the concentration of dissolved carbon present before cold rolling, high solute C levels lead to a decrease in the intensity of the {Ill} texture, with a simultaneous increase in the intensity of the {110} component in the final annealing texture. There is some evidence that the rotated cube {001}(110) oriented recrystallised grains also nucleate at deformation bands.P" The rates of recovery of the {Ill} and {112} orientations are much faster than that of the {001}.145,146 Therefore, when sufficient recovery takes place before recrystallisation and grain growth, the stored energy differences between regions of different orientations decreases, thereby reducing the driving force for SIBM. As a result, the {001} component will be weakened and the {Ill} strengthened in the annealing texture. In general, the texture of a cold worked material is expected to determine its annealing texture. In the past 60 years, a large volume of research has been carried out on the mechanism of formation of recrystallisation textures. In this connection, two theories, namely, the oriented nucleation147,148 and oriented growth 149-151 models have been proposed. In the former, it is suggested that only specific orientations of recrystallisation nuclei are formed in the cold
International Materials Reviews 1994 Vol. 39 No.4

< E{OOl}<l10) < E{1l2}(uvw> < E{1 < E{llO}(110)

ll} Cuvw)

When nucleation occurs by strain induced boundary migration (SIBM), the Goss and the rotated cube orientations grow into regions of high energy, such as the {112} or {Ill}, or into the matrix as a whole. This is a form of oriented nucleation, in which the other orientations do not succeed in developing viable nuclei. When the nucleation of recrystallised grains takes place around coarse second phase particles, such as cementite, the nuclei will have no orientation bias and, therefore, a nearly random texture will evolve from this volume fraction of the material on recrystallisation. Finally, when nucleation occurs by coalescence or subgrain growth, essentially in grains

164

Ray et al.

Textures in low and extra low carbon steels

<Pl,deg
(001)( 11 0 I0 .O--,-..,..----.---r--r-~-_r_____r___.____r_____r__r__ ... (001)[110] 90
.30 60
PHI2

= 450

MAX: 4.3

2.6
(114)[110] (113)(110) --------:::. (113){332] (112)(1 3.5 (111)[121) (111)(0111
LEVELS: 2.2 2.6 3.2 3.8 2.4 2.9 3.5 4.1

(112)[110) (445)[110) (111)(110]

i I)

tfi (223)[11 OJ

(111)[112) (554)[225] (332)[113J

Published by Maney Publishing (c) IOM Communications Ltd

(a)

(110) [11 OJ ......----"----'---'_--I..--JI--.L.-.L-....L..--L..-...L......l-L...L..-...L-L..-~ 90

(110) [00 II

<P2,deg
(001)[110] (001)[100)
0

___ ~--r-----rr----"'-

30

~---,rw---'---rT--____

60

90

(001) [01 OJ

PH 11 = MAX: 4.3

(013)(100)
30
CIJ
"'0

LEVELS: 2.2 2.4 2.6 3.2 3.8 2.3 2.5 2.9 3.5 4.1

tfi

(011)(100]

(101)[010)

60

(031)(1001

(301)1010)

(b)

(010)11001
90

(110)(1101
a 2 = 45 section; b , = 0 section Full constraint Taylor factor maps for bee materials (after Refs. 41, 44)

57

worked matrix, so that the annealing texture is characterised by the orientations of these nuclei. The latter theory advocates that recrystallisation nuclei are formed in a random fashion, orientation-wise, in the cold worked matrix; however, because of the orientation dependence of grain boundary mobility, only the nuclei possessing the highest growth rates grow rapidly. The annealing texture, in that case, is determined essentially by the orientations of the fastest growing nuclei. A considerable amount of work carried out by Lucke'Y and his associates has demonstrated the extent and effect of the orientation dependence of grain boundary mobility and therefore of growth selection during recrystallisation. From their bicrystal experiments, Ibe and Lucke"? found that an
International Materials Reviews 1994 Vol. 39 No.4

orientation relationship of 27 110) is applicable to high growth rates, which allows suitably oriented ND fibre grains to grow into the adjoining RD fibre region. An important consequence of selective growth in low carbon steels is an orientation dependence of the grain size in recrystallised materials. This is illustrated in Fig. 59, where the average {Ill} grain size is found to be larger than the mean grain size in a recrystallised low carbon steel.P? This feature of the recrystallised structure, where the {Ill} grains have a clear size advantage over the others, explains why rm increases during grain growth in different grades of low carbon steel. 153 The beneficial effect of grain growth on r m for three types of low carbon steels is shown in Fig. 60.

<

Ray et al.

Textures in low and extra low carbon steels


3.0

165

r-r---r---r----r----I
I I

2.5 -

[111]

Q)

~
z
o

::J ~ 2.0
>-E

d :>
ANNEALING TIME --+

1.5 -

1.0 10 20 GRAI N SIZE, 30

40

58

Schematic representation of nucleation rate v. annealing time for recrystallised grains of different low index orientations (after Ref. 3)

urn

60

Published by Maney Publishing (c) IOM Communications Ltd

Dependence of rm on grain size for three different steels cold rolled and annealed (after Ref. 99)

LUcke and his associates'Y have suggested that the ?ccurrence of selective growth can explain the rapid Increase, during annealing, in the intensity of the {111}(112) component and the simultaneous depletion of the {112}(110) in the deformed matrix. The {111}(112) is related to the {112}(110) by a 32 (110) relationship, which is not far away from the . ~7 (110) rel~tionship which confers high mobility to bee grain boundaries/'? This orientation relation allows growth of the {111}(112) component of t~e ND fibre at the expense of the {112}(110) portion of the RD fibre. Schlippenbach and Lucke':" suggested that a similar mechanism plays a role in the formation of the Goss component, {110}(001), of the annealing texture. The fact that this orientation has a 3?0 (110) r~lationship with the {111}(112) and their observatIon that, in the recrystallisation texture, one of these components is strong when the other is weak, has led to the conclusion that one source of the Goss component is the {111}(112). Lindh et al.154 have recently studied texture evolution during the recrystallisation of an ELC IF steel containing 0'0030/0C, 0080/0Mn, 0041 0/0 0'004%N, AI, and 0082 % Ti. They could not find any evidence for the (110) rotational misorientations that are supposed to confer exceptionally high mobility to the grain boundaries, at least in their material and in the

