the effect of solutes on grain boundary mobility during recrystallization and grain growth in some...

9
Materials Chemistry and Physics 132 (2012) 166–174 Contents lists available at SciVerse ScienceDirect Materials Chemistry and Physics jo u rn al hom epage : www.elsevier.com/locate/matchemphys The effect of solutes on grain boundary mobility during recrystallization and grain growth in some single-phase aluminium alloys Y. Huang a,, F.J. Humphreys b a BCAST, Brunel University, Kingston Lane, Uxbridge, Middlesex UB8 3PH, UK b Manchester Materials Science Centre, The University of Manchester, Grosvenor Street, Manchester M1 7JL, UK a r t i c l e i n f o Article history: Received 8 July 2011 Accepted 11 November 2011 Keywords: Alloys Annealing Electron microscopy Interfaces Thermodynamic properties. a b s t r a c t The effect of solutes (Si, Mn, Mg) in quantities typical of commercial aluminium alloys, on grain boundary mobility in aluminium, has been investigated with in situ annealing and electron backscattered diffraction in the SEM, and grain growth experiments. The in situ experiments provided information on the migration of the high mobility tilt boundaries of misorientations close to 40 1 1 1. Grain growth experiments were used to investigate boundary migration in alloys of high solute content (1–5wt%Mg), and a comparison between the in situ and bulk experiments is made. The relationship between boundary velocity and driving pressure was found to be linear in all cases, and the activation energies for boundary migration were higher than those controlled by lattice diffusion of the solutes at higher solute concentrations. © 2011 Elsevier B.V. All rights reserved. 1. Introduction It has long been recognised that solute elements segregate to grain boundaries in order to lower the energy and that dur- ing boundary migration, the solute atmosphere diffuses with, but lags behind the boundary, resulting in a dragging pressure on the boundary. Even very small additions of solute elements to high purity metals can drastically reduce the mobility (M) of bound- aries, e.g. [1–3]. In order to understand and control the processes of recovery, recrystallization and grain growth during thermome- chanical processing it is important that the effects of solute drag are quantified. Although the general effects of solute on bound- ary migration have long been known, quantitative measurements have been made only on a limited number of systems. Careful research carried out on curvature-driven boundary migration in bi-crystals of known orientation has clearly demonstrated the effects of boundary character, orientation and purity on boundary mobility, particularly at high temperature [4–8]. However, recent work [9] has suggested that there may be differences in boundary behaviour during recrystallization and curvature-driven growth, which could be attributed to differences in the type and magnitude of the driving pressure, the lower temperatures involved and the fact that a recrystallization front generally comprises small areas of boundary of different character, rather than a single boundary type. The prime reason for the present work was to assess the effect of the solutes of the type and quantity typically found in Corresponding author. Tel.: +44 0 1895266976; fax: +44 0 1895269758. E-mail address: [email protected] (Y. Huang). industrial aluminium alloys on boundary migration during recrys- tallization. Such a study should provide some of the measurements of boundary mobilities which are required as inputs to models of recrystallization, e.g. [10–13]. It should however be emphasised that it is much more difficult to interpret the results of such an investigation in terms of the fundamental physics of boundary migration than is the case for the fundamental studies on curva- ture driven growth referred to above, in which only trace additions of solute were made. 2. Experimental 2.1. Materials High purity (HP) aluminium and a series of high purity based single-phase alu- minium alloys were investigated in this study. The compositions of the alloys are given in Table 1. 2.2. Specimen preparation Cylindrical single crystals of size 16 × 100–140 mm were grown from the melt in a graphite mould under a vacuum of 10 3 Pa from 7 mm diameter cylindrical seed crystals of 1 0 0 orientation, and the crystal orientations were determined from X-ray diffraction to an accuracy of ±1 . The crystals were carefully cut and pol- ished into Goss-oriented {1 1 0}0 0 1 rectangular specimens of dimensions (4.5, 6.0, 10) mm × 10 mm × 15 mm and then given the homogenisation heat-treated speci- fied in Table 2. The crystals were cooled after homogenisation to 300 C at a rate of 10 C h 1 and then water quenched, except the high purity Al, which was cooled to room temperature at 10 C h 1 . The rectangular single crystals of Goss orientation were deformed in plane strain compression on an Instron 4045 test machine using a stainless steel channel-die. The deformation was carried out at room temperature and at a strain rate of 10 3 s 1 for all materials to reductions (strain): 30% (0.36), 50% (0.69) and 70% (1.2). In order to reduce friction, PTFE tape, which was replaced at an increment of 20% reduction, was used as a lubricant. 0254-0584/$ see front matter © 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.matchemphys.2011.11.018

Upload: y-huang

Post on 10-Sep-2016

214 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The effect of solutes on grain boundary mobility during recrystallization and grain growth in some single-phase aluminium alloys

Tg

Ya

b

a

ARA

KAAEIT

1

tilbpaocaahrbemwbwofb

e

0d

Materials Chemistry and Physics 132 (2012) 166– 174

Contents lists available at SciVerse ScienceDirect

Materials Chemistry and Physics

jo u rn al hom epage : www.elsev ier .com/ locate /matchemphys

he effect of solutes on grain boundary mobility during recrystallization andrain growth in some single-phase aluminium alloys

. Huanga,∗, F.J. Humphreysb

BCAST, Brunel University, Kingston Lane, Uxbridge, Middlesex UB8 3PH, UKManchester Materials Science Centre, The University of Manchester, Grosvenor Street, Manchester M1 7JL, UK

r t i c l e i n f o

rticle history:eceived 8 July 2011ccepted 11 November 2011

a b s t r a c t

The effect of solutes (Si, Mn, Mg) in quantities typical of commercial aluminium alloys, on grain boundarymobility in aluminium, has been investigated with in situ annealing and electron backscattered diffractionin the SEM, and grain growth experiments. The in situ experiments provided information on the migration

eywords:lloysnnealinglectron microscopy

of the high mobility tilt boundaries of misorientations close to 40◦〈1 1 1〉. Grain growth experiments wereused to investigate boundary migration in alloys of high solute content (1–5wt%Mg), and a comparisonbetween the in situ and bulk experiments is made. The relationship between boundary velocity anddriving pressure was found to be linear in all cases, and the activation energies for boundary migrationwere higher than those controlled by lattice diffusion of the solutes at higher solute concentrations.

nterfaces

hermodynamic properties.

