An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

download An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

of 15

Transcript of An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    1/15

    Materials Science and Engineering A283 (2000) 274288

    An experimental study of the recrystallization mechanism duringhot deformation of aluminium

    S. Gourdet, F. Montheillet *

    Ecole Nationale Superieure des Mines de Saint-Etienne, Centre Science des Materiaux et des Structures, URA CNRS 1884, 158 Cours Fauriel,

    42023 Saint-Etienne Cedex 2, France

    Received 10 February 1999; received in revised form 22 October 1999

    Abstract

    Discontinuous dynamic recrystallization (involving nucleation and grain growth) is rarely observed in metals with high stacking

    fault energies, such as aluminium. In this metal, two other types of recrystallization have been observed: continuous dynamic

    recrystallization (CDRX, i.e. the transformation of subgrains into grains); and geometric dynamic recrystallization (due to the

    evolution of the initial grains). The main purpose of this work was to bring clearly into evidence and to better characterize CDRX.

    Uniaxial compression tests were carried out at 0.7 Tm and 102 s1 on three types of polycrystalline aluminium: a pure

    aluminium (1199), a commercial purity aluminium (1200) and an Al-2.5wt.%Mg alloy (5052), and also on single crystals of pure

    aluminium. In addition, 1200 aluminium specimens were strained in torsion. The deformed microstructures were investigated at

    various strains using X-ray diffraction, optical microscopy, scanning electron microscopy and electron back-scattered diffraction.

    Observations of the single crystalline samples confirm that subgrain boundaries can effectively transform into grain boundaries,

    especially when the initial orientation is unstable. In the case of polycrystalline specimens, after separating the effects of the initial

    and new grain boundaries, it turns out that CDRX operates faster in the 1200 aluminium compared to the two other grades.

    Moreover, it appears that the strain path does not alter noticeably the CDRX kinetics. 2000 Elsevier Science S.A. All rights

    reserved.

    Keywords: Hot deformation; Aluminium; Dynamic recrystallization; Single crystals; Subgrain boundaries; Grain boundaries; Misorientations

    www.elsevier.com/locate/msea

    1. Introduction

    Aluminium and its alloys exhibit very high rates of

    dynamic recovery, which is generally expected to com-

    pletely inhibit dynamic recrystallization. However, the

    formation of new grains during hot deformation of

    aluminium has been frequently reported. Three types of

    dynamic recrystallization are likely to produce such a

    microstructure: (i) discontinuous dynamic recrystalliza-

    tion (DDRX), i.e. the classical recrystallization, whichoperates by nucleation and grain growth; (ii) continu-

    ous dynamic recrystallization (CDRX), which involves

    the transformation of low angle boundaries into high

    angle boundaries; and (iii) geometric dynamic recrystal-

    lization (GDRX), generated by the fragmentation of

    the initial grains.

    Discontinuous dynamic recrystallization, which is

    commonly observed in low stacking fault energy

    metals, remains exceptional in aluminium and alu-

    minium alloys. Nevertheless, it seems to occur in two

    specific cases, viz., in high purity aluminium and in

    aluminium alloys containing large particles:

    1. Single crystals and polycrystals of 99.999 wt.% alu-

    minium have been subjected to hot deformation in

    compression by Yamagata [15]. Stress-strain

    curves exhibit strong oscillations, typical of DDRX,although more irregular. The associated microstruc-

    tures generally display new grains without substruc-

    ture. It is therefore not excluded that these grains

    have grown after deformation, all the more as

    aluminium of such high purity recrystallizes stati-

    cally very rapidly, even at room temperature. How-

    ever, one micrograph [3] clearly displays the

    presence of several grains containing a substructure,

    in an initially monocrystalline sample. Moreover,

    * Corresponding author. Tel.: +33-4-77420026; fax: +33-4-

    77420157.

    E-mail address: [email protected] (F. Montheillet)

    0921-5093/00/$ - see front matter 2000 Elsevier Science S.A. All rights reserved.

    PII: S 0 9 2 1 - 5 0 9 3 ( 0 0 ) 0 0 7 3 3 - 4

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    2/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288 275

    the occurrence of DDRX in high purity aluminium

    has been recently confirmed by Ponge et al. [6].

    High purity produces two opposite effects. On the

    one hand, it favors DDRX by increasing grain

    boundary mobility. On the other hand, it can inhibit

    DDRX since the high level of recovery prevents

    dislocation accumulation, thus reducing the driving

    force. Experimental results indicate that the first

    effect prevails over the second [7].2. There is some evidence that DDRX can occur dur-

    ing hot deformation of AlMgMn alloys, because

    a high Mg solute addition raises the dislocation

    density and thus the driving force for DDRX, while

    large Al6Mn particles (\1 mm) stimulate nucle-

    ation. The presence of small new grains adjacent to

    large particles has been reported for instance after

    plane strain compression of Al-1wt.% Mg-1wt.%

    Mn [8] and extrusion of Al-5wt.% Mg-0.8wt.%Mn

    [9]. However, the volume fractions of recrystallized

    grains remained small and no effect on the shapes of

    the stress-strain curves was detected. This means

    that DDRX is a possible but limited restorationmechanism in aluminium alloys.

    Continuous dynamic recrystallization occurs in turn

    by the progressive accumulation of dislocations in low

    angle boundaries, leading to the increase of their mis-

    orientation and the formation of large angle grain

    boundaries when their misorientation angles reach a

    critical value qc (qc:15). This mechanism has been

    observed in several high stacking fault energy metals,

    such as aluminium and aluminium alloys [1013], b

    titanium alloys [1416], and ferritic steels [1720]. The

    microstructure of a commercial purity (1050 grade)

    aluminium strained in torsion has been investigated by

    Perdrix et al. [10], and Montheillet [11]. These authorsfound that, at small and medium strains (m:1), the

    microstructure consists of the deformed initial grains

    containing subgrains, which is typical of a recovered

    state. By contrast, strongly strained samples (m:40)

    exhibit a completely different microstructure: it is no

    longer possible to distinguish the initial grains, and the

    former subgrains now appear as crystallites bounded

    partly by low and partly by high angle boundaries.

    Furthermore, the misorientation angles, which display a

    bimodal distribution at small strains (with subgrain

    boundaries less than 15 and initial grain boundaries

    between 30 and 63) become uniformly distributed be-

    tween 0 and 63 at large strains. Perdrix et al. [10] have

    explained these results by assuming a progressive trans-

    formation of subgrain boundaries into grain

    boundaries. However, this mechanism remains contro-

    versial and some authors have suggested that the in-

    creased fraction of high angle boundaries could result

    from GDRX.