context of their analysis. This constitutes evidence against the oriented growth mechanism for the formation of recrystallisation textures. Clearly, this aspect of recrystallisation theory requires further and more rigorous investigation. . ~n th~ basis of the knowledge currently available, It IS possible to summarise the steps that lead to the formation of annealing textures in low and extra low carbon steels. This is also shown schematically in Fig. 61.145 Step 1: When recrystallisation starts because of the high stored energy, and therefore the high driving for~e for recovery, the first nuclei appear in the {Ill} regions and assume this orientation. Step 2: These nuclei initially grow at the expense of deformed regions of the same orientation. Step 3: Once they have reached a critical size, they grow at the expense of adjoining deformed zones to which they are related by (110) rotations. Step 4: Nuclei of lower stored energy orientations, such as .the. {11.2} and. {100}, start appearing. RecrystalhsatIon IS essentIally finished when the nu~lei, which have now become grains, occupy the entire volume of the material. Effect of elements in solid solution Elem~nts in soli~ solution in low carbon steel produce solution hardening, The increase in yield stress as a

25 o
N

w 20 ~

15

~ 10

5
11001

<

UYW UYW UYW

10

20 urn

30

/2111 11111

<
<

> 1111111 > 0i~{~


>
111111111111/11

GRAIN SIZE (0),

59

Grain size distribution for all grains and for those with {111} orientations in a just recrystallised low carbon steel sheet (after Ref. 139)

61

Schematic diagram illustrating steps leading to formation of recrystallisation texture in low carbon steel sheets (see text) (after Ref. 145) Materials Reviews 1994 Vol. 39 No.4

International

166

Ray et al.

Textures in low and extra low carbon steels


Z --------------------, acceleration of recrystallisation

C'oI

300

C and N

o en
~Si

z
w

en en
en
-J W

~ 225

:J -J

~ 150

o
75

a: u w a: '#.
o
I!)

~ en >-

single phase alloy retardation of recrystallisation

>= z
w
CJ

o u..

a:
w

:::c

~
L- __ ..L.__ -.L. __
-...ll-- __ .L- __

u -75

.....l--.J

0.5

2.0

2.5

~
INTERPARTICLE SPACING ~

62

Effect of alloying elements on yield stress of low carbon steel sheets (after F.B. Pickering, as quoted in Ref. 145)

63

Effect of interparticle spacing on time for 500/0 recrystallisation (after Ref. 156)

Published by Maney Publishing (c) IOM Communications Ltd

function of solute content, for the most common alloying elements, is illustrated in Fig. 62. It is tempting to correlate the solution hardening associated with individual elements with effects they may have on the cold rolling texture. Numerous investigations have shown, however, that elements in solid solution have no visible effect on the cold rolling texture,54,145 though there may be discernible changes in the cold rolled microstructure. 54 From a careful study of the results of investigations at IRSID, Meyzaud and Parnierel'" have come to the conclusion that, with the exception of some special situations, elements in solid solution affect the annealing textures of low and extra low carbon steels only marginally. These special situations are: 1. When dissolved carbon is present before cold rolling or during annealing, the resulting {111} recrystallisation texture is significantly weakened.Pt'" Similar observations apply to the presence of N, Mn, and P, in solution.71-73,81,117 2. Niobium, Ti, and to some extent Si, in solid solution appear to improve the {111} texture.28,29,98-108,120 Effect of precipitate particles Precipitate particles of various kinds, such as AIN, TiC, Fe3C, MnS, etc., are formed during the processing of low carbon steels. Some of these have been found to have a profound effect on the recrystallisation texture. Precipitates formed before cold rolling Interstitial free steels always contain certain amounts of second phase particles after hot rolling, such as TiC, TiN, Ti(CN), Ti4C2S2, Nb(C,N), NbC, AIN, and MnS. The precipitates present depend on the particular composition of the steel considered. Their influence on subsequent recrystallisation and the resulting texture is mainly a function of their morphology that is of their size, volume fraction, and mean spaci~g. Koster':" and Doherty and Martin1S7 have shown that recrystallisation can either be delayed or accelerated by the presence of small or large particles, respectively (see Fig. 63).
International Materials Reviews 1994 Vol. 39 No.4

When a few large particles are present in the matrix, a very heterogeneous deformation structure is produced, though the cold rolling texture is not significantly modified.?" The recrystallisation rate is accelerated in this case. Nuclei form preferentially around the particles without any orientation bias, giving rise to a random recrystallisation texture. This occurs when steel contains coarse cementite particles, which may have precipitated after high temperature coiling. This deleterious effect of coarse cementite particles is more than counterbalanced when, because of the much higher heating rate prevailing during continuous annealing, the dissolution rate of cementite is decreased, thus releasing less carbon into the matrix. When a large number of small particles is present in the hot band matrix (as in the case of Nb and/or Ti containing IF steels), the cold rolling texture again does not seem to be modified.P" It has also been shown28,29 that, for ordinary Nb stabilised IF steels, the NbC precipitates present in the hot rolled material do not exert any measurable influence on the develop' . ment of the recrysta 11isation texture. H u t chimson 158 has also suggested that second phase particles such as NbC or Ti(C,N) are largely irrelevant with respect to the formation of the {111} annealing texture. Precipitates formed during annealing When precipitation occurs during a recrystallisation anneal, there is a possible interaction between the two phenomena.156,159,160 This is illustrated schematically in Fig. 64. The diagram shows that for any temperature T, (i) if T> 1;., no precipitation takes place and recrystallisation occurs with the alloying elements in solid solution (ii) if 1;. > T> 12, recrystallisation occurs first, followed by precipitation. Interaction is possible if precipitation starts before the end of recrystallisation (iii) if 12 > T> 7;, precipitation will retard the onset of recrystallisation (iv) if T < 7;, precipitation is c?ml?leted befo~e the beginning of recrystallisation, In this case, the precipitates must coarsen before recrystallisation can take place.

Ray et al.

Textures in low and extra low carbon steels

167

i E
::>
w w a:

'? 10

E E

Annealing rate 12Ks-1

C3 8 :J

'co
C

.36Kh-1

~ a: T3

~ 0 ~

6
4

Q.

