Sunteți pe pagina 1din 8

Recrystallization kinetics of AI-Mg alloys

AA 5056 and
AA 5083 after hot deformation

Slabs of two commercial AI- Mg alloys were hot rolled on a laboratory mill. A wide range of processing parameters was used in the experimental design and the effect of those parameters on the annealing kinetics was established. The effects of each parameter are described by simple empirical relationships. The recrystallization time requiredfor a given volume fraction recrystallized was found to decrease with increasing total hot strain and with decreasing initial grain size. The recrystallization time also decreased with increasing temperature compensated strain rate. The annealing temperature also has a profound effect. Empirical relationships incorporating.all the process variables are presented, and it is shown that the alloy AA 5056 recrystallizes more readily than does alloy AA 5083. It is concluded that this observation arises because of the larger volume fraction of precipitates present in the homogenized 5083 alloy. MSTj357

N. Raghunathan

M. A. Zaidi T. Sheppard

1986 The Institute of Metals. Manuscript received 6 September 1985; in final form 24 February 1986. At the time the work was carried out the authors were in the Department of Metallurgy and Materials Science, Imperial College of Science and Technology, London. Dr Zaidi is now at the Alcoa Technical Center, Alcoa Center, Pa, USA.

Published by Maney Publishing (c) IOM Communications Ltd

Introduction
The recovery and recrystallization of deformed structures is an integral part of the thermomechanical process. This annealing process may occur wholly or partially between rolling passes during hot processing, and may be used in some cases to refine the grain size during such processing. The softening may also be utilized simply to facilitate further processing such as cold rolling or forming of the engineering component after the rolling operation. Clearly, a process which may occur several times during the production of the final component must be regarded as technologically important. Although some softening can occur by static recovery events, complete softening can be achieved only by replacing the deformed structure with new strain-free recrystallized grains. Most of the existing literature concerns itself with the mechanism of nucleation and growth of such grains, usually from a cold deformed structure, and provides little or no quantitative data which could relate the softening processes to the conditions of deformation. Furthermore, it is, in general, model systems which have been analysed, utilizing pure metals or simple binary alloys to provide experimental data; results concerning commercial alloys have, in general, not been reported. Recently, however, some parametric studies mostly concerning ferrous alloys have been published 1-8 which indicate the important role of the deformation variables on the softening process. In the previous work, direct metallographic techniques indicated that the recrystallization kinetics were dependent upon the total strain,1-8 the strain rate and temperature of deformation (a measure of the stored deformation energy), and the annealing temperature (the thermal energy supplied to assist the motion of high-angle boundaries).1-6 It has also been widely reported that nucleation occurs preferentially at grain boundaries,9-11 and hence it would be reasonable to expect that the initial grain size might influence the recrystallization kinetics. Further analysis of Refs. 1-11 reveals that the range of stress investigated has been limited and, in general, has not approached the strain
Table 1
Alloy

encountered in industrial practice, primarily owing to the mode of deformation utilized in the experimental trials. Hence, the effect of hot strain does not appear to have been adequately established. The work described in this paper is concerned with the effects of hot deformation parameters on the recrystallization kinetics of two commercial AI-Mg alloys (AA 5056 and AA 5083), in which the selected deformation mode is rolling, in order to establish the effects of a wide range of each variable.