    Nevertheless, there is a general agreement to consider

    that the transformation of low angle boundaries into

    high angle boundaries effectively takes place when the

    boundaries are pinned by small particles. This mecha-

    nism has been used to promote superplasticity in Zr

    bearing or high Mg aluminium alloys (Al-6wt.%Cu-

    0.4wt.%Zr [21,22], Al-0.25wt.% Zr-0.1wt.% Si [23,24],

    Al-10wt.% Mg-0.1wt.% Zr [25,26], Al-10wt.% Mg-

    0.5wt.% Mn [27,28]). These alloys are generally cold or

    warm rolled, to increase their dislocation density. Un-

    der these conditions, subgrains form very quickly dur-ing the subsequent hot tension testing. Since the

    boundaries are pinned by small particles (Al3Zr,

    (Al8Mg5)b or Al6Mn, according to the alloy composi-

    tion) and continuously absorb dislocations, the sub-

    grains transform into grains without growing. A fine

    and equiaxed grain structure is thus obtained in the

    early stages of superplastic deformation.

    Finally, geometric dynamic recrystallization has first

    been described by McQueen et al. [29] in a commercial

    purity aluminium. During deformation, the original

    grains flatten (compression) or elongate (tension, tor-

    sion), and their boundaries become progressively ser-

    rated while subgrains form. Consequently, the grainboundary area per unit volume grows strongly and an

    increasing fraction of subgrain facets is made of those

    initial grain boundaries. Ultimately, when the original

    grain thickness is reduced to about two subgrain sizes,

    the grain boundaries begin locally to come into contact

    with each other, causing the grains to pinch-off [29

    32]. In the case of Al-5Mg alloys, the tendency to the

    serration of boundaries is stronger and a secondary

    process of GDRX by pinching off of the serrations has

    been reported [33,34].

    The first purpose of this work was to bring forward

    further evidence for CDRX. The main objection to theresults of Perdrix et al. [10] was that the high fractions

    of large angle boundaries did not necessarily result

    from CDRX, but could also be due to the evolution of

    the initial grain boundaries. In order to exclude the

    latter possibility, single crystalline samples were used.

    The second objective was to better characterize CDRX.

    The influence of the following parameters was therefore

    investigated:

    crystalline orientation, by using single crystals of

    various orientations;

    purity, by testing three grades of polycrystalline alu-

    minium, ranging from 99.99 to 97 wt.% Al. Impuri-

    ties and alloying elements reduce the recoverycapacity of the material, since they decrease its stack-

    ing fault energy (although this effect seems to be

    relatively weak in aluminium alloys) and solutes

    as well as precipitates reduce the dislocation mobil-

    ity;

    strain path, by comparing the microstructures

    obtained from uniaxial compression and torsion

    tests.

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    3/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288276

    Table 1

    Chemical compositions of the three aluminium grades

    Mg Si Cr Mn Fe Cu Zn

    12 1 1 101199 (mg/g) 457 4

    1500 62 829 57001200 (mg/g) 59 200

    0.115052 (wt.%) 0.22.49 0.08 0.3 0.01

    2. Experimental procedure

    The three selected grades of aluminium, a pure 1199

    aluminium, a commercial purity 1200 aluminium and

    an AlMg 5052 alloy (Table 1), were provided by the

    Centre de Recherches de Voreppe (Pechiney) in the

    form of hot rolled plates. The initial structures were

    fully recrystallized, exhibiting equiaxed grains of 220

    mm in the 1199 aluminium, flattened grains of average

    size 200 mm in the 1200 aluminium, and again equiaxed

    grains of 80 mm in the 5052 alloy (Table 2). The

    textures displayed a strong cube component, especially

    in the 1200 grade. Cylindrical compression specimens

    were machined with their axes parallel to the rolling

    direction. In addition, torsion specimens were prepared

    from the 1200 aluminium with the torsion axis parallel

    to the rolling direction. Finally, single crystals were

    grown from the 1199 aluminium and sectioned to ob-

    tain cubic specimens with a 001, 011 or 111

    direction parallel to the compression axis (Table 3).

    Hot compression tests were carried out at a constant

    strain rate m;=102 s1 and 0.7 Tm (i.e. 380, 368 and

    333C for the 1199, 1200, and 5052 grades, respec-

    tively). The specimens were lubricated with graphite to

    minimize strain inhomogeneities, and water quenched

    within 1 s after deformation. The torsion specimenswere strained under the same strain rate and tempera-

    ture conditions, and water quenched within 5 s. The

    compression specimens were then sectioned parallel to

    the axis, ground and electropolished (5 ml HClO4, 95

    ml C2H5OH, 20 V, 0C, 60 s); some of them were

    subsequently anodized (10 ml HBF4, 90 ml H2O, 30 V,

    20C, 120 s). Since deformation was not uniform, local

    strains were estimated using a mechanical model [35].

    In what follows, the strain values correspond to the

    center part of each specimen, where all observations

    were carried out. However, even at large strains, the

    strain gradient was quite low, e.g. in the compression

    direction, Dm/Dz:0.2 mm1 at m=1.5 [35]. The tor-sion cylinders were ground to obtain a flat surface, and

    then prepared the same way. The strain and strain rate

    undergone by these samples were estimated to approxi-

    mately 80% of their nominal (surface) values.

    The deformed microstructures and textures were in-

    vestigated using polarized optical microscopy (POM)

    on the anodized specimens, scanning electron mi-

    croscopy (SEM) in the channeling contrast mode, elec-

    tron back-scattered diffraction (EBSD) and X-raydiffraction on the electropolished specimens. POM dis-

    plays both the initial grains and the new crystallites

    (subgrains or new grains), whereas only the crystallites

    are generally revealed by SEM. This is due to the fact

    that POM colors are related to the crystalline orienta-

    tions, and the original grains are associated with re-

    gions of similar colors. This is not the case in SEM,

    where crystallites of similar orientations can display

    quite different grey levels, and conversely. However,

    POM is known to overestimate subgrain sizes

    [31,36,37], and is therefore inadequate for quantitative

    analyses of the substructure. For this purpose, SEMmicrographs were therefore used. Moreover, the color

    contrast in POM is not precise enough to distinguish a

    new grain from a subgrain and even less to estimate the

    misorientation between two crystallites. The orientation

    of each crystallite was thus determined using EBSD and

    the misorientation associated with each boundary sub-

    sequently calculated. On account of detection limitation

    in hot worked structures, boundaries with a misorienta-

    tion angle of less than 1 were not taken into consider-

    ation. This omission is likely to be of no consequence,

    however, since the study is focused on the transition

    from low angle to high angle boundaries around 15. In

    addition to these local texture measurements, global

    textures were investigated by X-ray diffraction.