.w 64

~ a: z

LOGTIME~ Schematic diagram illustrating interaction between recrystallisation (B) and precipitation (C) (after Ref. 159)

::>
0 10 20 30 40 50 60 RECRYSTALLISATION FRACTION, % 70

w ....J U

65

Published by Maney Publishing (c) IOM Communications Ltd

Influence of heating rate during annealing on nucleation rate of grains in a 700/0 cold rolled low C steel sheet (after Ref. 167)

Classic examples of the phenomena described above are provided by the interaction between AIN particles (and clusters) and recrystallisation during the annealing of low carbon AI-killed steels. The best results are obtained when annealing is carried out at relatively low temperatures (below 540C). Recrystallisation is delayed until the precipitation of AIN is complete and the resulting annealing texture then contains strong {Ill} components. During the industrial processing of AI-killed low carbon steels, the sequence of precipitation and recrystallisation is normally controlled by the heating rate during annealing. Low heating rates lead to the precipitation of AIN during recovery, resulting in a strong {Ill} texture after recrystallisation. The heating rates which lead to the highest rm values, also known as the peak heating rates (PHRs), can be calculated from the following relationship.'?' 10g(PHR)

= 183 +27 log ([Al] [N] [Mn]/RcR)


. . . . . . . . (3)

The conventional batch annealing of AI-killed steels provides strong {Ill} textures and correspondingly high rm values. In order to produce comparable properties by continuous annealing, it is necessary to modify some of the processing parameters. One of the changes requires increasing the grain size, because excessively fine microstructures are detrimental to formability. This is done by decreasing the nucleation rate, which involves lowering the recrystallisation temperature. The former can be achieved by: (a) decreasing the carbon and alloying and impurity levels in the steel; and (b) increasing the stored energy of deformation, either by resorting to higher cold rolling reductions, or through adjustment of the compositional and hot rolling parameters to obtain the desired distribution of second phase particles (TiC, Nb(C,N), Fe3C, MnS, ... ) in the hot band.':" Texture control in low carbon steels A critical assessment of the effects of various compositional and processing parameters on the development of annealing textures in low and extra low carbon steels has been presented above. An overall view of the textures formed during cold rolling and annealing, and of the factors that influence their intensities, is depicted schematically in Fig. 66. Unlike the cold rolling texture, which is affected mainly by the hot band texture and amount of cold reduction, the annealing texture is controlled by a host of parameters pertaining to steel chemistry and the entire processing history. The situation with regard to the formation of cold rolling textures is, therefore, much simpler and can, in fact, be satisfactorily modelled by the methods of crystal plasticity.P':" By contrast, simulation of the development of annealing textures presents a much greater challenge. Recently, the formation of recrystallisation textures in IF steels has been modelled in terms of a selective growth theory in which those {Ill} nuclei that possess favourable (110) axis rotation relationships with respect to the deformed matrix undergo preferential growth. By adopting a suitable (110) axis variant
International Materials Reviews 1994 Vol. 39 No.4

where PHR is the heating rate in K h -1 corresponding to the peak in rm value, [AI], [N], and [Mn] are the solute concentrations in weight per cent, and RCR the percentage cold rolling reduction. Similar interactions between precipitation and recrystallisation have also been observed during the annealing of a low carbon steel containing 12%Cu (Refs. 162-164) and also in IF steels containing different amounts of Ti (Refs. 165, 166). Effect of heating rate Lebrun et al.146 have carried out a detailed analysis of the influence of heating rate on the recrystallisation of low carbon steel sheets. They observed that the main consequence of an increase in heating rate is to raise the temperature of primary recrystallisation. Microstructural observations have indicated that this is due to the decreased time available for recovery. To compensate, recrystallisation takes place at a higher temperature and, since higher recrystallisation temperatures give rise to higher nucleation rates.l"? the resulting grain size will be much finer (see Fig. 65).

168

Ray et a/.

Textures in low and extra low carbon steels

'Y

Rolling and recrystallisation

texturel

Y -to- a transformation

I Transformed
Sharpened by higher cold reductions and sharper hot band textures

hot band texturel Cold rolling


Affected a little by steel chemistry but not by morphology of precipitates
J

Cold rolling texture

I
RD fibre, NO fibre,

< 110> !lRD


>
or

< 111 > 1.1 O N

Maxima at {OO 1 } < 11 0 at {112} < 110 >

I
Published by Maney Publishing (c) IOM Communications Ltd

I
Annealing Annealing texture

I
1
RD fibre weakens,
significant decline in {112} < 110> intensity

ND fibre, {111}
sharpens

< uvw >

appreciably

I
In {111} sharpened by fine hot band batch annealing, Texture forms by oriented nucleation of {001} < 110> and {111} is sharpened by low coiling temperatures grain size; high cold reductions; low concentrations of C, N, P, and {11 O} < 001 > at shear bands or by SIBM, and of {111}<uvw> grain boundaries. The latter is sharpened by selective growth into deformed matrix grains related by < 110> rotations. at

Mn; high annealing temperatures and long holding times ( to promote grain growth)

and slow heating rates. In continuous annealing,

{1 1 1} is sharpened by high coiling temperatures and high heating rates.

66

Schematic representation of texture formation influence them

during cold rolling and annealing and of factors that

selection rule and employing it in conjunction with a nucleus availability factor, it has been shown that recrystallisation textures can be successfully predicted from experimental rolling textures." Bunge and Kohler168 have also suggested a general model for recrystallisation texture, valid for both bee and fcc materials. They have assumed nucleation at randomly distributed sites and simultaneous growth of these nuclei into all the components of the deformation texture. The average growth rate was considered to be a 'compromise' of the local growth rates; the latter are given by an orientation dependent driving force and the local mobility, which depends in turn on the orientation difference between growing grains and the
International Materials Reviews 1994 Vol. 39 No.4

matrix. However, further work is necessary before an acceptable model for the formation of recrystallisation textures in steels can emerge. Over the years, great advances have been made in elucidating the effects of different variables on the formation of the {111} texture. For example, the influence of alloying elements such as C is now well documented. In addition, the way the hot band texture and grain size influence the annealing texture is reasonably well understood. It has also been possible to identify, with reasonable accuracy, the nucleation sites of the different annealing texture components in the deformed matrix. However, in order to achieve a more quantitative understanding of the

Ray et al.