Experimental procedure
The alloys used were supplied in the form of direct chill semicontinuous cast logs having a cross-section of 210 x 84 mm. The composition of the two alloys is given in Table 1. Both materials were homogenized in an air circulating furnace, the homogenization procedure involving a soak at 575C for 12 h, slow cool to 515C, hold for 4 h, and furnace cool to room temperature. Rolling billets of crosssection 45 x 70 mm and 200 mm length were machined from the homogenized slab. Rolling experiments were conducted on a two-high laboratory mill having rolls 300 mm in diameter and 350 mm in width. The roll speed was maintained at 10 rev min -1, giving an average strain rate of about 2 32 s - 1. Each billet was cleaned and degreased before heating to the rolling temperature in an electric resistance furnace. Rolling to the required reduction was achieved in multipass schedules; the billet was quenched and reheated after each pass and inspected metallographically to ascertain that recrystallization had not occurred during the quench and reheat procedure. To achieve reductions greater than 75%, the billets were sandwiched between 5 mm aluminium sheets and rolled to give a maximum reduction of 90%. Experiments were thus designed to study the effect of total hot strain, rolling temperature, and initial grain size on the annealing characteristics of the alloys. The initial grain size was varied by repeated rolling and cross-rolling combined with intermediate anneals.

Composition
Cu Fe

of AI-Mg
Mg

alloys, wt_% Mn 013 070


Si

Ti 0014 0016

Zn 001 001

Cr

B (ppm)

AI

AA 5056 AA 5083

0013 0002

016 018

498 455

005 005

012 017

8 14

Bal. Bal.

938

Materials Science and Technology

September 1986

Vol. 2

Raghunathan

et al.

Recrystallization

kinetics of AI-Mg

alloys

939

Published by Maney Publishing (c) IOM Communications Ltd

Structure constituent at 575C

of alloy 5056 indicating differing particles present after homogenization

Structure constituent at 575C

of alloy 5083 indicating differing particles present after homogenization

For the annealing experiments, small samples of dimensions 10 x 10 x 3 mm were extracted from the centre of the rolled billets. The samples were then either annealed at constant temperature and the development of recrystallization studied with respect to time or annealed at varying temperature for a constant time period. The annealed samples were mounted, polished, and anodized using 5 vol.-% HBF 4 aqueous solutions at 20 V for a minimum of 2 min, and the volume fraction recrystallized established using an Omnicon image analyser. Anodizing produced a complex grain contrast which rendered the samples unsuitable for direct use in the image analyser. It was therefore necessary to trace the recrystallized grains from the optical micrographs and darken them before using the digital image analyser to calculate the area of the dark region and subsequently the area fraction on the tracings. The volume fraction was calculated from the above data, since the volume fraction is directly proportional to the area fraction.

Results STRUCTURAL CONSIDERATIONS

A detailed structural analysis resulting from this investigation is presented elsewhere.12 However, for the purposes of the present paper, it will be necessary to review some of the microstructural features of the alloys and these are shown in Figs. 1-4.

Figure 1 shows the structure of the AA 5056 alloy after homogenization. The grains were equiaxed with a mean grain size of 144 11m. Analysis of the homogenized structure indicated that three differing types of constituent particle size were present. The coarse grain boundary particles were identified using energy dispersive X-ray and electron diffraction analysis to be Al1s(Fe, MnhSi and are the remnants of the cast structure but in spheroidized form. These particles are potentially active nucleation sites during the recrystallization process. The matrix of C( solid solution also contains precipitates of spherical and rod like morphology. Samples quenched from the homogenization soak temperature revealed only the rodlike precipitates, indicating that the spherical precipitates formed during the cooling sequence. Hence, the spherical precipitates may well redissolve during the heating sequence before rolling. The rodlike precipitates were identified as an MgMnAI phase, while the majority of the spherical precipitates were of the Mg2Al3 type and some were identified to be MnAI6. These particles would act primarily as growth inhibitors during the annealing process. Figure 2 is a micrograph of the AA 5083 alloy after homogenization and the structure is in many respects similar to that of the AA 5056 alloy, the mean grain size being larger at 163 11m. However, the matrix contains predominantly globular precipitates which were identified as both MnAl6 and Mg2A13 The micrograph also indicated that the volume fraction of particles is much higher in the AA 5083 alloy which is a direct consequence

940

Raghunathan

et al.