    Table 2

    Mean intercept lengths of the grains in the hot rolled plates along the

    rolling (RD), transverse (TD), and normal (ND) directions

    RD TD ND

    2261199 (mm) 204 227

    1200 (mm) 285 215 98

    8678 595052 (mm)

    Table 3Crystallographic directions parallel to the compression axis (CA) and

    perpendicular to the lateral faces (TD1 and TD2) of the cubic single

    crystals

    TD1 TD2CA

    [100][001] [010]

    [100] [011(][011]

    [112(][11(0][111]

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    4/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288 277

    Fig. 1. Stress-strain curves of the 001, 011, 111 single crystals

    and polycrystals of 1199 aluminium.

    3. Experimental results

    3.1. Pure aluminium single crystals

    During the compression tests, the cross-sections of the

    001 specimens remain square, as expected since the

    001 crystallographic axis and the geometric axis

    (square cross-section) are both fourfold, and the trans-

    verse directions TD1 and TD2 are crystallographicallyequivalent (Table 3). The 011 (twofold crystallo-

    graphic axis) specimens lengthen in the 100 direction.

    It has been shown [38] that this shape change can be

    explained by the activation of the four octahedral slip

    systems (111)[101(], (111)[11(0], (1(11)[1(01(], and (1(11)[1(1(0].

    Finally, the cross-sections of the 111 specimens be-

    come irregular, which is due to the combination of their

    threefold crystallographic axis and four-fold geometric

    axis.

    Fig. 1 shows that the flow stresses reach a steady state

    value of about 10.5 MPa in the case of the 001 and

    111 specimens (note that the polycrystal flow stress

    tends to the same value); by contrast, it is much higherfor the 011 specimens (about 14 MPa). It will be

    shown below that these values are closely related to the

    subgrain sizes. The flow stress drop associated with the

    111 crystals can be mainly attributed to the evolution

    of the Taylor factor: indeed, at m=0, M=33/2:3.67,while at m=1.5 (101 orientation, see below) M=6:2.45.

    Global texture measurements show that the 011

    orientation is perfectly stable (Fig. 2(b)). The 001

    orientation is also fairly stable, since its decomposition

    only starts at m=1.5 (Fig. 2(a)). On the other hand, the

    111 orientation is very unstable (Fig. 2(c)). At m=0.3,

    the compression axis is roughly parallel to 112 andeventually reaches the 011 stable orientation at m=

    1.5. These results are in good general agreement with the

    literature, although the rotation amplitudes are larger

    than those observed by Mecif et al. [39], after uniaxial

    compression of aluminium single crystals at the same

    temperature. This difference can be attributed, to a large

    extent, to the higher strain (m=1.5 vs 0.35) and strain

    rate (102 vs. 2104 s1) applied in the present work.

    The aspect of the sections perpendicular to the com-

    pression axis does not change significantly with increas-

    ing strain for the 100 specimens observed by POM

    (Fig. 3(a, b)). However, the band structure observed on

    the lateral sections at m=0.3 transforms into a subgrain

    structure at larger strains. It should be noted that SEM

    reveals the presence of many subgrains within these

    bands. The 011 specimens exhibit symmetrical mosaic

    patterns on their (011() lateral sections (Fig. 3(c, d)), with

    dislocation walls parallel to the planes of the

    activated slip systems. SEM also reveals the formation

    of small subgrains inside the cells at m=1.5. Such a

    mosaic microstructure of the 011 specimens (which

    Fig. 2. Global textures of the monocrystalline specimens strained to

    m=1.5. (a) 001; (b) 011; (c) 111. The horizontal and vertical

    axes are associated with the TD1 and TD2 directions, respectively

    (Table 3).

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    5/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288278

    elongate only in the [100] direction) has already been

    observed in the case of single crystals of various orien-

    tations deformed by channel die compression [40]. This

    suggests that such a particular structure is due to the

    geometry of the slip systems, which allows a unidirec-

    tional strain to take place. By contrast to the previous

    orientations, the 111 specimens exhibit an inhomoge-

    neous microstructure. Horizontal and tilted bands are

    displayed on the TD1 lateral sections at m=0.3. How-ever, this inhomogeneity tends to vanish at larger

    strains.

    Misorientation maps were plotted from EBSD mea-

    surements for each single crystal strained to m=1.5. In

    the 001 specimen (Fig. 4(a)), a large number of

    boundaries exhibit a misorientation greater than 15,

    which indicates that subgrain boundaries have trans-

    formed into grain boundaries. By contrast, the

    boundaries in the 011 specimen (Fig. 4(b)) exhibit

    low misorientations, generally smaller than 6, so that

    no high angle boundaries have formed. The behavior of

    the 111 specimen (Fig. 4(c)) is intermediate, since

    only a small fraction of the interfaces consists of high

    angle boundaries.

    The evolutions of the misorientation distributions

    with increasing strain are compared for the three orien-

    tations in Fig. 5. In the case of the 001 crystal, the

    average misorientation strongly increases with strain.

    At m=0.9, a significant fraction of misorientations al-

    ready exceeds 15, which clearly means that part of the

    low angle boundaries have transformed into large angle

    boundaries. At m=1.5, this trend is more pronounced,

    since almost 20% of the interfaces are now grainboundaries. However, the microstructural steady state

    is not yet attained, which suggests that a more recrys-

    tallized microstructure (with crystallites bounded

    mainly by large angle boundaries) could form at larger

    strains. By contrast, for the 011 orientation, all the

    measured misorientations are lower than 15, even at

    m=1.5. Furthermore, no evolution is noticeable be-

    tween m=0.9 and 1.5, and thus, the formation of grain

    boundaries is unlikely, even at larger strains. The be-

    havior of the 111 specimens is more complex. In

    particular, the specimen strained to m=0.9 has devel-

    oped a large amount of very high angle boundaries

    (30 60). This is due to the splitting of the initial

    orientation into two components (Fig. 6), which was

    Fig. 3. POM microstructures of the monocrystalline specimens. (a) 001, m=0.3; (b) 001, m=1.5; (c) 011, m=0.3; (d) 011, m=1.5.

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    6/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288 279

    Fig. 4. SEM micrographs and associated misorientation maps of the monocrystalline specimens strained to m=1.5. (a) 001; (b) 011; (c) 111.

    The vertical and horizontal axes are associated with the CA and TD 1 (001 or 011) or TD2 (111) directions, respectively (Table 3).

    observed in only one specimen of this orientation. It

    should also be pointed out that, in the case of this

    unstable orientation, large angle boundaries are formed

    in the early stages of the deformation: at m=0.3, they

    already represent more than 8% of the boundaries.