. Introduction

It has long been recognised that solute elements segregateo grain boundaries in order to lower the energy and that dur-ng boundary migration, the solute atmosphere diffuses with, butags behind the boundary, resulting in a dragging pressure on theoundary. Even very small additions of solute elements to highurity metals can drastically reduce the mobility (M) of bound-ries, e.g. [1–3]. In order to understand and control the processesf recovery, recrystallization and grain growth during thermome-hanical processing it is important that the effects of solute dragre quantified. Although the general effects of solute on bound-ry migration have long been known, quantitative measurementsave been made only on a limited number of systems. Carefulesearch carried out on curvature-driven boundary migration ini-crystals of known orientation has clearly demonstrated theffects of boundary character, orientation and purity on boundaryobility, particularly at high temperature [4–8]. However, recentork [9] has suggested that there may be differences in boundary

ehaviour during recrystallization and curvature-driven growth,hich could be attributed to differences in the type and magnitude

f the driving pressure, the lower temperatures involved and theact that a recrystallization front generally comprises small areas of

oundary of different character, rather than a single boundary type.

The prime reason for the present work was to assess theffect of the solutes of the type and quantity typically found in

∗ Corresponding author. Tel.: +44 0 1895266976; fax: +44 0 1895269758.E-mail address: [email protected] (Y. Huang).

254-0584/$ – see front matter © 2011 Elsevier B.V. All rights reserved.oi:10.1016/j.matchemphys.2011.11.018

© 2011 Elsevier B.V. All rights reserved.

industrial aluminium alloys on boundary migration during recrys-tallization. Such a study should provide some of the measurementsof boundary mobilities which are required as inputs to models ofrecrystallization, e.g. [10–13]. It should however be emphasisedthat it is much more difficult to interpret the results of such aninvestigation in terms of the fundamental physics of boundarymigration than is the case for the fundamental studies on curva-ture driven growth referred to above, in which only trace additionsof solute were made.

2. Experimental

2.1. Materials

High purity (HP) aluminium and a series of high purity based single-phase alu-minium alloys were investigated in this study. The compositions of the alloys aregiven in Table 1.

2.2. Specimen preparation

Cylindrical single crystals of size �16 × 100–140 mm were grown from the meltin a graphite mould under a vacuum of 10−3 Pa from 7 mm diameter cylindricalseed crystals of 〈1 0 0〉 orientation, and the crystal orientations were determinedfrom X-ray diffraction to an accuracy of ±1◦ . The crystals were carefully cut and pol-ished into Goss-oriented {1 1 0}〈0 0 1〉 rectangular specimens of dimensions (4.5, 6.0,10) mm × 10 mm × 15 mm and then given the homogenisation heat-treated speci-fied in Table 2. The crystals were cooled after homogenisation to 300 ◦C at a rate of10 ◦C h−1 and then water quenched, except the high purity Al, which was cooled toroom temperature at 10 ◦C h−1.

The rectangular single crystals of Goss orientation were deformed in plane strain

compression on an Instron 4045 test machine using a stainless steel channel-die. Thedeformation was carried out at room temperature and at a strain rate of 10−3 s−1 forall materials to reductions (strain): 30% (0.36), 50% (0.69) and 70% (1.2). In order toreduce friction, PTFE tape, which was replaced at an increment of ∼20% reduction,was used as a lubricant.
Page 2: The effect of solutes on grain boundary mobility during recrystallization and grain growth in some single-phase aluminium alloys

Y. Huang, F.J. Humphreys / Materials Chemistry and Physics 132 (2012) 166– 174 167

Fig. 1. Partly recrystallized sample of Al–0.1%Mg: (a) Electron backscattered image; (b) EBSD map of the same area; (c) 1 1 1 pole figure of all the grains in the map. Thecluster of orientations close to the Goss orientation are due to the recovered material; (dto the recovered matrix.

Table 1Compositions of the alloys.

Alloy Compositions, wt%

Si Fe Cu Mn Mg Zn Ti B Al

HP–Al <.01 <.01 .001 <.001 <.001 <.001 <.001 <.001 BalAl–0.05Si .05 <.01 .001 <.001 <.001 <.001 <.001 <.001 BalAl–0.05Mn <.01 <.01 .001 .05 <.001 <.001 <.001 <.001 BalAl–0.3Mn <.01 <.01 .001 .30 <.001 <.001 <.001 <.001 BalAl–0.1Mg <.01 <.01 .001 <.001 0.13 <.001 <.001 <.001 BalAl–1Mg <.01 <.01 .001 <.001 0.91 <.001 <.001 <.001 BalAl–3Mg <.01 <.01 .001 <.001 3 <.001 <.001 <.001 Bal

Ba

odp

2

t

influenced by the deformation close to the scratch, the boundary velocity was found

TT

Al–5Mg <.01 <.01 .001 <.001 5 <.001 <.001 <.001 Bal

old values in Table 1 represent the amount (in weight percentage) of the purposelydded individual solute element in the corresponding binary aluminium alloy.