Recrystallization

kinetics of AI-Mg

alloys

a
3

Published by Maney Publishing (c) IOM Communications Ltd

Banded structure in alloy 5056 preferential recrystallization near bands

indicating

of the manganese addition. Another important feature, common to both alloys, is the presence of precipitate-free regions close to the grain boundaries which, combined with the decoration of the boundaries by coarse particles, will be an important structural feature influencing postdeformation annealing. One of the important features of the deformed material was the presence of deformation bands. Figure 3 shows a typical example of this banded structure in the 5056 alloy which was observed in all large grained material subjected to moderate deformations (i.e. < 50%). Such banding was difficult to detect in material deformed at the higher deformation temperatures and also in the 5083 alloy where the presence of a larger volume fraction of precipitate tended to mask this phenomenon. Figure 4 illustrates some of the features of the partially recrystallized hot rolled samples. The recrystallized grains have nucleated at the grain boundaries and at the large second phase intermetallics, at shear bands, and within grains in each alloy; this was observed for all deformation conditions. Nucleation can thus be seen to vary from grain to grain and is heterogeneous. RECRYSTALLIZATION KINETICS The volume fraction recrystallized Xv obtained at varying times during the isothermal annealing experiments produced a curve of the sigmoidal type as shown in Fig. 5a for the alloy 5056 at varying temperatures. Similar curves were also observed when the annealing temperature was held constant and the true strain and critical (initial) grain size were respectively held constant (Figs. 5c and e). This is consistent with the recovery, nucleation, and growth equation proposed by Johnson and Mehp3 and Avrami,14-16 which is of the form Xv
=

b
5056 annealed at 330C for 120 s after deformation at 300C to equivalent strain of 052; b alloy 5083 annealed at 430C for 60 s after deformation at 300C to equivalent strain of 051

a alloy

Partially

recrystallized

structures

1-exp (-fJtn)

(1)

which is an isothermal expression in which fJ and n are constants at anyone temperature and t is the annealing time. The data were thus processed and the slope of the In In [1/(1- Xv)] versus time graphs evaluated to yield the values of n for differing deformation conditions (Figs. 5b, d, andf). Recrystallization kinetics in the 5083 alloy also showed sigmoidal behaviour, but nucleation proved more difficult in this alloy. It was observed that the annealing temperature had to be varied with the deformation conditions in order to obtain results which yielded useful analytical data. Figures 6a and b indicate the combination of experimental procedures which was found necessary.

The results in Fig. 6 show the effect of varying annealing temperature, total hot strain, and initial grain size for the 5083 alloy. The value obtained for the time sensitivity n varied from 05 to 20 depending on the deformation conditions. The sensitivity was observed to increase with increasing strain and decreasing grain size and to remain substantially invariant with the rolling temperature. When the plots were non-linear, the time sensitivity was determined using data only from the linear portion of the graphs. The nonlinearity was usually observed at very high or very low volume fractions recrystallized which are inherently points for experimental error. Since recrystallization is a thermally activated process, with it there is associated a unique activation energy. Glover and Sellars17 have observed that if the activation energy required to sustain the recrystallization process is to be meaningful, then it must be derived from isothermal tests in which a sufficient driving force has been imparted to the material by its deformation history. Hence, in this investigation, the activation energy was obtained from data derived from rolling experiments in which the slabs were rolled at 300C to a total strain of o 52. The reaction kinetics can best be described by considering times required for a given fraction of the total reaction to occur, and hence the times required for different fractions of the structure to recrystallize were utilized to calculate the activation energy. The time required to achieve 05 volume

Materials Science and Technology

September 1986

Vol. 2

Raghunathan

et al.