Textures in low and extra low carbon steels

169

Published by Maney Publishing (c) IOM Communications Ltd

recrystallisation process, further work is necessary on the following lines: 1. The interactions between alloying elements and the way they affect the annealing texture should be investigated in detail. This will require working with alloys of strictly controlled chemical compositions. 2. The process of recrystallisation itself, especially the nature and mobility of grain boundaries, is still not fully understood. Also of importance are the precise effects of second phase particles and the rules of variant selection during selective growth. Experiments using bicrystals deformed in channel dies and the very precise determination of TTT diagrams for recrystallisation and precipitation will be needed to further understanding of these phenomena. 3. The ideal orientation relationship of 27 (110) (and of other CSL or coincidence site lattice relationships) suggested for the rapid growth rate of suitably oriented nuclei into the deformed matrix should be looked into carefully. The different manifestations of these relationships and their dependence on alloy chemistry, amount of cold deformation, etc., should also be properly investigated. 4. Perhaps the most promising line of attack will be to undertake very detailed microstructural studies of both the cold rolled and annealed states and to correlate these observations with crystallographic data. This will pave the way for greater understanding of the overall phenomenon and allow accurate modelling of the process. The sound scientific models obtained in this way will lead to the more effective industrial control and optimisation of annealing textures in low and extra low carbon steels.

Summary and conclusions


In this review, the extensive literature on the development of cold rolling and annealing textures in low and extra low carbon steels has been summarised. Texture is an important property as it induces plastic anisotropy, which can have both beneficial and detrimental effects on the formability. In the preceding sections, the effects of different compositional and processing variables on texture formation have been evaluated critically and the complex interactions between the various parameters have been looked into carefully. Finally, efforts were directed towards increasing the understanding of the basic mechanisms of texture formation in order to provide a scientific basis for industrial texture control. The major conclusions of this work are enumerated below. 1. The cold rolling texture of low carbon steels is mainly composed of a nearly perfect ND fibre and a peak component situated at or near {001}(110) or at or near {112}(110) on the RD partial fibre axis, depending on the type of steel and the amount of cold rolling reduction. This texture is affected most significantly by the hot band texture and not particularly by such metallurgical parameters as steel chemistry or the morphology of precipitates. 2. The deformation textures predicted theoretically by means of the 'relaxed constraint' versions of the Taylor model of crystal plasticity are in reasonably good agreement with experimental cold rolling textures. Thus, the observed textures can be readily

accounted for by the geometric features of crystallographic glide. 3. When heavily cold rolled material is recrystallisation annealed, the ND fibre is strengthened and the orientation density of the RD fibre, particularly that of the {112}<110) component, decreases to some degree, ranging between modest and large depending on the type of steel, amount of cold reduction, and the location in the sheet where texture is evaluated. The annealing texture, especially the intensity of the {Ill} component, depends critically on the hot band texture and grain size, certain processing variables, such as the amount of cold reduction, and the alloy chemistry. 4. During annealing, {111}<uvw) grains nucleate at grain boundaries, and those that have favourable (110) axis rotation relationships with respect to the deformed matrix, particularly with respect to the RD fibre, grow rapidly. Oriented nucleation followed by strain induced boundary migration lead to the formation and growth of the {001}(110) oriented grains. Oriented nucleation is also responsible for the appearance of the {110}(001) grains, which form preferentially in deformation bands. Textural changes during recrystallisation annealing are also affected by the orientation dependence of the stored energy of cold working, which increases in the order E{110}<OOD < E{OOl}<110> < E{112}<uvw> < E{111} <UVW> <E{110}<110>' 5. For conventional steels processed by batch annealing, the optimum amount of cold reduction is around 70%. This leads to the most desirable annealing textures and the correspondingly highest r rn values. However, in the Nb or Ti stabilised interstitial free (IF) steels, the optimum reduction can be increased to 900/0.In all these steels, finer hot band grain sizes prod uce sharper {Ill} textures after annealing because they are less susceptible to shear band formation during cold rolling. 6. Low coiling temperatures and slow heating rates during annealing impart high r rn values to Al-killed steels. The r rn values of Nb or Ti stabilised steels do not appear to be particularly sensitive to either coiling temperature or heating rate. 7- Grain growth after recrystallisation generally leads to sharper {Ill} fibres and increased rrn values. This is why longer annealing cycles and higher annealing temperatures are beneficial. In the case of box annealing, the practical upper limit of annealing temperature is ~ 720C. The higher heating and cooling rates inherent in the continuous annealing process allow temperatures above 720C to be used. Thus coils can be annealed in the intercritical y + a range, which can lead to intensification of the {Ill} components. 8. Carbon in solution and/or in the form of carbides is the single most deleterious element in that it retards the development of sharp {Ill} annealing textures. Dissolved N, P, and Mn, have effects similar to that of C. Niobium, Ti, and Si in solid solution enhance the intensity of the {Ill} fibre texture. 9. The interaction between substitutional and interstitial solutes may give rise to the formation of complexes or dipoles of the type Mn-C, Mn-N, and AI-N. Such complexes can pin dislocations at elevated temperatures, thus delaying recovery; they
International Materials Reviews 1994 Vol. 39 No.4

170

Ray et al.