Disk specimens were spark-cut from the deformed single crystals, with the planef the disc parallel to the compression (rolling) plane. The disk specimens wereeeply etched to remove any damage from the spark erosion, and mechanicallyolished and electropolished to a final size of �3 mm × 1 mm.

.3. In situ annealing experiments

The in situ annealing experiments were carried out using similar procedures tohose of an earlier investigation [9]. A single fine scratch was initially made across a

able 2he homogenisation treatment of the single crystals.

HP–Al Al–0.05Mn Al–0

Temperature (◦C) 630 600 580

Time (h) 4 5 12

) inverse pole figure showing the misorientation axes of the fastest growing grains

disc sample, which was mounted in a hot stage in a CAMSCAN MX4000 field emissiongun scanning electron microscope (FEGSEM). The disk was then heated to a temper-ature of 270–280 ◦C and held for ∼10 min to promote recovery of the dislocationsto form subgrain structures. The disk was then heated to the required temperature(between 300 ◦C and 420 ◦C) and isothermally annealed. Recrystallization was ini-tiated at the scratch (Fig. 1a), and the experiment was continued until grains hadgrown a distance of ∼500 �m from the scratch.

The partly recrystallized disk was then cooled to room temperature and EBSDmapping was carried out using an HKL Channel EBSD system. In some cases wherethe size of recrystallized grains was suitable for further annealing, the disk wasannealed again at a different (generally lower) temperature and measurements weremade on the same boundaries as those at the previous higher temperature. Afterannealing at the second temperature, the subgrain size and misorientation in therecovered matrix were again measured. During in situ annealing, the microstruc-tural evolution was monitored using backscattered electron imaging. A U-maticvideo recorder was used to record the microstructural changes in samples whereboundary migration was too rapid for normal image recording.

2.4. Microstructural characterisation and EBSD data analysis

The rate of growth of the recrystallizing grains was measured in the rollingplane. After the first stages of growth (∼50 �m) in which the recrystallization was

to remain almost constant, although a slight decrease due to recovery in the matrixwas sometimes observed [9].

Because all the recrystallized grains originated close to the surface at the scratch,it is reasonable to assume that measurements of boundary velocity in the disc plane

.3Mn Al–0.05Si Al–0.1Mg Al–1Mg

580 600 5804 5 12

Page 3: The effect of solutes on grain boundary mobility during recrystallization and grain growth in some single-phase aluminium alloys

168 Y. Huang, F.J. Humphreys / Materials Chemistry and Physics 132 (2012) 166– 174

Fsa

aateti

ttfdat

V

e

2

ssmdtcnma

gAt

aaroA

bpcwda

3

3

ctaaa

0

0.2

0.4

0.6

0.8

1

10 20 30 40 50 60

Me

an

sp

ec

ific

bo

un

dary

ve

loc

ity

Misorientation ang le(°)

hp-AlAl-0.05MnAl-0.3MnAl-0.05 SiAl-0.13 MgAl-0.91 Mg

deviation from<111 >~15

0

0.2

0.4

0.6

0.8

1

0 20 40

Me

an

sp

eic

fic

bo

un

dary

ve

loc

ity

Deviation from <11 1>(°)

hp-AlAl-0.05MnAl-0.3MnAl-0.05 SiAl-0.13 MgAl-0.91 Mg

misorientaion35-45

(a)

(b)

ig. 2. Electron backscattered image of the cross section of a partly recrystallizedample of Al–1%Mg showing limited interaction between the migrating grain bound-ry and the specimen surface.

re representative of the true boundary velocity. In order to check this assumptionnd to assess the influence of the free surface on the recrystallization, partly recrys-allized samples were sectioned normal to the disc plane and metallographicallyxamined. In most cases, the migrating boundary was approximately perpendicularo the specimen surface as shown in Fig. 2, indicating that the free surface had littlenfluence on the boundary migration.

The EBSD data in the form of orientation maps (e.g. Fig. 1b), were analysed usinghe post-processing software, VMAP [14]. The subgrain size and misorientation inhe recovered matrix were calculated, and from these, the driving pressures (P)or recrystallization was calculated as discussed in Section 3.4. Correlation of theriving pressures with the boundary velocities (V) measured just before the in situnnealing was stopped, enabled the boundary velocities (M) to be determined fromhe relationship:

= MP (1)

The crystallographic relationships between recrystallized grains and the recov-red matrix were also determined from the EBSD data.

.5. Grain growth experiments on bulk samples

It was of interest to measure the boundary mobilities of alloys containing largeolute concentrations, and Al–Mg, which is of industrial importance, is an ideal alloyystem. However, it was not possible to carry out in situ recrystallization experi-ents on alloys with larger concentrations of aluminium than 1%. This was mainly

ue to the difficulty in determining accurately the driving pressures for recrystalliza-ion from the deformed and recovered microstructures in alloys containing largeroncentrations of magnesium, and in which well-recovered subgrain structures mayot be formed before recrystallization. It was also thought that the evaporation ofagnesium at the surface via bulk or grain boundary diffusion might result in surface

rtefacts.Therefore boundary migration in these alloys was investigated by measuring

rain growth rates in polycrystals of Al–3%Mg and Al–5%Mg in addition to thel–1%Mg alloy used for the in situ experiments. By using the Al–1%Mg alloy for both

ypes of experiment, it was also possible to compare results from the two methods.The 50 mm thick cast ingots were homogenised at 500 ◦C and slow cooled

t 10 ◦C h−1 to room temperature. They were then hot rolled to 20 mm slab andnnealed for 2 h at 360 ◦C for Al–1Mg, 400 ◦C for Al–3Mg and 440 ◦C for Al–5Mgespectively. The hot rolled and annealed slabs were then cold rolled to a reductionf 60%, and annealed at 360 ◦C for 1 h. The average grain sizes were Al–1Mg: 64 �m,l–3Mg: 92.1 �m, Al–5Mg: 58.7 �m.