Recrystallization

kinetics of AI-Mg

alloys

941

1-0 06 06
x
>-

lO

1-0

05
04

05

02
(a)

3000e o 3500e 400C 450't


G

05 -05 -"5

:R~~4 H~:w/0
TIME, s

(et

(el

05 -05 -15 -25


(f)

-2

-25

Published by Maney Publishing (c) IOM Communications Ltd

103 TIME I S

a, b isothermal recrystallization after annealing at 300C and deformation to equivalent strain of 052 at various temperatures, do = 144 ~m; c, d effect of strain on isothermal recrystallization after annealing at 290C and deforming at 300C, do = 144 ~m; e, f effect of initial grain size on isothermal recrystallization after annealing at 290C and deformation to equivalent strain of 052

Recrystallization

behaviour of alloy 5056

fraction recrystallized during isothermal annealing of samples having the same deformation history, but annealed at two different temperatures, was thus established. The values derived were found to be 21029kJmol-1 for the 5056 alloy and 183 22 kJ mol-1 for the 5083 alloy. The error given above is not a statistical error. It arises because differing volume fractions are utilized to calculate the activation energy and the errors are given with respect to the energy calculated using the time required for 0,5 volume fraction transformation. The influence of strain on the recrystallization kinetics is shown in Fig. 5c for alloy 5056 and in Fig. 6c for alloy . 5083. The strain is defined as e = In [1/(1- r)], where r is the percentage reduction in thickness. The curves for each alloy are similar and the A vrami exponent n for these cases was calculated to vary from 1 to 2 with an increasing tendency toward 2 as the strain increased. The effect of lO
1-0

strain is more clearly shown by Fig. 7a, in which the time" required for 05 volume fraction recrystallization to'5 to occur is plotted against the total hot strain and is a loglog relationship. The shape of the curves indicates the probability of a saturation strain beyond which any increase in strain would not affect recrystallization. The curves suggest that a simple power law of the form (2) would adequately fit the data and regression analysis yielded values of m of - 226 for alloy 5056 and - 212 for alloy 5083. An interesting feature which can be clearly seen in Figs. 5c and 6c is that, irrespective of total strain, recrystallization begins very rapidly in each of the alloys. Figures 5e and 6e show the influence of initial grain size when the material has been deformed to a true thickness strain e = 0'52. Increasing grain size clearly delays
1-0

.. .

~05

~_::.=: ~fi~/
~-15 -25

oG

Ta/oC 346

4
~.
G..

05

05

(al Tr C 0.300 x 450

(e)

t~g

05 500 -05

.051 093 G 1-63 o 243

05 -05 -1-5 -25


(d)

027~ EJ 49~ x163~

/
103
TIME ,s

(el

'?7
( b)

-1-5 -25

10 TIME,
S

103

104

102

103 TIME ,s

104

10'

a, b

isothermal recrystallization after annealing and deformation to true thickness strain of 051 at various temperatures, do = 163 ~m; C, d effect of hot strain on isothermal recrystallization after annealing at 400C and deformation at 300C, do = 163~m; e, f effect of initial grain size on isothermal recrystallization after annealing at 330C and deformation to equivalent strain of 052

Recrystallization

behaviour of alloy 5083; Ta and Tr are annealing and rolling temperature,

respectively

942

Raghunathan

et al.

Recrystallization

kinetics of AI-Mg

alloys

G 5056 o 5083

5083

Published by Maney Publishing (c) IOM Communications Ltd

163~

o
(a)

102 10 GRAIN SIZE, ~ 102

10

TRUE STRAIN a strain;


7
b grain size (e = 0'53)

Effect of strain and grain size on to'5

commencement of recrystallization. The Avrami exponent was found to vary from 1-4 for a grain size of 144 flm to 183 for a grain size of 22 flm for the 5056 alloy. Similar data for the 5083 alloy (Fig. 6e) yielded values of 08 for the 163 flm material to 20 for the 27 flm material. The time for 50% recrystallization is shown varying with grain size

in Fig. 7b and a simple power law fit is again suggested. Thus, we may write to'5 oc d~
(3)

EI 5056

o
-20
0

50B3

ex:

c5
I

u
CD

163~

~
-oJ

c..

c::.