    3.2. Polycrystalline specimens

    The mechanical behavior of the various specimens

    obtained from the compression and torsion tests was

    investigated (see Fig. 1 for the 1199 grade). In all cases,

    the flow stress seems to reach a plateau at m:0.3

    (although a small decrease of the torsion stress is

    expected at very large strains [10,11]). The flow stresses

    reach steady state levels of 10.5, 21 and 77 MPa for the

    1199, 1200 and 5052 grades, respectively. Global tex-

    ture measurements were carried out on the compression

    specimens. The pole figures of the three grades (Fig. 7

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    7/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288280

    Fig. 5. Strain dependence of the misorientation distributions for the three single crystals (N: number of measurements, q(: average misorientation

    angle).

    Fig. 6. Formation of deformation bands in the 111 specimen strained to m=0.9.

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    8/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288 281

    Fig. 7. Global texture of the 1199 polycrystalline specimen strained to

    m=1.5. The horizontal and vertical axes are associated with the TD

    and ND directions, respectively (Table 2).

    All the microstructures are quite homogeneous. At

    m=0.3, subgrains are well formed in the 1199 and 1200

    grades (Fig. 8(a)), whereas they are not visible in the

    5052 grade, because their formation is delayed by the

    strong solute content of this alloy. Note also that at

    m=1.5, the grain thickness is still much larger than the

    subgrain size (Fig. 8(b)), which indicates that the initial

    grains are far from decomposing into smaller grains by

    a GDRX process. In Fig. 9, the mean intercept lengthsD of the subgrains measured from the SEM micro-

    graphs are compared with a theoretical value calculated

    from the flow stress |, using the relationship |/G=hb/

    D. The value of h=28 was chosen here, since it is

    commonly used for aluminium [34]. In the case of the

    1199 and 1200 grades, the calculated subgrain sizes are

    consistent with the measured ones, although slightly

    larger. The coefficient 28 seems therefore overestimated,

    and a better agreement between the two sets of data is

    obtained by using a value of 24. However, according to

    Blum et al. [34], this difference is due to the condensa-

    tion of free dislocations into additional subgrain

    boundaries between the end of deformation andquenching, which causes a decrease of the average size

    D. By contrast, the calculated subgrain size is much

    smaller than the experimental value for the 5052 alloy.

    This is due to the slow formation of the subgrains in

    that case, since the presence of regions without well-

    formed subgrains leads to an overestimation of the

    mean intercept length.

    Fig. 10 displays misorientation maps of the polycrys-

    tals strained to m=1.5. In the 1199 specimen (Fig.

    10(a)), two initial grain boundaries can still be iden-

    tified because they form a continuous chain of high

    misorientation segments (\30), whereas isolated high

    angle boundaries with lower misorientations (15 30)are very probably new grain boundaries. In the case of

    the 1200 aluminium strained in compression (Fig.

    10(b)), the map shows the presence of a large fraction

    Fig. 8. POM microstructures of the 1200 polycrystalline specimens.

    (a) m=0.3; (b) m=1.5. The compression axis is vertical.

    Fig. 9. |/G versus b/D plot for the single crystals and polycrystals

    strained to 1.5 (the shear moduli G=20.6, 20.8, 21.2 GPa for the

    1199, 1200, 5052 grades, respectively, at 0.7 Tm, and the Burgers

    vector length b=2.861010 m).

    illustrates the texture of the 1199 aluminium) show that

    the texture consists mainly of a 011 fiber component,

    which is especially strong for the 5052 alloy. Since a

    strong cube component was observed in the initial

    state, this means that the crystallites have strongly

    rotated during compression.

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    9/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288282

    Fig. 10. SEM microstructures and associated misorientation maps of the polycrystalline specimens strained to m=1.5. (a) 1199 specimen; (b) 1200

    compression specimen; (c) 1200 torsion specimen. The compression axis is vertical.

    of high angle boundaries, but the original grainboundaries are no longer recognizable. It is likely that

    subgrains have rotated more in the 1200 than in the

    1199 grade, thus leading to a fragmentation of the

    initial grain boundaries. In the torsion specimen of Fig.

    10(c), the fraction of high angle boundaries is very

    similar, although some chains of segments exhibiting

    very large misorientations at the top of the map look

    like initial grain boundaries. The misorientation distri-

    butions of the 1199 aluminium are depicted in Fig. 11.For this aluminium, as for the two other grades, there

    is a progressive shift of the small misorientations to-

    wards the larger ones, which leads to an increase of the

    average misorientation of about 8 between m=0.3 and

    m=1.5. The fraction of subgrain boundaries with small

    misorientations (B6) is strongly reduced; the number

    of subgrain boundaries with larger misorientations in-

    creases at first, and then decreases, whereas the fraction

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    10/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288 283

    of high angle boundaries continuously increases, reach-

    ing more than 30% at m=1.5. This indicates that the

    very low angle boundaries (B6) are continuously con-

    verted into medium angle boundaries (615), which in

    turn transform into high angle boundaries. Such an

    evolution may be referred to as continuous

    recrystallization.

    3.3. Comparison of the geometric and continuous

    dynamic recrystallization kinetics

    In the polycrystalline specimens, the increase of the

    high angle boundary fraction is due not only to the

    formation of new grain boundaries, but also to the

    expansion of the initial grain boundary area. In order

    to get a quantitative estimation of the respective effects

    of CDRX and GDRX, a geometrical model was used.

    The evolutions of the lengths per unit area of initial

    grain boundaries, new high angle boundaries, and low

    angle boundaries were determined from a simple

    derivation detailed in the Appendix A.Fig. 12(ae) illustrate the GDRX and CDRX kinet-

    ics. The strain dependence of the initial grain boundary

    fraction is not monotonic since it grows due to the

    flattening of the initial grains, whereas the formation of

    subgrain boundaries makes it decrease. At m=1.5, the

    initial grain boundaries represent about 8% of the

    interfaces for the 1200 and 5052 grades. The 1199

    aluminium differs by a much higher fraction, about

    17%, which is due to a smaller initial grain size/sub-

    grain size ratio. By comparing the thickness h of the

    initial grains at m=1.5 (the latter were estimated by

    calculation to 68, 80, and 33 mm for the 1199, 1200, and

    5052 grades, respectively) and the subgrain sizes D atthe same strain (14.3, 6.6 and 3.4 mm, respectively), it

    appears that the h/D ratios are about 5 for the 1199

    grade, and 1012 for the 1200 and 5052 grades. This

    confirms that GDRX is more developed in the pure

    aluminium. But this ratio still remains far from 2,

    which means that the strain achieved by compression is

    not large enough to obtain a complete GDRX

    structure.