Textures in low and extra low carbon steels 20. u. von SCHLIPPENBACH, F. EMREN, and K. LUCKE: Acta Metall., 1986, 34, (7), 1289. 21. K. LUCKE and M. HOLSCHER: Textures Microstruct., 1991, 14-18, 585. 22. M. SHIMIZU, K. MATSUDA, Y. SADAMURA, N. TAKAHASHI, and M. KAWAHARADA: French Pat. 1511529, Jan. 1968: this corresponds to Br. Pat. 1176863, Jan. 1970 and US Pat. 3522110, July 1970. 23. J. A. ELIAS and R. E. HOOK: US Pat. 3 761 324, Sept. 1973. 24. J. A. ELIAS and R. E. HOOK: US Pat. 3 765 874, Oct. 1973. 25. T. OBARA, S. SATOH, M. NISHIDA, and T. IRIE: Scand. J. Metall., 1984, 13, 201. 26. A. J. HECKLER and w. G. GRANZOW: Metall. Trans., 1970, 1, 2089. 27. R. E. HOOK: Unpublished results, Armco, Inc., Research and Technology, Middletown, OH. 28. R. E. HOOK, A. J. HECKLER, and J. A. ELIAS: Metall. Trans., 1975, 6A, 1683. 29. R. E. HOOK: in 'Metallurgy of vacuum-degassed steel products', (ed. R. Pradhan), 263; 1990, Warrendale, PA, Metallurgical Society of AIME. 30. R. E. HOOK: Metall. Trans., 1993, 24A, 2009. 31. R. J. JESSEMAN:Unpublished results, AK Steel Corp., Research Center, Middletown, OH. 32. s. L. LOPATA and E. B. KULA: Trans. AIME, 1962, 224, 865. 33. S. LEBER: Rev. Sci. Instrum., 1965, 36, 1747. 34. R. J. HAZEL and R. c. HUDD: 'A laboratory investigation of the processing behaviour and properties of niobium bearing EDD steels'. Report SM/734/A, British Steel Corp., South Wales Group, 1970. 35. H. INAGAKI: Z. Metallkd., 1984, 75, 510. 36. H. INAGAKI: Z. Metallkd., 1988, 79, 716. 37. M. P. BUTRON-GUILLEN, J. J. JONAS, and R. K. RAY: Acta Metall. Mater., 1994, 42, in press. 38. R. E. HOOK and H. NYO: Metall. Trans., 1975, 6A, 1443. 39. w. B. HUTCHINSON, K.-I. NILSSON, and J. lllRSCH: in 'Metallurgy of vacuum-degassed steel products', (ed. R. Pradhan), 109; 1990, Warrendale, PA, Metallurgical Society of AIME. 40. O. KWON, G. KIM, and R. W. CHANG: in 'Metallurgy of vacuumdegassed steel products', (ed. R. Pradhan), 215; 1990, Warrendale, PA, Metallurgical Society of AIME. 41. J. J. JONAS and T. URABE: in Proc. Int. Forum on 'Physical metallurgy of IF steels', Tokyo, 1994, The Iron and Steel Institute of Japan, 77-94. 42. F. H. SAMUEL, S. YUE, J. J. JONAS, and B. A. ZBINDEN: in 'Metallurgy of vacuum-degassed steel products', (ed. R. Pradhan), 395; 1990,Warrendale, PA, Metallurgical Society of AIME. 43. A. NAJAFI-ZADEH, S. YUE, and J. J. JONAS: ISIJ Int., 1992,32,213. 44. T. URABE and J. J. JONAS: ISIJ Int., 1994, 34, 435-442. 45. L. S. TOTH, J. J. JONAS, D. DANIEL, and R. K. RAY: Metall. Trans., 1990, 21A, 2985. 46. I. L DILLAMORE and H. KATOH: Met. Sci., 1974,8,21. 47. J. L. RAPHANEL and P. VAN HOUTTE: Acta Metall., 1985,33, 1481. 48. M. ARMINJON: Acta Metall., 1987,35,615. 49. P. GILORMINI: Acta Metall., 1989, 37, 2093. 50. w. B. HUTCHINSON: in Proc. 10th Int. Conf. on 'Textures of materials', 1917-1928; 1994, Aedersmannsdorf, Switzerland, Trans Tech Publications. 51. c. DASARATHY: 'A review of recent observations on the deformed, recovered and recrystallized states in iron and low carbon ferrite steels', Report SM/668/ A, British Steel Corp., 1973. 52. M. FUKUDA: Tetsu-to-Hagane (J. Iron Steel Inst. Jpn), 1967, 53,559. 53. S. NOMURA, T. YUTORI, and T. FUKUTSUKA: Tetsu-to-Hagane (J. Iron Steel Inst. Jpn), 1975, 61, 3092. 54. J. J. LAVIGNE, T. SUZUKI, and H. ABE: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 2, 749; 1981, Tokyo, The Iron and Steel Institute of Japan. 55. H. ABE, T. SUZUKI, and K. TAKAGI: Trans. Iron Steel Inst. Jpn, 1981, 21, 100. 56. T. SENUMA, H. YADA, R. SHIMIZU, and J. HARASE: Acta Metall. Mater., 1990, 38, 2673. 57. H. KUBOTERA, K. NAKAOKA, K. ANAKI, K. WATANABE, and
K. IWASE: Tetsu-to-Hagane

also seem to retard the nucleation and growth of the {111} fibre. 10. The interaction between precipitation and recrystallisation can be used to advantage to control the annealing texture. In Al-killed low carbon steels, the precipitation of AIN occurs before the onset of recrystallisation, and this produces a strong {111} annealing texture. The precipitation, before cold rolling, of fine second phase particles such as TiC and Nb( C,N) in Ti and Nb IF steels is largely irrelevant to the development of the {111} texture. 11. The extra low carbon steels (C + N ~ 0006%) are remarkably tolerant of many compositional and processing variables. The very low level of carbon in these steels can in itself enable excellent deep drawabilities to be attained.

Acknowledgments
Published by Maney Publishing (c) IOM Communications Ltd The authors are grateful to the Natural Sciences and Engineering Research Council of Canada (NSERC), the Canadian Steel Industry Research Association (CSIRA), and the Ministry of Education of Quebec (FCAR Program) for financial support. They would like to thank Drs J. Savoie and M. P. Button-Guillen of the Department of Metallurgical Engineering, McGill University for numerous helpful discussions. Thanks are also due to T. Urabe, research student of the same department, for kindly making available some of his unpublished results. Appreciation is also extended to R. P. Jesseman, Research and Technology, AK Steel Corp., for making available some of his unpublished results and to AK Steel Corp., for permission to use the results. The extremely able secretarial help of Ms Lorraine Mello and Priti Wanjara is also gratefully acknowledged.