~-30

where do is the initial grain size, and evaluate k to be 158 and 245 for the 5056 and 5083 alloy, respectively. Figures 5a and 6a clearly indicate the effect of rolling temperature on the annealing process. As the rolling temperature is increased, recrystallization becomes significantly -more difficult. Figure 5a in particular shows that when the rolling temperature is increased from 400 to 450C the kinetics are very considerably retarded, which might suggest that recrystallization is occurring conjointly with other processes or quite simply that the stored energy is insufficient at this temperature. Such anomalous behaviour was not observed for the 5083 alloy, although most of the data (Fig. 6a) indicate much slower kinetics for this alloy. However, this can be adequately explained by the larger volume fraction of both the rodlike and the fine precipitates shown in Fig. 2, which will clearly inhibit the recrystallization process. The data presented in Figs 5a and 6a can be more conveniently analysed by introducing the conceptI7 of a temperature compensated time WXv' defined by an equation of the form
WXv = t Xv exp ( - Qrec/ R 1;ec)

(4)

-~O

where txv is the time required to recrystallize the volume fraction Xv; Qrec and 1;ec are the activation energy and annealing temperature, respectively; and R is the universal gas constant. Such a temperature compensated time is shown plotted against the temperature compensated strain rate Z in Fig. 8 in which Z is defined as
30 35 LN Z 40 45
for

25
8

Z = e exp (QderlRTdef)

(5)

Effect of Z on temperature compensated time 05 volume fraction recrystallized (E = 053)

Here Qdef and Tdef are the activation energy for deformation (obtained from torsion test results) and

Materials Science and Technology

September 1986

Vol. 2

Raghunathan

et al.

Recrystallization kinetics of AI-Mg alloys

943

05

>-

(al

Published by Maney Publishing (c) IOM Communications Ltd

05

(b)

a
9

alloy 5056; b alloy 5083

Isochronal curves for alloys 5056 and 5083 deformed to true strain of 053 at various temperatures

nucleation of new grains occurs predominantly at grain boundaries where the substructure has a fine morphology,12 but can also be observed in grain interiors, especially at heterogeneities such as deformation bands and large precipitates. Similar observations have been reported for ferrous materials 1 and for an Al-IMg alloy.19,20 The recrystallization functions presented in Figs. Sa, b and 6a, b indicate a conventional sigmoidal behaviour indicative of a nucleation and growth process. The use of Avrami type equations is thus justified. The magnitude of the exponent in the Avrami equation is dependent upon the nature of nucleation, and for grain boundary nucleation after 60% cold work a value of n = 2 has been reported.21 In the present investigation, it has been shown that this exponent varies from 05 to 20 and this variation cannot be due to experimental errors. Indeed, comparing Figs. Sa, band 6a, b it is evident that in many cases the plots deviate from the linearity defined by equation (1), suggesting that the equation is slightly inadequate in modelling the kinetics at higher volume fraction recrystallized; similar behaviour has been reported for vanadium.22 This deviation is greatest for coarse grained material which has received least deformation before the anneal. Under these conditions, deformation will be concentrated at grain boundaries and heterogeneities and this will not be typical of more normal experimental conditions. Although nucleation can occur readily during the early stages of the process, further recrystallization by normal subgrain coalescence occurs later23 and the continuation of the process is thus delayed. Hence, considering this heterogeneous nature of recrystallization, and the fact that the data are statistically insufficient to suggest an alternative value of the Avrami exponent, it is considered justified to fit the data to a value of n = 20. This is consistent with the predominant grain boundary nucleation observed.

temperature of deformation, respectively. Because i; is the local strain rate at different locations in the deformation zone, it becomes a function of location in the deformation zone. Hence, an average strain rate e in the roll gap is used defined by 18

FACTORS AFFECTING

RECRYSTALLIZATION

~ 2nN -1 e=-- [.SIn (ai1h - (i1h/2)2)]-ll ----60 a

n-

hI ho

(6)

where a is the roll radius, N the roll speed, hI and ho the final and initial slab thicknesses, and i1h = hI - hOt The linearity of the plot on a log-log scale suggests a power function of the type to. 5 exp ( - Qrec/ R 7;ec) ex Z - A (7)