    With regard to CDRX, the 001 single crystal (Fig.

    12(d)) and the polycrystal of same purity (and strong

    initial cube texture, Fig. 12(a)) display similar behav-

    iors: at m=1.5, roughly 1518% of their interfaces

    consist of new high angle boundaries. When deforma-

    tion bands form in the 111 crystal, the fraction of

    new grain boundaries rises very quickly; if not, it

    remains rather small, about 5 8%. Among the poly-

    crystalline specimens, the highest fraction of new grainboundaries is observed in the 1200 specimens (35% in

    compression, 39% in torsion at m=1.5), then in the

    5052 alloy (about 24%), and last in the pure aluminium

    (only 15%). This can be explained by the very high

    recovery rate in pure aluminium, which lowers the

    accumulation rate of dislocations in the subgrain

    boundaries. On the other hand, in the Al-Mg alloy, the

    solute atmospheres impede dislocation movements,

    thereby delaying the formation of subgrain boundaries

    [7]. It should be noted that the grain boundaries present

    in the 1199 aluminium originate in equal parts from

    GDRX and CDRX, while those present in the 1200

    and 5052 grades have mainly developed by CDRX. Thepresent results also indicate that the compression and

    torsion specimens display quantitatively the same be-

    havior. Therefore, the strain path does not seem to

    modify the CDRX kinetics significantly.

    4. Discussion

    4.1. New grain boundaries and deformation bands

    Experiments carried out on single crystals confirm

    that the subgrain boundary misorientations stronglyincrease with strain: a maximum of at least 15 is

    reached for the three investigated orientations at m=

    0.9. Moreover, the transformation of low angle

    boundaries into high angle boundaries is clearly demon-

    strated for the two unstable orientations, as well as in

    the polycrystalline specimens where the fraction of high

    angle boundaries is much larger than expected for the

    deformed initial grain boundaries.

    Fig. 11. Strain dependence of the misorientation distributions for the 1199 polycrystals ( N: number of measurements, q(: average misorientation).

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    11/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288284

    Fig. 12. Strain dependence of the surface fractions pertaining to the

    various kinds of interfaces, viz. low angle boundaries (LAB), new

    high angle boundaries (nHAB) [including the deformation band

    boundaries (DB)], and old initial high angle boundaries (oHAB). (a)

    1199 polycrystals; (b) 1200; (c) 5052; (d) 1199 001; (e) 1199 111.

    However, the misorientation increase of the subgrain

    boundaries and their transformation into high angle

    boundaries has been controversial for a long time,

    principally on the basis of the work by Kassner and

    McMahon [30]. These authors studied the microstruc-

    tural evolution of high purity (99.999 wt.%) polycrys-

    talline aluminium specimens strained in torsion (370C,

    5104 s1). The observations were mainly carried

    out by transmission electron microscopy, and the mis-orientations were estimated from either dislocation

    spacings (subgrain boundaries) or selected area diffrac-

    tion patterns (grain boundaries). These authors noticed

    only very limited increases in subgrain misorientations

    with strain: the mean values varied from 0.5 at m=0.2

    to a saturation value of 1.2 from m=1.2 to 16. Most of

    the discrepancies between this prior work and the

    present investigation can be explained by differences in

    the experimental procedures. First of all, the aluminium

    used by the authors was purer, and the strain rate,

    lower. This means that recovery was more efficient, and

    thus the increase in misorientation was slowed down.

    Furthermore, by contrast to the SEM-EBSD techniqueused here, TEM allowed to account for boundaries

    with very low misorientations (B1), therefore decreas-

    ing the average misorientation value. However, some

    boundaries with medium (about 10), and large misori-

    entations (\30) were also observed by Kassner and

    McMahon [30]. Since a limited number of measure-

    ments (about 20 for each specimen) were carried out,

    the authors misorientation distributions are not

    smooth and discontinuous values are observed within

    the range 515, instead of the continuous distributions

    displayed here (Figs. 5 and 11). This is a reason why

    these medium angle boundaries, which were also

    present in single crystalline specimens strained undersimilar conditions [41], were interpreted as interfaces

    between persistent deformation bands (although such

    bands were not clearly identified), and not as former

    subgrain boundaries transformed into new grain

    boundaries.

    It should be noted that the classification of the

    various boundary types has been developed in the case

    of cold deformation and its application to hot worked

    structures in not obvious. Indeed, at temperatures be-

    low 0.4 Tm, several kinds of interfaces are observed.

    The initial grains decompose into deformation bands

    separated by a thin (12 mm) transition band contain-

    ing dislocation cells. Inside the deformation bands,

    random low misorientation cells are grouped together

    in cell blocks separated by dense dislocation walls of

    higher misorientations [43]. However, as temperature

    increases, dislocation walls become thinner and a much

    more homogeneous microstructure develops. Generally,

    above 0.6 Tm, only equiaxed subgrains are observable

    inside the deformed initial grains. Even when a decom-

    position occurs, the deformation bands are not easily

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    12/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288 285

    observed on micrographs (Fig. 6). The reason is that

    they appear bounded by true grain boundaries, instead

    of the straight above-mentioned transition bands. Dif-

    ferentiation between several types of low and high angle

    boundaries is therefore difficult in hot worked struc-

    tures, and the relevance of such a classification even

    questionable.

    4.2. Influence of crystalline orientation

    Results obtained on single crystals confirm that the

    CDRX kinetics strongly depend on the crystallographic

    orientation. The fully stable 011 orientation enables

    only a limited increase of the misorientations. No high

    angle boundary creation was observed, although some

    misorientations are close to 15 at m=0.9. In the case of

    the quasi-stable 001 orientation, the formation of

    new high angle boundaries, with 15 30 misorienta-

    tions, was observed in the specimens strained to m=0.9

    and 1.5. For the 111 orientation, which is fully

    unstable, new grain boundaries in the same misorienta-

    tion range were observed, starting even at m=0.3. Suchresults are in total agreement with previous work on

    aluminium single crystals. Theyssier et al. [40] observed,

    after channel die compression of single crystals of the

    same purity, that some orientations (e.g., {112}111

    or {421}112) led to the formation of high angle

    boundaries, whereas other ones (e.g. {110}112) did

    not. More recently, Ponge et al. [6] investigated the

    misorientations developed during hot uniaxial compres-

    sion (260C, 4104 s1) of a very high purity alu-

    minium (99.9995 wt.%), which recrystallized by DDRX.

    In the unrecrystallized matrix, these authors observed

    the occurrence of smaller misorientations in the 011

    single crystal (B9), than in the 112 specimen (up to26). The misorientation ranges were thus very close to

    those obtained in the present work.