References
in 'Mechanical working and steel processing IV,' (ed. D. A. Edgecombe), 3; 1965, New York, American Institute of Mining, Metallurgical and Petroleum Engineers. 2. s. MISHRA and c. DARMANN: Int. Met. Rev., 1982, 27, (6), 307. 3. w. B. HUTCHINSON: Int. Met. Rev., 1984, 29, (1), 25. 4. R. K. RAY and J. J. JONAS: Int. Mater. Rev., 1990,35, (1), 1. 5. L. G. SCHULZ: J. Appl. Phys., 1949, 20, 1030. 6. B. F. DECKER, E. T. ASP, and D. HARKER: J. Appl. Phys., 1948, 19, 388. 7. B. D. CULLITY: 'Elements of X-ray diffraction'; 1978, Reading, MA, Addison-Wesley. 8. M. HATHERLY and w. B. HUTCHINSON: 'An introduction to textures in metals'; 1979, London, The Institution of Metallurgists. 9. K. LUCKE: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 1, 14; 1981, Tokyo, The Iron and Steel Institute of Japan. 10. H. J. BUNGE: Z. Metallkd., 1965, 56, 872. 11. R. J. ROE: J. Appl. Phys., 1965, 36, 2024, 2069. 12. H. J. BUNGE: 'Mathematische Methoden der Texturanalyse'; 1969, Berlin, Akademie Verlag. 13. H. J. BUNGE: 'Texture analysis in materials science'; 1982, London, Butterworths. 14. J. HANSEN, J. POSPIECH, and K. LUCKE: 'Tables for texture analysis of cubic crystals'; 1978, Berlin, Springer-Verlag. 15. D. DANIEL and J. J. JONAS: Metall. Trans., 1990, 21A, 331. 16. R. K. RAY, J. J. JONAS, M. P. BUTRON-GUILLEN, and J. SAVOIE: ISIJ Int., 1994, 34, in press. 17. D. DANIEL, K. SAKATA, and J. J. JONAS: ISIJ Int., 1991, 31, 696. 18. D. DANIEL and J. J. JONAS: Textures Microstruct., 1991, 14-18, 1165. 19. R. HILL: Proc. R. Soc., 1952, A65, 349.
J. F. HELD:

1.

(J. Iron

Steel

Inst. Jpn),

1976, Cryst.

62,846. 58.
K. MATSUDO, T. SHIMOMURA,

and o.

NOZOE: Texture

Solids, 1978, 3, 53.

International

Materials Reviews

1994

Vol. 39

No.4

Ray et al. 59. w. B. HUTCHINSON and K. USHIODA: Scand. J. Metall., 1984, 3,269. 60. D. T. GAWNE and G. T. HIGGINS: in 'Textures in research and practice', 319; 1969, Berlin, Springer-Verlag. 61. K. USHIODA, J. AGREN, and w. B. HUTCHINSON: in Proc. 7th Int. Conf. on 'Textures of materials', 301; 1984, Noordwijkerhout, Netherlands Society for Materials Science. 62. K. TODA, H. GONDOH, H. TAKEUCHI, M. ABE, N. UEHARA, and K. KOMIYA: Trans. Iron Steel Inst. Jpn, 1975, 15, 305. 63. K. USHIODA, W. B. HUTCHINSON, J. AGREN, and u. von SCHLIPPENBACH: Mater. Sci. Technol., 1986, 2, 807. 64. w. B. HUTCHINSON and K. USHIODA: in Proc. 7th Int. Conf. on 'Textures of materials', 409; 1984, Noordwijkerhout, Netherlands Society for Materials Science. 65. I. L. DILLAMORE, C. J. E. SMITH, and T. W. WATSON: Met. Sci. J., 1967, 1,49. 66. K. USHIODA and M. ABE: Tetsu-to-Hagane (J. Iron Steel Inst. Jpn), 1984, 70, 96. 67. G. IBE and K. LUCKE: Arch. Eisenhiittenwes.; 1968, 39, 693. 68. S. HASHIMOTO, T. KASHIMA, and T. INOUE: Textures Microstruct., 1991, 14-18, 841. 69. S. HASHIMOTO, T. YAKUSHIJI, T. KASHIMA, and K. HOSOMI: in Proc. 8th Int. Conf. on 'Textures of materials', 673; 1988, Warrendale, PA, Metallurgical Society of AIME. 70. Y. HOSOYA, T. SUZUKI, and A. NISHIMOTO: in 'Metallurgy of vacuum-degassed steel products', (ed. R. Pradhan), 291; 1990, Warrendale, PA, Metallurgical Society of AIME. 71. M. TAKAHASHI and A. OKAMOTO: in Proc. 5th Int. Conf. on 'Textures of materials', Vol. 2, 265; 1978, Berlin, SpringerVerlag. 72. M. TAKAHASHI, A. OKAMOTO, S. INO, and T. NAKATA: Trans. Iron Steel Inst. Jpn, 1979, 19, 144. 73. M. TAKAHASHI and A. OKAMOTO: Trans. Iron Steel Inst. Jpn, 1979, 19, 391. 74. A. OKAMOTO and N. MIZUI: in Proc. 7th Int. Conf. on 'Textures of materials', 427; 1984, Noordwijkerhout, Netherlands Society for Materials Science. 75. K. TAGASHIRA, W. B. HUTCHINSON, and I. L. DILLAMORE: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 2, 778; 1981, Tokyo, The Iron and Steel Institute of Japan. 76. J. E. ENRIETTO: Trans. AIME, 1962, 224, 43, 1119. 77. w. C. LESLIE, R. L. RICKETT, C. L. DOTSON, and c. S. WATSON: Trans. ASM, 1954,46,1470 .. 78. P. N. RICHARDS: J. Aust. Inst. Met., 1967, 12, 2. 79. P. R. MOULD: in 'Metallurgy of continuous-annealed sheet steel', (ed. B. L. Bramfitt and P. L. Mangonon), 3; 1982, Warrendale, PA, Metallurgical Society of AIME. 80. N. TAKAHASHI, M. SHIBATA, Y. FURUNO, H. HAYAKAWA, K. KAKUTA, and K. YAMAMOTO: in 'Metallurgy of continuousannealed sheet steel', (ed. B. L. Bramfitt and P. L. Mangonon), 133; 1982, Warrendale, PA, Metallurgical Society of AIME. 81. H. HU and s. R. GOODMAN: Metall. Trans., 1970, 1, 3057. 82. I. F. HUGHES and E. W. PAGE: Metall. Trans., 1971, 2, 2067. 83. N. OHASHI, M. KONISHI, and Y. ARIMA: Kawasaki Steel Tech. Rep., 1973,5, 164. 84. N. TAKAHASHI, M. ABE, O. AKISUE, and H. KATOH: in 'Metallurgy of continuous-annealed sheet steel', (ed. B. L. Bramfitt and P. L. Mangonon), 51; 1982, Warrendale, PA, Metallurgical Society of AIME. 85. H. KOBAYASHI, T. SHIMOMURA, and K. MATSUDO: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 2, 729; 1981, Tokyo, The Iron and Steel Institute of Japan. 86. H. HU: Metall. Trans; 1977, 8A, 1567. 87. H. HU: in Proc. 5th Int. Conf. on 'Textures of materials', Vol. 2, 3; 1978, Berlin, Springer-Verlag. 88. w. B. HUTCHINSON: Met. Sci., 1974, 8, 185. 89. K. OSAWA, K. MATSUDO, K. KURIHARA, and T. SUZUKI: Tetsuto-Hagane (J. Iron Steel Inst. Jpn), Mar. 1984, 70, S552. 90. W. KOCH and H. KELLER: Arch. Eisenhiittenwes., 1964, 35, 1173. 91. P. PARNIERE: Report PA3 2340, IRSID, 1980. 92. T. SUZUKI and H. ABE: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 2, 797; 1981, Tokyo, The Iron and Steel Institute of Japan. 93. H. ABE, T. SUZUKI, and J. J. LAVIGNE: Trans. Iron Steel Inst. Jpn, 1981, 21, 332. 94. J. FRIEDEL: 'Dislocations', 407; 1964, Oxford, Pergamon Press. 95. L. S. DARKEN and R. W. GURRY: 'Physical chemistry of metals', Chap. 16; 1953, New York, McGraw-Hill. 96. T. NISHIZAWA: Bull. Jpn Inst. Met., 1973, 12, 401.