For the grain growth experiments, samples were annealed in air at temperaturesetween 300 ◦C and 450 ◦C. The annealed samples were sectioned in the ND–RDlane, metallographically prepared and examined in the FEGSEM using channellingontrast imaging and EBSD. The grain size, misorientation and their distributionsere determined from the reconstructed EBSD data using VMAP. Grain sizes wereetermined from the grain area as an equivalent circle diameter (ECD), which, forn equiaxed grain structure is 0.816 of the diameter of spherical grains.

. Results of in situ experiments

.1. General features of the annealing

The Goss orientation is stable during channel die plane strainompression [15] and the resultant subgrain structure in which

here are neither short range nor long-range orientation gradientsnd therefore no sites for originating recrystallization, providesn ideal matrix into which the recrystallized grains originatingt the scratch, can grow. The microstructure in the deformed and

Fig. 3. The mean specific boundary velocity as a function of (a) misorientation angleand (b) deviation from a 〈1 1 1〉 axis.

recovered matrix was a uniform subgrain structure, for details see[15,16], and the range of microstructural parameters of the variousrecovered materials is given in Table 3.

The basic features of recrystallization and grain boundary migra-tion were found to be similar for all the materials, and are typifiedby those shown in Fig. 1. As many aspects of the recrystallizationbehaviour are qualitatively similar to those reported in an earlierinvestigation on Al–0.05%Si [9], we will not discuss them in detail,but will concentrate on the quantitative effects of the alloying ele-ments.

Recrystallization originated along the scratch. The initial grainswere found to have a reasonably random orientation relationshipwith the matrix, and the distribution of misorientation axes for thespecimen of Fig. 1 is shown in Fig. 1c. However, a few grains areseen to have grown more rapidly and anisotropically.

3.2. Grain shape

The fast growing grains appeared to be lenticular, and thiswas confirmed by microstructural examination after progressive

removal of the surface layers. The fastest growth direction of therapidly growing grains was found to be perpendicular to the axis ofmisorientation between the grains and the recovered matrix, show-ing that the most mobile boundaries were of tilt character, whereas
Page 4: The effect of solutes on grain boundary mobility during recrystallization and grain growth in some single-phase aluminium alloys

Y. Huang, F.J. Humphreys / Materials Chemistry and Physics 132 (2012) 166– 174 169

Table 3Typical microstructural parameters for the recovered specimens.

Al(5N) Al–0.05Si Al–0.05Mn Al–0.3Mn Al–0.1Mg Al–1Mg

True strain 0.36–1.4 0.36–1.2 0.36–1.2 0.36–1.3 0.36–1.2 0.36–1.2Subgrain size (�m) (mean intercept) 1.8–4.8 2.4–4.3 2.1–4.2 1.5–3.4 1.5–3.78 1.5–4.5

tlmgmagpsa

3

htwiadtaowabpgawobbp

dmbhai∼a

sg

TT

Mean subgrain misorientation ( ) 1.5–2.5 1.7–2.6

Orientation spread (◦) ∼10 ∼8

Net driving pressure (MPa) 0.03–0.26 0.01–0.13

he twist boundaries along the broad faces of the grains were theeast mobile. For those grains whose rotation axes were approxi-

ately parallel to the sample surface, both the fastest and slowestrowth directions are parallel to the sample surface and thereforeeasurement of the grain aspect ratio gives a good measure of the

nisotropy of the boundary growth rate. As shown in Table 4, therowth anisotropy ratio was found to be ∼10 and to be almost inde-endent of the solute type and concentration. However, it may beeen that there is some small increase with increasing strain for thelloy crystals, and a larger effect for the high purity aluminium.

.3. Orientation dependence of growth

The fast growing grains in all the materials were found toave orientation relationships with the recovered matrix closeo 40◦〈1 1 1〉 and no other fast growing orientation relationshipsere detected except for the HP–Al in which the fastest grow-

ng grains had a 28–38◦〈1 0 0〉 relationship to the recovered matrixt recrystallization temperatures below 150 ◦C. Fig. 3a shows theependence of the boundary velocity of the fast growing grains onhe misorientation angle to the matrix for grains of misorientationxes <15◦ from 〈1 1 1〉. Fig. 3b shows the dependence of velocityn the deviation of the misorientation axis from 〈1 1 1〉, for grainshose misorientation to the matrix is 35–45◦. In Fig. 3, the bound-

ry velocity is expressed as a ratio of the velocity of a particularoundary relative to that of the fastest growing grain in the sam-le. Note that the measurements of the 28–38◦〈1 0 0〉 misorientedrains found at annealing temperatures below 150 ◦C in high purityluminium are excluded from Fig. 3, but will be discussed else-here. Fig. 3 shows that although the maximum boundary mobility

ccurs close to a 40◦ rotation about a 〈1 1 1〉 axis as has commonlyeen observed during recrystallization [9,17,18], the peaks are veryroad, and Table 5 shows the half-height half-widths of the mobilityeaks for the various materials.

The half widths of the mobility peaks as a function of angulareviation (Fig. 3a) are seen to be ∼10◦, and generally similar for allaterials. Although there is scatter in the data, some trends may

e seen. The Al–0.3Mn peak is the narrowest (9.1◦) and that of theigh purity aluminium the widest (11.2◦), with those of the otherlloys being intermediate and similar. The half widths of the mobil-ty peaks as a function of axis deviation from 〈1 1 1〉 (Fig. 3b) are also10◦. However in this case, the Al–0.3Mn peak is broadest (13.4◦)

nd that of the high purity aluminium is sharpest (7.3◦).