    4.3. Continuous dynamic recrystallization and grain

    rotations

    In order to clarify the origin of the new high angle

    boundaries, it is interesting to look more closely at the

    relationship between global texture and local misorien-

    tations. In the case of uniaxial compression, the follow-

    ing remarks can be made:

    1. Stable orientation. During straining, the global tex-

    ture of the 011 crystals remains unchanged, ex-

    cept that it becomes slightly less sharp. This is due

    to strain hardening and dynamic recovery: disloca-

    tions accumulate in the subgrain boundaries, thus

    altering the initial orientation. Misorientations up to

    15 have been reported in pure aluminium single

    crystals. The results obtained on polycrystals indi-

    cate that in a less pure aluminium, such as the 1200

    grade, the misorientations are probably larger, thus

    leading to the formation of high angle grain

    boundaries. A question still remains open: which

    mechanism leads to the increase in misorientation?

    If one considers that dislocations of opposite signs

    are created in equal densities during straining, each

    dislocation wall will absorb dislocations of both

    signs, keeping its misorientation at a low level. A

    misorientation increase can only occur if a subgrain

    boundary absorbs an excess of dislocations of onesign, which supposes that the various types of dislo-

    cation are not uniformly distributed in the material.

    This assumption does not seem unrealistic, however,

    all the more as misorientations between adjacent

    subgrains will increase such inhomogeneities. In-

    deed, slip system activities can be affected by small

    orientation changes. Furthermore, among the vari-

    ous orientations introduced by strain, those belong-

    ing to the 011 fiber are fully stable and will

    probably not disappear, thus leading to increased

    and permanent misorientations. A slight trend to-

    wards fiber formation in the pole figure of the 011

    crystal strained to m=1.5 can be noticed (Fig. 2(b)).If compression specimens could be deformed to very

    large strains, a 011 fiber texture would certainly

    be observed.

    2. Unstable orientations. New interfaces can be intro-

    duced by both strain hardening and lattice rota-

    tions. Two cases must be distinguished, according to

    whether deformation bands occur or not. Let us first

    consider the case in which the whole crystal rotates

    towards the same orientation. In the 111 single

    crystal strained to m=0.3, the EBSD local pole

    figure clearly shows that some orientations still re-

    main close to the initial one, whereas others are

    already located near the final one. That is why highangle boundaries are observed at such a low strain.

    However, at m=1.5, all the crystallites have reached

    their final 011 orientation, which explains the

    lower fraction of high angle boundaries. The subse-

    quent behavior has been described in (i). In the

    second case, when the initial orientation splits into

    symmetrical components, very large and permanent

    misorientations are rapidly built up. Deformation

    bands are expected to occur only for specific initial

    orientations, e.g. 001, 111, or the intermediate

    uuw orientations, and they are more likely to

    occur in single crystals than in polycrystals (except

    for very large initial grains).

    Lyttle and Wert [23] have formulated three models

    based on dislocation glide, boundary sliding, and neigh-

    bor switching to account for the increased misorienta-

    tions during straining of superplastic alloys. They

    concluded that combination of the boundary sliding

    model and the neighbor switching model most closely

    reproduced the misorientations measured experimen-

    tally. In the case of the dislocation glide model, the

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    13/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288286

    misorientations tended to decrease rather than increase,

    therefore reflecting texture development. However, it

    can be inferred from the above discussion that the

    convergence of the crystallite orientations and the cre-

    ation of large misorientations are not incompatible. A

    reason is that strain hardening tends to move the

    crystallites away from their ideal orientations. This

    effect was probably very pronounced in this case be-

    cause of the low recovery level associated with highalloying element content. Moreover, it was experimen-

    tally found that, for some reason, the lattice rotation

    rate varied from one crystallite to another, thus intro-

    ducing high misorientations during the transient. All

    these disturbing factors were not taken into account in

    the model. The increase in subgrain boundary misorien-

    tation by accumulation of dislocations was probably

    predominant in the early steps of straining. However,

    since the crystallite size was very small, grain boundary

    sliding and grain switching probably prevailed as soon

    as large enough misorientations were built up.

    4.4. Kinetics of the continuous dynamic recrystallization

    It is worth noting that the various mechanical and

    structural parameters tend to their steady state values

    at different rates. The flow stress remains approxi-

    mately constant from m=0.3, except for the 111

    crystal. In this case, the crystallographic rotation delays

    the steady state up to m=0.9. The subgrain size is

    generally well established at m=0.3. However, a sub-

    grain refinement is still observed in the 5052 alloy at

    m=1.5. By contrast, the misorientations are not yet

    stabilized at m=1.5, with the only exception of the

    011 crystal. This is due to the combined effects of the

    increasing misorientations of the subgrains and newlyformed grain boundaries on the one hand, and the

    increasing surfaces of the original grain boundaries on

    the other hand. Apart from some minor cases, the flow

    stress and the average subgrain size thus reach their

    steady state values quite rapidly (m=0.3), while the

    average misorientation is still increasing at m=1.5.

    These experimental results are thus in contradiction

    with the similitude principle, which stipulates that the

    various microstructural spacings are inversely propor-

    tional to the flow stress. For instance, McQueen and

    Blum [42,44] proposed that there is a unique relation-

    ship between the average dislocation spacing s in the

    subgrain boundaries, (or, equivalently, the average sub-

    grain boundary misorientation q:b/s) and the flow

    stress. The fact that the average subgrain boundary

    misorientation increases without affecting the flow

    stress significantly can be explained if one considers, as

    established by several authors [45,46], that the strength-

    ening associated with dislocations inside the subgrains

    is larger than that due to dislocations in the subgrain

    boundaries. The former is thus the main controlling

    parameter of the flow stress. The assumption of Mc-

    Queen and Blum was based mostly on creep data from

    specimens deformed to low or moderate strains, which

    can explain why these authors did not observe any

    misorientation increase of the subgrain boundaries.

    4.5. Elementary mechanisms of the continuous dynamic

    recrystallization

    Continuous dynamic recrystallization has sometimes

    been labeled extended dynamic recovery, either because

    some authors restrict the term recrystallization to the

    classical discontinuous recrystallization, or because the

    microstructures generally consist of crystallites only

    partially bounded by high angle boundaries. This termi-

    nology is somewhat misleading, however, since dynamic

    recovery can have detrimental as well as beneficial

    effects on CDRX. Indeed, the above experimental ob-

    servations indicate that the CDRX process results from

    the combination of three elementary mechanisms:

    1. The formation of subgrain boundaries. These

    boundaries are created with a very low misorienta-tion angle (about 1), as a result of dynamic

    recovery.