The power exponent A was found to be - O35 for the 5056 alloy and -058 for the 5083 alloy. In order to ascertain the effect of annealing temperature on the kinetics, and also to determine the temperature at which complete recrystallization may be achieved for any given time, samples of both alloys were annealed for 1 h at varying temperatures between 250 and 500ae. The results are shown in Fig. 9, indicating the opposing effects of increasing annealing temperature and increasing rolling temperature. It is also apparent that for similar conditions the alloy 5083 shows a greater resistance to recrystallization. The reader should of course be aware that the Avrami equation describes a mixed recovery, nucleation, and growth mechanism, and neither recovery nor growth has been separately investigated as functions of time in this work.

Discussion KINETICS
The structural evidence presented in Fig. 4a indicates that

The effect of strain on to'5 shown in Fig. 7a is in agreement with that reported for an Al-l Mg alloy6 and for ferrous alloys. 1 The exponent in the power function is, however, much lower for the 5056 and 5083 alloys than the value reported for steel. An earlier publication 1 also indicated that the Avrami exponent was independent of strain, while in these experiments this is clearly not the case. There are two possible reasons for this contradiction. The most likely is the simple observation that a much wider range of strains to far higher deformations were used in the experiments reported here. The other major difference was that this work utilized coarser grained material which at low strains introduces larger strain inhomogeneities tending to concentrate deformation at grain boundaries and large precipitates. This will tend to increase the time required for recrystallization, thus lowering the power exponent, while the former observation with increments of strain having less effect with increasing strain will also result in a lower exponent. Thus far, a simple power function has been utilized to fit all the data points. Inspection of Fig. 7a, however, shows that data obtained at higher strains do not fit this relationship particularly well. It has also been suggested above that a saturation strain may exist, beyond which further increments may not affect the kinetics of recrystallization. Alternative correlations were thus investigated based upon the form of stress-strain relationships in which stress asymptotically achieves a steady state value. This was because, in the present investigation, it was observed that time required for 05 volume fraction recrystallized obtained a constant value asymptotically with respect to strain. This would appear to be logical since the high recovery rates in the alloys investigated imply that the stored energy available for recrystallization will not

944

Raghunathan

et a/.

Recrystallization

kinetics of AI-Mg

alloys

105

5056 o 5083

104
In

e
0

0 0

.:3
103

je
1014 1015 Z>. d~ exp(Qrec/RTrec)(X+YE2rl 1016 1017

10

10

Master curves for recrystallization kinetics; values of constants A.,k, x, yare given in text

Published by Maney Publishing (c) IOM Communications Ltd

increase after some finite amount equations took the form to. 5 oc (00286

of strain.