The broad mobility peaks of Fig. 3 are in contrast to the veryharp peaks (<1◦) found for boundaries moving by curvature drivenrowth, e.g. [3,7]. However, the broad peaks are similar to those

able 4he average aspect ratio of fast-growing grains.

Reduction (%) Aspect ratio

HP–Al Al–0.05Mn Al–0.3

30 8.8 9.6 9.2

50 11.3 9.5 9.6

70 14.7 11.6 9.3

Average 11.6 10.2 9.4

1.5–2.8 1.8–2.9 1.6–2.75 1.6–2.6∼9 ∼11 ∼7 ∼90.03–0.21 0.035–0.3 0.038–0.2 0.035–0.2

found in an earlier investigation of recrystallization [9], where itwas suggested that they were attributable to the effect of boundaryexcess volume rather than to the crystallography of high coinci-dence boundaries.

3.4. Grain boundary velocity and mobility

The velocities of the fast growing grains in their fast growingdirection (i.e. the tilt boundaries of the ∼40◦〈1 1 1〉 misorientedgrains) were measured from the in situ annealing experiments forvarious driving pressures and temperatures as described earlier.The driving pressure (PD ) is provided by the recovered subgrains,and is given by [9]

PD = 2�m

L

�m

(1 − ln

�m

)(2)

where �m is the energy of a high angle grain boundary (taken as0.324 J m−2 [19]), � is the low angle misorientation, �m is the mis-orientation above which the boundary energy become independentof misorientation (taken as15◦) and L is the mean linear subgrainintercept. Such a relationship, which is based on the Read–Shockleyrelationship, is valid only if the low angle boundaries are well recov-ered (i.e. not cell walls), and if the dislocation density within thesubgrains is low, and this has been verified.

The migrating boundaries have a curvature (R), and this providesa retarding pressure (PC) given, e.g. [2] by

PC = 2�m

R(3)

and therefore the resultant pressure on the boundary (P) is PD − PC.PD was calculated from Eq. (2) using values of � and L determined

from EBSD of the unrecrystallized microstructure after the in situannealing, and PC was calculated from the measured curvatures ofthe growing grains. PC was found to be typically ∼0.02 MPa and tovary only slightly during the growth of the grains.

The variation of boundary velocity with temperature and netdriving pressure (P) for the various materials is shown in Fig. 4. Thedata show some scatter, but give reasonable fits to a straight lineas predicted by Eq. (1). As was found in a previous investigation[9], there appears to be a small threshold pressure (∼0.02 MPa)for migration which varies only slightly with material and tem-

perature, and whose origin is unclear. It could be associated withinteraction of the migrating boundaries with the surface by groov-ing or oxide film formation, although as shown in Fig. 2, there islittle evidence of such interaction.

Mn Al–0.05Si Al–0.1Mg Al–1Mg

9.5 8.2 8.511.6 10.7 9.510.5 10.5 10.6

10.5 9.8 9.5

Page 5: The effect of solutes on grain boundary mobility during recrystallization and grain growth in some single-phase aluminium alloys

170 Y. Huang, F.J. Humphreys / Materials Chemistry and Physics 132 (2012) 166– 174

) HP–

dit

Fig. 4. The correlation between boundary velocity and driving pressure: (a

The boundary mobilities were calculated from Eq. (1), using theata from Fig. 4, and the effect of temperature on mobility is shown

n Fig. 5. It is seen that all the solute elements have strongly reducedhe boundary mobilities in the alloys and the mobility values for the

Al, (b) Al–0.05Si, (c) Al–0.1Mg, (d) Al–1Mg, (e) Al0.05Mn and (f) Al–0.3Mn.

alloys are relatively similar compared to the large differences to themobilities of the high purity aluminium. However, if the mobilitiesof the alloys alone are plotted on an expanded scale as in Fig. 5b,differences between the alloys are clearly seen. Fig. 5 shows that

Page 6: The effect of solutes on grain boundary mobility during recrystallization and grain growth in some single-phase aluminium alloys

Y. Huang, F.J. Humphreys / Materials Chemistry and Physics 132 (2012) 166– 174 171

Table 5Analysis of the growth rate peaks of Fig. 3.

Material Misorientation at themaximum rate (◦)

The half width of the rate peak –function of angle of misorientation (◦)

The half width of the rate peak – function ofaxis deviation (◦) from 〈1 1 1〉

HP–Al 40 34.3–45.5 (11.2◦) 7.3Al–0.05Mn 40.4 35.5–46 (10.5◦) 8.7Al–0.3Mn 39 34.7–43.8 (9.1◦) 13.4

(10◦)

(9.5◦)(9.6◦)

tt

M

at

4

cmfb

F2

Al–0.05Si 41.9 37.1–47.1

Al–0.1Mg 40.6 36.1–45.6

Al–1Mg 41 36.6–46.2

he variation of the mobilities with temperature is a reasonable fito the equation

= M0 exp(−Q

RT

)(4)

nd the calculated activation energies (Q) and the pre-exponentialerm (M0) for each material are listed in Table 6.

. Results of bulk grain growth experiments

For the reasons discussed in Section 2.5, boundary migration

ould not be investigated by in situ annealing in alloys of highagnesium content. However, some information could be obtained

rom bulk grain growth experiments, and for the Al–1%Mg alloy,oth techniques were employed for comparison.

1E-14

1E-13

1E-12

1E-11

1E-10

1E-9

1E-8

1 1.5 2 2.5 3 3.5

1000/T, K-1

Mo

bil

ity

, m

4J

-1s

-1

5N-Al<100>5N-Al<11 1>Al-0.05MnAl-0.3MnAl-0.05SiAl-0.13MgAl-0.91Mg

(a)

1E-13

1E-12

1E-11

1E-10

1E-9

1E-8

1.4 1.5 1.6 1.7 1.8

1000/T, K-1

Mo

bil

ity

, m

4J

-1s

-1

Al-0.05Mn

Al-0.3Mn

Al-0.05Si

Al-0.13Mg

Al-0.91Mg

(b)

ig. 5. Temperature dependence of mobility, (a) all data including those of8–38◦〈1 0 0〉 boundaries in the HP–Al and (b) data only for the alloys.