Textures in low and extra low carbon steels


B. HUTCHINSON, T. W. WATSON,

171

97. w. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108.

and

I. L. DILLAMORE: J. Iron

Steel I11St., 1969, 207, 1479.


I. S. BRAMMAR, T. R. THOMSON,

and

R. M. HOBBS: J. Aust. I11St.

Met., 1972, 17, 147.

Published by Maney Publishing (c) IOM Communications Ltd

109. 110. 111.

112. 113. 114. 115. 116.

and J. L. FORAND: in 'Mechanical working and steel processing VII', 127; 1969, New York, Metallurgical Society of AIME. P. R. V. EVANS, J. C. BITCON, and I. F. HUGHES: J. Iron Steel Inst., 1969, 207, 331. J. A. ELIAS and R. E. HOOK: in 'Mechanical working and steel processing IX', 348; 1970, New York, Metallurgical Society of AIME. R. H. GOODENOW and J. F. HELD: Metall. Trans; 1970, 1,2507. I. F. HUGHES and R. c. HUDD: Br. Pat. 1 236 598, 1971. P. R. MOULD and J. M. GRAY: Metall. Trans., 1972, 3, 3121. R. GILLANDERS, C. DASARATHY, and R. c. HUDD: in 'Textures and the properties of materials', 245; 1976, London, The Metals Society. P. MESS lEN and T. GREDAY: in 'Textures and the properties of materials', 266; 1976, London, The Metals Society. D. J. WILLIS and M. HATHERLY: in 'Textures and the properties of materials', 48; 1976, London, The Metals Society. U. LOTTER, W. MOSCHENBORN, and R. KNORR: in Proc. 5th Int. Conf. on 'Textures of materials', Vol. 2, 285; 1978, Berlin, Springer-Verlag. P. MESSIENand T. GREDAY: CRM Metall. Rep., Dec. 1976,(49),3. E. P. ABRAHAMSON, IT and B. S. BLAKENEY, Jr: Trans. AIME, 1960, 218, 1101. W. BLECK and u. LOTTER: in Proc. 7th Int. Conf. on 'Textures of materials', 383; 1984, Noordwijkerhout, Netherland Society for Materials Science. T. MATSUOKA and M. TAKAHASHI: Tetsu-to-Hagane (J. Iron Steel I11St.Jpn), 1971, 57, 1134. N. FUKUDA and M. SHIMIZU: Tetsu-to-Hagane (J. Iron Steel I11St.Jpn), 1975, 61, 817. H. ABE and K. TAKAGI: Tetsu-to-Hagane (J. Iron Steel Inst. Jpn), 1975, 61, S141.
D. A. KARLYN, R. W. VEITH, I. KOKUBO, M. SUDO, K. KAMENO, S. HASHIMOTO, I. TSUKATANI,

117. 118. 119. 120. 121.

122. 123. 124. 125. 126. 127. 128. 129. 130. 131. 132.

and T. IWAI: Tetsu-to-Hagane (J. Iron Steel Inst. Jpn), 1973, 59,469. T. SUZUKI: in Proc. 7th Int. Conf. on 'Textures of materials', 439; 1984, Noordwijkerhout, Netherlands Society for Materials Science. K. MATSUDO, T. SHIMOMURA, K. OSAWA, M. SAKOH, and s. ONO: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 2, 759; 1981, Tokyo, The Iron and Steel Institute of Japan. H. HU: Texture Cryst. Solids, 1979, 3, 215. N. OHASHI, M. KONISHI, A. YASUDA, S. SATO, and T. IRIE: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 1, 195; 1981, Tokyo, The Iron and Steel Institute of Japan. M. SUDO, S. HASHIMOTO, and I. TSUKATANI: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 2, 1076; 1981, Tokyo, The Iron and Steel Institute of Japan. M. SUDO and I. TSUKATANI: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 2, 807; 1981, Tokyo, The Iron and Steel Institute of Japan. D. A. LEAK and G. M. LEAK: J. Iron Steel Inst., 1958, 189, 256. K. TODA, H. GONDOH, H. TAKEUCHI, M. ABE, N. UEHARA, and
K. KOMIYA: Testu-to-Hagane