The resulting

+ 18t; - 2) 2 )-1

for alloy 5056

to'5 oc (9'73+3'82e-

for alloy 5083

In view of the predominance of grain boundary nucleation, it was assumed that initial grain size would have a significant effect on recrystallization kinetics. Indeed, it has been reported 11 in a theoretical investigation that to'5 is proportional to dg'75 for grain boundary nucleated recrystallization. In the present investigations, such power laws fit the data of Fig. 8 very well but exhibit exponents of 158 and 2-45 for the alloys 5056 and 5083, respectively. The obvious explanation, since a greater dependency on grain size is indicated, is that events other than grain boundary nucleation occur quite readily. It has been shown21 that for some AI-Mg alloys the recrystallized grain size is strongly related to the subgrain size obtaining before the annealing process, indicating the strong possibility of sub grain coalescence or growth within grains. If it is assumed that subgrain distributions are homogeneous, then it is likely that the number of nucleation events necessary would also be related to grain size, thus apparently increasing the influence of initial grain size on the annealing process. It should also be observed, however, that there is an alternative explanation. The differing grain sizes were prepared by mechanical working and annealing before commencement of the experiments. Thus, each specimen would exhibit an annealed texture which has been reported to influence recrystallization after deformation in steel. 18 The relationship indicated in Fig. 8 could therefore be due to the combined effect of grain size and texture. Finally, in this context, the reader should recall that both large and small precipitates exist in both alloys and these features would also ensure that nucleation is not purely a grain boundary phenomenon. The Zener-Hollomon parameter is very closely related to the substructure produced in a rolling pass and hence by producing varying stored energy might be expected to influence recrystallization. Figure 8 shows the relationships obtained between the temperature compensated time and the temperature compensated strain rate plotted on a log-log axis. A linear fit to the data is obvious. However, with reference to the isothermal curves, Figs. 5a, band 6a, b indicate that recrystallization is retarded rather more than would be expected when the rolling temperature exceeds 400C. This may be explained by considering the interaction between precipitate particles and the stored energy providing the driving force for recrystallization. It

has previously been reported23 that in AI-2Mg alloys the migration of high-angle grain boundaries is preceded by the dissolution of the Mg5Al8 particles usually located at boundaries. Thus, in a structure in which stored energy, and hence the driving force for recrystallization, exceeds the drag force exerted by precipitate particles, the presence of such particles is unlikely to affect the recrystallization process profoundly. If, however, the reverse situation applies, as happens when the deformation temperature is raised, then a contrary conclusion may be reached. Moreover, precipitate dissolution is a diffusion assisted transport process, and hence it is only the early stages of recrystallization which are likely to be affected when the latter condition obtains. This clearly occurs in alloy 5056 (Fig. 5a, b) and is prevalent to a larger extent in alloy 5083 in which the recrystallization kinetics are hindered both by soluble precipitates and those which show little solubility at annealing temperatures. The temperature of annealing has a profound influence on recrystallization (Fig. 9). This is not surprising because it is the combination of stored deformation energy and supplied thermal energy which provides the necessary driving force for recrystallization. The activation energy for recrystallization calculated from the experimental results is much greater than that for self-diffusion. It is not possible to postulate a clear physical significance for the value obtained because an average activation energy has been derived for a nucleation and growth process partly modified by precipitate dissolution kinetics. The value is much higher than that reported for pure aluminium but, importantly, the activation energy derived from experiments such as those reported here for commercial alloys is of significant industrial import. From the results presented above and in Figs. 5-9, it is possible to propose relationships describing recrystallization in the alloys investigated. Of course, the effect of second-phase particles, precipitates, etc., cannot be specifically included in such a relationship because separate experiments would be required to dissociate their effect from the parameters which have been investigated. Nevertheless, the form and definition of the equations are practical and should be applicable in the industrial case. Combining the data from all experiments the following equations are obtained: Xv = 1-exp[

-0693(t/tO.5)2J
0.0286

91 x 10-12db'58Z-0'35exp(212000/R4eJ

to'5 =
for alloy 5056

+ 18e2
exp (183 000/R4eJ

27 x 10-10d5'45Z-0'58
to'5

=
for alloy 5083

9.73

+ 382e2
with the experimental

These relationships results in Fig. 10.-

are

shown

Conclusions
The experiments have shown that recrystallization after hot deformation is complex. The hot rolling parameters have considerable influence on recrystallization kinetics with the total hot strain and temperature compensated strain rate of the final pass being particularly dominant. It has also been shown that the initial grain size is an important parameter. Hence, the annealing temperatures required for recrystallization can encompass a wide spectrum, and, therefore, if fully soft material is required for further processing in industry, the metallurgist must carefully control both casting and thermo mechanical processing.

Materials Science and Technology

September 1986

Vol. 2

Raghunathan

et al.

Recrystallization

kinetics of AI-Mg

alloys

945

References
1.
D. R. BARRACLOUGH

and c.