8 9.3 9

4.1. Grain growth kinetics and boundary mobility

The material processing routes (Section 2.5) produced polycrys-tals with weak textures which were similar in all cases. The dataare summarised in Table 6, where it is seen that the only significanttexture component is S {1 2 3}〈6 3 4〉. The distribution of bound-ary angles was similar in all cases, and an example is shown inFig. 6 where it can be seen that the distribution is rather similarto that predicted for a randomly oriented grain assembly [20], butwith a slightly larger frequency of boundaries with misorientationsbetween ∼10◦ and 20◦.

Grain growth was observed in all the bulk annealed samples. Thegrain structures remained equiaxed and no abnormal growth wasdetected. As seen from Table 7, there were no significant changesto the texture or grain boundary distribution during grain growth,and the standard deviation of the grain size relative to the meangrain size remained almost constant. The results discussed in Sec-tion 3 referred to the migration rates of boundaries close to the40◦〈1 1 1〉 or �7 relationship. In the bulk annealed samples, thenumber of such boundaries was small and the length percentagesof �7 boundaries (defined according to the Brandon criterion) wasbetween 0.6% and 0.8%, with no significant change during graingrowth.

The isothermal grain growth kinetics are shown in Fig. 7. Thegood fit of the data to straight lines indicates that the variation ofmean grain size (D) with time (t) follows the well known relation-ship, see e.g. [2,3]

D2 − D2 = kt (5)

o

where k is a constant equal to ˛M� , where ∼ 2 [2].It is interesting that this relationship, which is predicted

for an ideal grain assembly (equiaxed grains with identical

0

20

40

60

80

100

120

140

160

0 10 20 30 40 50 60

Fre

qu

en

cy

Misorientation (°)

Fig. 6. A typical misorientation distribution in the Al–Mg alloys used for graingrowth experiment. The data are from Al–5%Mg annealed at 400 ◦C for 1 h.

Page 7: The effect of solutes on grain boundary mobility during recrystallization and grain growth in some single-phase aluminium alloys

172 Y. Huang, F.J. Humphreys / Materials Chemistry and Physics 132 (2012) 166– 174

Table 6Parameters of grain boundary migration obtained in this investigation.

Material Driving pressure (MPa) Temperature (◦C) Orientation (±10◦) Q (kJ mol−1) M0 (M4 J−1 s−1)

HP–Al 0.03–0.15 50–170 28∼38◦/〈1 0 0〉 66 0.0022Al(5N) 0.03–0.15 100–340 40◦/〈1 1 1〉 87 0.15Al–0.05Si 0.01–0.12 295–400 40◦/〈1 1 1〉 151 184Al–0.1Mg 0.05–0.23 300–370 40◦/〈1 1 1〉 136 3.16Al–1Mg 0.05–0.2 300–400 40◦/〈1 1 1〉 171 376Al–1Mga 0.0032–0.016 300–450 Random 170 367Al–3Mga 0.0034–0.012 300–450 Random 189 3533Al–5Mga 0.0052–0.018 300–450 Random 196 3715Al–0.05Mn 0.05–0.24 300–400 40◦/〈1 1 1〉 175 44,355

bbidotdaT

5

dvrgcsmdaastrsdb

5

bbmia

TT(

Al–0.3Mn 0.05–0.32 300–400

a Measured by grain growth during bulk annealing of recrystallized structures.

oundary energies in a single-phase alloy) [21] is often found not toe obeyed [2]. However, as the grain growth kinetics in the present

nvestigation clearly follow Eq. (5), the growth rates may be used toetermine the boundary mobilities using this equation and the dataf Fig. 7. The effect of temperature on the mobility of boundaries inhe Al–Mg alloys is shown in Fig. 8, where it may be seen that theata are a good fit to Eq. (4). The calculated activation energies (Q)nd the pre-exponential term (M0) for the materials are shown inable 6.

. Discussion

The boundary velocities were found to be proportional to theriving pressures for all the conditions investigated, confirming thealidity of Eq. (1) under the range of conditions investigated. Thisesult supports the recent Monte Carlo simulation outcomes thatrain boundary mobility is independent of the driving force underertain conditions [22,23] and casts doubts on the widely acceptedolute drag theory [24–26], which implies that grain boundaryobility is a function of driving force. In addition, there is no evi-

ence in the materials and conditions of the present investigation of transition to high boundary velocity associated with the break-way of a boundary from its solute atmosphere predicted by theolute drag theory. The driving pressures and annealing tempera-ures in the present work are typical of those encountered duringecrystallization of aluminium, and the type and concentration ofolutes are similar to some industrial alloys. This suggests thaturing the recrystallization of most aluminium alloys, boundaryreakaway (if there is any) is unlikely to play a significant role.