(J. Iron Steel lnst. Jpn), 1975,

61,2363. and R. J. ARSENAULT: in 'Treatise on materials science and technology', Vol. 1, 179; 1972, New York, Academic Press. K. OSAWA and K. KURIHARA: in 'Memoirs of the research committee of low carbon sheet steels', 108; 1987, Tokyo, The Iron and Steel Institute of Japan. H. ABE, T. SUZUKI and s. OKADA: Testu-to-Hagane (J. Iron Steel Inst. Jpn), 1983, 69, S1415. H. ABE: in Proc. 8th Int. Conf. on 'Textures of materials', 661; 1988, Warrendale, PA, Metallurgical Society of AIME. A. OKAMOTO and M. TAKAHASHI: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 2, 739; 1981, Tokyo, The Iron and Steel Institute of Japan. c. J. E. SMITH and I. L. DILLAMORE: Met. Sci. J., 1970, 4, 161. I. L. DILLAMORE, P. L. MORRIS, C. J. E. SMITH, and w. B. HUTCHINSON: Proc. R. Soc., 1972, A239, 405. R. L. EVERY and M. HATHERLY: Texture, 1974, 1, 183. J. BOURGEOT, J. L. LEBRUN, Y. MEYZAUD, P. PARNIERE, and B. J. THOMAS: Report RE 770, IRSID, Aug. 1980.
D. F. HASSON

International

Materials Reviews

1994

Vol. 39

No.4

172 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144.

Ray et al.

Textures in low and extra low carbon steels and s.


NAGASHIMA: Trans. AIME,

H. TAKECHI, H. KATO,

1968,

242,56.
M. MATSUO, S. HAYAMI,

and s. NAGASHIMA: in Adv. X-ray Anal.,

1974, 14, 214.


1. L. LEBRUN, G. MAEDER, F. MOLIEXE, and P. PARN-mRE: EEC Technical Research on Steel, Report No. EUR 6652, 1980. D. BENOIT, Y. MEYZAUD, P. PARNI:ERE, and R. TIXIER: in 'Texture and the properties of materials', 13; 1976, London, The Metals Society. H. KUBOTERA and K. NAKAOKA: in Proc. Conf. on 'Mechanical working and steel processing', 101; 1967, Metals Park, OH, American Society for Metals. B. J. DUGGAN: MSc thesis, University of Birmingham, 1970. 1. L. DILLAMORE and w. B. HUTCHINSON: in Proc. ICSTIS (suppl. Trans. Iron Steel Inst. lpn), 1971, 11, 877. H. INAGAKI: Trans. Iron Steel Inst. Jpn, 1984, 24, 266. w. B. HUTCHINSON: Acta Metall., 1989, 37, 1047. K. USHIODA, H. OHSONE, and M. ABE: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 2, 829; 1981, Tokyo, The Iron and Steel Institute of Japan. w. B. HUTCHINSON: in Proc. 8th Int. Conf. on 'Textures of materials', 603; 1988, Warrendale, PA, Metallurgical Society of AIME. u. von SCHLIPPENBACH and K. LUCKE: in Proc. 8th Int. Conf. on 'Textures of materials', 861; 1988, Warrendale, PA, Metallurgical Society of AIME. P. PARNrERE: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 1, 181; 1981, Tokyo, The Iron and Steel Institute of Japan. J. L. LEBRUN, G. MAEDER and P. PARNIERE: in Proc. 6th Int. Conf. on 'Textures of materials', Vol. 2, 787; 1981, Tokyo, The Iron and Steel Institute of Japan. w. G. BURGERS and P. c. LOUWERSE: Z. Physik, 1931, 67, 605.

Published by Maney Publishing (c) IOM Communications Ltd

145. 146. 147.

148. w. G. BURGERS and T. J. TIEDEMA: Acta Metall., 1953, 1, 234. 149. P. A. BECK: Adv. Phys., 1954, 3, (11), 245. 150. P. A. BECK: Acta Metall., 1953, 1, 230. 151. P. A. BECK and H. HU: in 'Recrystallization, grain growth and textures', (ed. H. Margolin), 393; 1966, Metals Park, OH, American Society for Metals. 152. K. LUCKE: Can. Metall. Q., 1974, 13, 261. 153. D. J. BLICKWEDE: Trans. ASM, 1968, 61, 653. 154. E. LINDH, W. B. HUTCHINSON, and P. BATE: in Proc. 10th Int. Conf. on 'Textures of materials', 997-1002; 1994, Aedersmannsdorf, Switzerland, Trans Tech Publications. 155. Y. MEYZAUD and P. PARNI:ERE: Report RFP 238, IRSID, Jan. 1975. 156. U. KOSTER: in 'Recrystallization of metallic materials', 215; 1971, Berlin, Dr Riederer Verlag. 157. R. D. DOHERTY and J. W. MARTIN: J. Inst. Met., 1962-63,91,332. 158. w. B. HUTCHINSON: Personal communication. 159. E. HORNBOGEN and H. KREYE: in 'Textures in research and practice', 274; 1969, Berlin, Springer-Verlag. 160. E. HORNBOGEN: Metall. Trans., 1979, lOA, 947. 161. M. TAKAHASHI and A. OKAMOTO: Sumitomo Met., 1974,27,40. 162. P. AUBRUN and P. ROCQUET: Mem. Sci. Rev. Metall., 1975,72,1. 163. B. J. DUGGAN and w. B. HUTCHINSON: in 'Textures and the properties of materials'; 292; 1976, London, The Metals Society. 164. w. B. HUTCHINSON and B. J. DUGGAN: Met. Sci., 1978,12,372. 165. Y. MEYZAUD, P. PARNIERE, and B. J. THOMAS: Report RE 434, IRSID, Mar. 1977. 166. Y. MEYZAUD, P. PARNI:ERE, B. J. THOMAS, and R. TIXIER: in Proc. 5th Int. Conf. on 'Textures of materials', Vol. 2, 243; 1978, Berlin, Springer-Verlag. 167. E. BOMMIER and F. MOLIEXE: Unpublished results, IRSID. 168. H. J. BUNGE and u. KOHLER: Scr. Metall. Mater., 1992,27, 1539.

International

Materials Reviews

1994

Vol. 39

No.4

You might also like