M. SELLARS:

Met. Sci., 1979, 13,

257. 2. D. J. TOWLE and T. GLADMAN: Met. Sci., 1979, 13, 246. 3. c. M. SELLARS and J. A. WHITEMAN: Met. Sci., 1979, 13, 187. 4. J. J. JONAS, C. M. SELLARS, and w. M. McG. TEGART: Me tall. Rev., 1969, 14, 1. 5. H. J. McQUEEN and J. J. JONAS: in 'Plastic deformation of metals', (ed. R. J. Arsenault), 281; 1975, New York, Academy Press. 6. c. M. SELLARS: in 'Hot working and hot forming', (ed. C. M. Sellars and W. J. Davies), 3; 1979, London, The Metals Society, 7. R. A. P. DJAIC and J. J. JONAS: J. Iron Steel Inst., 1972, 210, 256. 8. R. A. P. DJAIC and J. J. JONAS: Me tall. Trans., 1973, 4, 621. 9. A. T. ENGLISH and w. A. BACKOFEN: Trans. AIME, 1964, 230, 369. 10. G. R. SPEICH and R. M. FISHER: in 'Recrystallisation, grain

growth and texture', (ed. H. Margolin), 563; 1966, Metals Park, Ohio, The American Society for Metals. 11. J. w. CAHN: Acta Metall;, 1956, 4, 449. 12. N. RAGHUNATHAN and T. SHEPPARD: in 'Aluminium technology '86', (ed. T. Sheppard), 357; 1986, London, The Institute of Metals. 13. w. A. JOHNSON and R. F. MEHL: Trans. AIME, 1939, 135, 416. 14. M. AVRAMI: J. Chern. Phys., 1939, 7, 1103. 15. M. AVRAMI: J. Chern. Phys., 1940,8,8. 16. M. AVRAMI: J. Chern. Phys., 1941,9,117. 17. G. GLOVER and c. M. SELLARS: Metall. Trans., 1972, 3, 2271. 18. H. J. WHITTAKER: PhD thesis, The University of Sheffield, 1973. 19. D. S. WRIGHT and T. SHEPPARD: Met. Technol., 1981, 8, 180. 20. E. D. CABRARA: PhD thesis, The University of Sheffield, 1983. 21. P. R. MOULD and P. COTTERILL: in 'Recrystallisation and grain growth in metals', 43; 1976, London, Surrey University Press. 22. c. w. PRICE: Scr. Metall., 1985, 19, 669. 23. M. A. ZAIDI and T. SHEPPARD: Met. Sci., 1982, 16, 229.

Published by Maney Publishing (c) IOM Communications Ltd

DISLOCATIONS AND PROPERTIES OF REAL MATERIALS


In view of the important role that dislocations are now known to play in the behaviour of materials as diverse as semiconductors and the earth's crust and in determining the properties of everyday substances, this international conference, sponsored and organised by The Metals Society to celebrate the Fiftieth Anniversary of the Concept of Dislocation in Crystals and held in London in December 1984, assessed important contributions of the experimental and theoretical study of dislocations, as well as future developments in the subject. Contents include: perspectives on the early days of dislocations on the plastic proper~ies

Historical

Effect of dislocation core structure of metallic materials Dislocations Solution and phase transformations

hardening in fatigue and the plasticity and geological of ionic crystals

Dislocations Dislocations Dislocation

deformation

BOOK 323 ISBN 0 904357 74 0 280 x 210mm 348pp Paperback Pub 1 ished
PRICE

Please

send order,

with

remittance,

to:

1985

THE INSTITUTE OF METALS Subscriber Services Department 1 Carlton House Terrace London SW1Y 5DB Tel. Telex 01-839 8814813 4071

UK 29.50 (Institute

OVERSEAS US$41.50 of Metals members deduct

20%)

Materials Science and Technology

September 1986

Vol. 2

S-ar putea să vă placă și