.1. Comparison of in situ and bulk experiments

Fig. 8 and Table 6 show that for the 1%Mg alloy, for which theoundary mobilities were measured by both in situ annealing and

ulk grain growth experiments, there is a large difference in theobilities. Although the activation energies are similar, the mobil-

ties at any temperature measured from the in situ experimentsre larger by a factor of ∼10. The reason for this is that the in situ

able 7he initial microstructures and textures of the Al–Mg alloys used for grain growth expAl–3Mg) and 180 min (Al–5Mg) when the grain sizes are approximately double the initia

D (�m) (sd/D) � (◦)

Al–1Mg Initial 64 (0.55) 33.2

Annealed 151 (0.59) 33.4

Al–3Mg Initial 92 (0.54) 33.8

Annealed 165 (0.51) 34

Al–5Mg Initial 58 (0.48) 35.1

Anneal 115 (0.51) 35.5

40◦/〈1 1 1〉 212 37,731,279

experiments measure the mobilities of tilt boundaries with ori-entation relationships close to 40◦〈1 1 1〉 to the recovered matrix(M�7), whereas the grain growth experiments measure the mobil-ities of “random” boundaries (MR). Previous measurements of theratio M�7:MR from in situ [9] or bulk [16] experiments show it tobe ∼10 and to be close to the aspect ratio of the 40◦〈1 1 1〉 plateletgrains. As seen from Table 4, the average value of the aspect ratiofor the 1%Mg alloy is 9.5, and if the in situ measurements in Fig. 8are divided by this factor so as to obtain the mobility of “random”boundaries, the mobilities obtained by in situ and bulk methods arecomparable as shown by the crosses in Fig. 8.

5.2. Effect of temperature on mobility

The behaviour of the high purity aluminium alloy was quite dif-ferent to the alloys, and the boundary mobility was not only muchlarger, but showed two regimes, both of low activation energy, inwhich boundaries of different type grew most rapidly. The mobili-ties of boundaries in the alloys appear to be controlled by the samethermally activated process throughout the temperature rangeinvestigated. It is seen from Table 6 that the reduction of mobility inthe alloys with increasing solute is not due to a decrease in M0 butto an increase of both M0 and Q, and as a result, the effect of soluteson mobility is less at higher temperatures. This is in agreement withthe experiments of Gordon and Vandermeer [27] on Al–Cu alloys.Recent work of the Aachen group [3,7] has clearly demonstratedthat for boundaries of a particular structure, M0 and Q are oftenrelated by the relationship

Q = ˛ ln(M0) + (6)

where and are constants. The boundary mobilities mea-sured in the present work (Table 6), apart from the Al–0.3Mn

alloy, are in agreement with such a relationship, with = 6.7 and

= 119 kJ mol−1. This coupling of M0 and Q, which is often referredto as the “compensation effect”, can be explained in terms of theo-ries of thermally activated boundary migration [3,7,28].

eriments, compared with those annealed at 400 ◦C for 30 min (Al–1Mg), 120 minl ones.

HAGB (%) % material within 15◦ of important texturecomponents (no other components were >5%)

Brass S

77.6 5.4 10.278.4 5.1 9.480.5 6.3 8.981.2 5.5 9.583.3 6.7 10.785.4 6.2 10.3

Page 8: The effect of solutes on grain boundary mobility during recrystallization and grain growth in some single-phase aluminium alloys

Y. Huang, F.J. Humphreys / Materials Chemistry and Physics 132 (2012) 166– 174 173

0

40000

80000

120000

160000

0 10000 0 20000 0 30000 0 40000 0

Annealing time (sec )

D2-D

02 (

µm2)

300°C

350°C

400°C

450°C

Al-1Mg

0

40000

80000

120000

0 20000 0 40000 0 60000 0 80000 0

Annealing ti me (se c)

(D2-D

02)

( µm

)2

300°C

350°C

400°C

450°C

Al-3Mg

0

8000

16000

24000

32000

40000

0 2000 00 40000 0 60000 0 80000 0

Annealing time (sec )

D2-D

02 (

µm2)

300°C

350°C

400°C

450°C

Al-5Mg

(a)

(b)

(c)

Fig. 7. Isothermal grain growth kinetics of Al–Mg polycrystals plotted according toEq. (4): (a) Al–1Mg, (b) Al–3Mg, (c) Al–5Mg.

5

saF

1.E-16

1.E-14

1.E-12

1.E-10

1.E-08

1.3 1.4 1.5 1.6 1.7 1.8

1000/T, K-1

Mo

bil

ity

, m

4J

-1s

-1

Al-0.1Mg(in-situ)

Al-1Mg(in-situ)

Al-1Mg

Al-3Mg

Al-5Mg

Al-1Mg (in-situ)/9.5

Fig. 8. Temperature dependence of mobility in Al–Mg alloys. Data from bulk grain

results showed much larger reductions in mobility with increasingmagnesium concentration as shown in Fig. 9. The calculated acti-vation energies at high magnesium concentrations (>1%) are muchlarger than the values obtained from lattice diffusion controlled

1E+7

1E+9

1E+11

1E+13

0.01 0.1 1 10

Conce ntration (at%)

(Mo

bil

ity

)-1,

m-4

J1s

1

Al0.0 5Si

Al-0.1 Mg

Al-1Mg (in-situ)

Al-0.05 Mn

Al-0.3 Mn

300°C

350°C

400°C

Fig. 9. Grain boundary mobility (random boundaries) as a function of the concen-

.3. Effect of magnesium concentration on boundary mobility

Due to the limited data available, a direct correlation betweenolute concentration and mobility could only be made for Al–Mglloys. The data from the Al–Mg alloys, which is shown in

ig. 9, shows that the relationship between mobility (M) and

growth experiments are represented by solid shapes and the in situ results by openshapes. The crosses represent the mobilities of “random” boundaries in Al–1Mgcalculated from the in situ data as discussed in the text.

concentration (X) at a given temperature can be described by therelationship

ln M = aX + b

where a = 0.7332 (300 ◦C), 0.5989 (350 ◦C) and 0.05512 (400 ◦C),b = 29.08 (300 ◦C), 26.267 (350 ◦C) and 23.95 (400 ◦C) respectively.

Fully quantitative theories of the effects of solutes on the mobil-ity of grain boundaries have not yet been developed. For the caseof boundaries whose migration rate is controlled by the diffusionof a solute atmosphere, it is predicted [3,24–27] that

M = q

X(8)

where q is a constant and relationships close to this have been foundfor metals containing trace additions of solute, see [2,3,28], and inparticular, by the work of Gordon and Vandermeer [26]. The present

tration of Mg in Al–Mg alloys.