    2. The transformation of subgrain boundaries into

    grain boundaries. The increase in misorientation of

    the subgrain boundaries is more or less rapid, ac-

    cording to the material and the experimental condi-

    tions. It was shown that it is accelerated by medium

    recovery levels and when the initial orientation is

    unstable. If such favorable conditions are brought

    together, the subgrain boundaries can be gradually

    transformed into grain boundaries.

    3. The elimination of subgrain and grain boundaries.

    Measurements carried out on a b titanium alloy[13,47] and some aluminium alloys have recently

    shown that the grain boundaries migrate, even in

    the absence of classical DDRX, although at much

    lower rates. All the interfaces present in the volume

    swept by the migrating boundaries disappear, which

    certainly plays a major role in the establishment of

    the steady state, for both GDRX and CDRX. It has

    been shown for instance that, in compression, the

    initial grains can reach a quasi-steady state thick-

    ness, which is an increasing function of the

    boundary velocity [13,47]. Moreover, the elimina-

    tion of interfaces allows the subgrain size and the

    misorientation distribution to stabilize after a tran-

    sient period.

    From the previous experimental observations, a

    model of CDRX has recently been proposed, in which

    the above three mechanisms are combined in order to

    predict the microstructural evolutions [38,48]. The main

    features are well reproduced: during the transient, the

    subgrain size decreases while the misorientation grows,

    and the subgrain boundaries are gradually converted

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    14/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288 287

    into grain boundaries; ultimately, the crystallite size

    reaches a steady state value, as well as the misorienta-

    tion distribution, although at larger strains.

    5. Conclusions

    Various aluminium specimens were submitted to uni-axial compression and torsion testing at 0.7 Tm and

    102 s1, up to m=1.5. The main results obtained after

    investigation of the hot worked structures are the

    following.

    (1) Experiments carried out on 001, 011 and

    111 single crystals confirm that the subgrain

    boundary misorientations strongly increase with strain.

    For the three investigated orientations, the largest val-

    ues are close to or beyond 15. In the case of the

    stable 011 orientation, the increase in misorien-

    tation is not sufficient to transform low angle into

    high angle boundaries. Among the two other orienta-

    tions, the conversion occurs earlier for the most un-stable one: it starts at m=0.3 in the unstable 111

    crystal, but only at m=0.9 in the metastable 001

    crystal.

    (2) Most of the new high angle boundaries have

    1530 misorientations. They originate either from dif-

    ferences in the rotation rate towards the final orienta-

    tion, or from fluctuations around the average

    orientation introduced by strain hardening and dy-

    namic recovery. In some cases, the initial orientation

    splits into symmetrical components, causing the occur-

    rence of deformation bands. In this case, very high

    misorientations (\30) are rapidly built up.

    (3) In the case of polycrystals, the high angleboundary fractions strongly increase with strain. Since

    both continuous and geometric dynamic recrystalliza-

    tion are likely to occur, calculations were carried

    out in order to estimate the surface fraction of the

    initial grain boundaries. It turns out that the high angle

    boundary fraction is much larger than could be ac-

    counted for by the expanded initial boundaries, thus

    confirming the presence of many new high angle

    boundaries.

    (4) Continuous dynamic recrystallization is more effi-

    cient in the commercial purity aluminium than in the

    pure aluminium and the Al-Mg alloy. This

    suggests that the transformation of low angle

    boundaries into high angle boundaries is faster when

    the recovery level is neither too high (the accumulation

    rate of dislocations in the subgrain boundaries de-

    creases), nor too low (the formation of subgrains is very

    slow or they do not form at all). Comparisons between

    compression and torsion specimens also indicate that

    the strain path does not alter CDRX kinetics notice-

    ably.

    Acknowledgements

    The authors are indebted to Professor J.J. Jonas,

    McGill University, Montreal, for providing access to

    the torsion facility. They are also grateful to Professor

    H.J. McQueen, Concordia University, Montreal, for

    many fruitful discussions. The work of S. Gourdet was

    supported in part by the Region Rhone-Alpes, France,

    through a scientific fellowship (Avenir program).

    Appendix A

    The evolution of the fraction of the various interface

    types (i.e. low angle boundaries, initial and new high

    angle boundaries) is addressed here in the case of

    compression. Similar equations apply for torsion (see

    Ref. [38] for more details). Since experimental measure-

    ments provide interface lengths per unit area, a two-di-

    mensional approach was chosen. Initial grains and

    subgrains are approximated by ellipses with semiaxes

    a=y/4 DCA and b=y/4 DTD, where DCA and DTDdenote the measured intercept lengths parallel to the

    compression axis and the transverse direction,

    respectively.

    The mean intercept lengths of the initial grains can

    be accurately measured only at m=0. Their evolutions

    with strain are evaluated by taking into account the

    flattening of the grains and the migration of the initial

    boundaries, which leads to an increase of the average

    thickness [13,38]. Along the direction parallel to the

    compression axis:

    D:CA=DCAm;+26 (1)

    where the migration rate 6=0.1 mm/s at m;=0.01 s1

    [13,38]. This yields after integration:

    DCA=(D0DS) exp(m)+DS (2)

    where D0 is the initial length and DS=26/m;, or, for the

    semiaxis:

    ag=(a0aS) exp(m)+aS (3)

    where aS=y6/2m;. Moreover, perpendicular to the com-

    pression axis:

    bg=b0 exp(m/2) (4)

    (in the above equations, a0 and b0 are the semiaxes

    lengths at m=0).

    In addition, the initial grain boundary length is af-

    fected by the presence of serrations. Measurements

    from POM micrographs showed that the length ratio k

    between a serrated and a straight boundary first in-

    creases with strain (as subgrains form) and then

    remains approximately constant. This evolution can be

    accurately described by the equation k=1+

    0.25 [1 exp( 4m)] . The current length per unit

  • 7/29/2019 An Experimental Study of the Recrystallization Mechanism During Hot Deformation of Aluminium

    15/15

    S. Gourdet, F. Montheillet /Materials Science and Engineering A283 (2000) 274288288

    area of the initial grain boundaries is then given by

    kPg/2yagbg, where the ellipse perimeter Pg is calculated

    numerically from the lengths of the semiaxes ag and bg.

    Since the subgrain semiaxes asg and bsg can be mea-

    sured at various strains, the total interface length per

    unit area is simply given by Psg/2yasgbsg. The total

    (old+new) high angle grain boundary length is then

    estimated by multiplying the total interface length by

    the fraction of high angle boundaries measured byEBSD (see the misorientation distributions). Finally,

    the length per unit area of the new high angle

    boundaries is obtained by difference.

    References

    [1] H. Yamagata, Scripta Metall. Mater. 27 (1992) 201 203.

    [2] H. Yamagata, Scripta Metall. Mater. 27 (1992) 727 732.

    [3] H. Yamagata, Scripta Metall. Mater. 27 (1992) 1157 1160.

    [4] H. Yamagata, Scripta Metall. Mater. 30 (1992) 411 416.

    [5] H. Yamagata, Acta Metall. Mater. 43 (1992) 723 729.

    [6] D. Ponge, M. Bredehoft, G. Gottstein, Scripta Mater. 37 (1997)

    17691775.

    [7] C. Chovet, S. Gourdet, F. Montheillet, Int. Conf. UHPM 98,

    France, pp. 6774.

    [8] F.R. Castro-Fernandez, C.M. Sellars, Mat. Sci. Tech. 4 (1988)

    621627.

    [9] T. Sheppard, M.G. Tutcher, Met. Sci. 14 (1980) 579 589.

    [10] C. Perdrix, M.Y. Perrin, F. Montheillet, Mem. Sci. Rev. Metall.

    78 (1981) 309320.

    [11] F. Montheillet, in: P. Costa, et al. (Eds.), 24eme Colloque de

    Metallurgie, INSTN, Saclay, 1981, pp. 57 70.

    [12] S. Gourdet, E.V. Konopleva, H.J. McQueen, F. Montheillet,

    Mat. Sci. Forum 217 (1996) 441446.

    [13] S. Gourdet, A. Girinon, F. Montheillet, in: T. Chandra, T. Sakai

    (Eds.), Int. Conf. Thermec 97, The Minerals, Metals and Mate-

    rials Society, Wollongong, Australia, 1997, pp. 21172124.

    [14] F. Chaussy, J.H. Driver, in: A. Vassel, D. Eylon, Y. Combres(Eds.), Les Alliages de Titane b, vol. 8, Editions de la Revue de

    Metallurgie, Paris, 1994, pp. 57 64.

    [15] A.-M. Chaze, F. Montheillet, in: A. Vassel, D. Eylon, Y. Com-

    bres (Eds.), Les Alliages de Titane b, vol. 8, Editions de la Revue

    de Metallurgie, Paris, 1994, pp. 41 48.

    [16] F. Chaussy, J.H. Driver, Rev. Metall.-CIT/Sci. Genie Mater. 93

    (1996) 10571066.

    [17] L. Lombry, C. Rossard, B.J. Thomas, Rev. Metall./CIT 78

    (1981) 975988.

    [18] C.G. Schmied, C.M. Young, B. Walser, R.H. Klundt, O.D.

    Sherby, Metall. Trans. 13A (1982) 447456.

    [19] A. Belyakov, R. Kaibyshev, T. Sakai, Metall. Trans. 29A (1998)

    161167.

    [20] P. Cizek, B.P. Wynne, Mater. Sci. Eng. A230 (1997) 88 94.

    [21] R.H. Bricknell, J.W. Edington, Metall. Trans. 10A (1979) 1257

    1263.

    [22] R.H. Bricknell, J.W. Edington, Acta Metall. 27 (1979) 1303

    1311.

    [23] M.T. Lyttle, J.A. Wert, J. Mater. Sci. 29 (1994) 3342 3350.

    [24] H. Gudmundson, D. Brooks, J.A. Wert, Acta Metall. Mater. 39

    (1991) 1935.

    [25] E.-W. Lee, T.R. McNelley, Mater. Sci. Eng. 93 (1987) 45 55.

    [26] S.J. Hales, T.R. McNelley, Acta Metall. 36 (1988) 1229 1239.

    [27] T.R. McNelley, E.-W. Lee, M.E. Mills, Metall. Trans. 17A(1986) 10351041.

    [28] E.-W. Lee, T.R. McNelley, A.F. Stengel, Metall. Trans. 17A

    (1986) 10431050.

    [29] H.J. McQueen, O. Knustad, N. Ryum, J.K. Solberg, Scripta

    Metall. 19 (1985) 7378.

    [30] M.E. Kassner, M.E. McMahon, Metall. Trans. 18A (1987) 835

    846.

    [31] J.K. Solberg, H.J. McQueen, N. Ryum, E. Nes, Phil. Mag. 60A

    (1989) 447471.

    [32] G.A. Henshall, M.E. Kassner, H.J. McQueen, Metall. Trans.

    23A (1992) 881889.

    [33] W. Blum, Q. Zhu, R. Merkel, H.J. McQueen, Mater. Sci. Eng.

    205A (1996) 2330.

    [34] W. Blum, Q. Zhu, R. Merkel, H.J. McQueen, Z. Metallkde 87

    (1996) 1423.

    [35] B. Dumanowski, Thesis, 1997, Ecole des Mines de Saint-Etienne,

    France.

    [36] C. Perdrix, F. Montheillet, G. Wyon, F. Weill, Metallography 13

    (1980) 289297.

    [37] Q. Zhu, H.J. McQueen, W. Blum, High Temp. Mat. Proc. 17

    (1998) 289297.

    [38] S. Gourdet, Thesis, Ecole des Mines de Saint-Etienne, France,

    1997.

    [39] A. Mecif, B. Bacroix, P. Franciosi, Acta Mater. 45 (1997)

    871881.

    [40] M.-C. Theyssier, B. Chenal, J.H. Driver, N. Hansen, Phys. Stat.

    Sol. (a) 149 (1995) 367378.

    [41] M.E. Kassner, Metall. Trans. 20A (1989) 2182 2185.

    [42] H.J. McQueen, W. Blum, ICAA6, Japan Inst. Metals, pp. 99

    112.

    [43] J.H. Driver, 16th Ris Int. Symp., Danemark, 1995, pp. 25 34.[44] H.J. McQueen (section 9, R.D. Doherty, D.A. Hughes, F.J.

    Humphreys, J.J. Jonas, D. Juul Jensen, M.E. Kassner, W.E.

    King, T.R. McNelley, H.J. McQueen, A.D. Rollett), Mat. Sci.

    Eng. A238 (1998) 219274.

    [45] M.E. Kassner, A.K. Miller, O.D. Sherby, Metall. Trans. 13A

    (1982) 19771986.

    [46] F.R. Castro-Fernandez, C.M. Sellars, Phil. Mag. 60A (1989)

    487506.

    [47] A. Girinon, Memoire CNAM, 1999, Conservatoire National des

    Arts et Metiers, Saint-Etienne, France.

    [48] S. Gourdet, C. Chovet, F. Montheillet, in: T. Sakai, H.G. Suzuki

    (Eds.), 4th Int. Conf. on Recrystallization and Related Phenom-

    ena, The Japan Institute of Metals, Tsukuba, 1999, pp. 259264.

    .