Page 9: The effect of solutes on grain boundary mobility during recrystallization and grain growth in some single-phase aluminium alloys

174 Y. Huang, F.J. Humphreys / Materials Chemi

1E+8

1E+10

1E+12

1E+14

1E+16

0 1 2 3 4 5 6

Conce ntration (at%)

(Mo

bil

ity

)-1,

m-4

J1s

1

300°C

350°C

400°C

Al-Mg alloys

Fr

mtia

5

tmtdwteatTfibsetots

tpstcc

6

1

[[

[[[[[[[

[

[[[[

[[[[[

[

ig. 10. The effect of solute type and concentration on boundary mobility (in situesults only).

igration process [29,30]. Unfortunately, there is little experimen-al data with which the current results can be compared, as mostnvestigations of the effects of solute on the mobilities of bound-ries in aluminium have considered only trace additions.

.4. The effect of different solutes on boundary mobility

The in situ annealing experiments measured the mobilities ofilt boundaries close to 40◦〈1 1 1〉 for a number of solutes: silicon,

agnesium and manganese, over a limited range of concentra-ions. The migration of boundaries under the influence of soluterag is expected to be affected by a number of solute parameters,hich will depend on the specific solute and solvent. In addition to

he solute concentration which is discussed above, the interactionnergy between solute and boundary, the shape of the solute profilend the diffusivity of the solute atoms are all predicted to be impor-ant parameters in determining boundary mobility [3,23,31,32].he boundary character is also particularly important, and as shownrst by the classic work of Aust and Rutter [4,5] the mobilities ofoundaries with misorientations close to coincidence site relation-hips are less affected by solute than other boundaries. There isxperimental evidence, see e.g. [3] that for trace additions of soluteo aluminium, different elements may have greatly different effectsn boundary mobility. However, even for very low solute concen-rations it is difficult to quantitatively predict the effect of a specificolute on boundary mobility.

The present investigation is concerned with large solute addi-ions to aluminium. Attempts to correlate boundary mobility witharameters such as solubility and solute diffusivity have not beenuccessful, and, as shown in Fig. 10, for the in situ experiments,he boundary mobilities correlate reasonably well with soluteoncentration alone, suggesting that for these elements and con-entrations, other factors are secondary.

. Conclusions

. In-situ recrystallization experiments on deformed aluminiumcrystals containing large solute additions have shown that the

[

[

[

stry and Physics 132 (2012) 166– 174

orientation dependence of tilt boundaries with misorientationsclose to, but deviating from 40◦〈1 1 1〉 is dependent on the natureof the solute.

2. Grain growth experiments have been used to measure the mobil-ities of “random” boundaries in a range of Al–Mg polycrystallinealloys. The results can be correlated with those from the in siturecrystallization studies.

3. A relationship between the activation energies and pre-exponential terms similar to that found for more dilute alloyshas been found.

4. The mobility of boundaries in the Al–Mg alloys decreases morerapidly with solute concentration than the inverse relationshippredicted by early theories.

5. The mobilities of alloys containing Mg, Si or Mn solutes werefound to depend primarily on solute concentration and temper-ature, the specific solute type having limited influence.

Acknowledgements

The authors would like to acknowledge the financial support ofthe EPSRC for this project and Alcan International, Banbury, for theprovision of materials, facilities and helpful discussions.

References

[1] D.A. Smith, C.M.F. Rae, C.R.M. Grovenor, in: R.W. Balluffi (Ed.), Grain-BoundaryStructures and Kinetics, ASM, Metals Park, OH, 1979, p. 337.

[2] F.J. Humphreys, M. Hatherly, Recrystallization Related Annealing Phenomena,Pergamon Press, Oxford, 1995.

[3] G. Gottstein, L.S. Shvindlerman, Grain Boundary Migration in Metals: Thermo-dynamics, Kinetics Applications, CRC Press, Boca Raton, FL, 1999.

[4] K.T. Aust, J.W. Rutter, Trans. AIME 215 (1959) 119.[5] K.T. Aust, J.W. Rutter, Ultra-High Purity Metals, ASM, Metals Park, OH, 1962, p.

115.[6] B.B. Rath, H. Hu, Trans. AIME 245 (1969) 1577.[7] D.A. Molodov, U. Czubayko, G. Gottstein, L.S. Shvindlerman, Acta Mater. 46

(1998) 553.[8] M. Winning, G. Gottstein, L.S. Shvinderlerman, Acta Mater. 49

(2001) 211.[9] Y. Huang, F.J. Humphreys, Acta Mater. 47 (1999) 2259.10] E. Nes, in: N. Nansen (Ed.), Proc. 21st Riso Symp. Riso, 2000, p. 139.11] G. Gottstein, V. Marx, R. Sebald, in: T. Sakai, H. Suzuki (Eds.), Proc. 4th Inter.

Conf. of Recrystallization, Japan Inst. Metals, Tsukuba, 1999, p. 15.12] A.L. Garcia, V. Tikare, E.A. Holm, Scr. Mater. 55 (2008) 661.13] H. Hallberg, M. Wallin, M. Ristinman, Mater. Sci. Eng. A 527 (2010) 1126.14] F.J. Humphreys, J. Mater. Sci. 36 (2001) 3833.15] M. Ferry, F.J. Humphreys, Acta Mater. 44 (1996) 1293.16] Y. Huang, F.J. Humphreys, Acta Mater. 48 (2000) 2017.17] B. Liebmann, K. Lücke, G.Z. Masing, Metallk 47 (1956) 47.18] G. Ibe, K. Lücke, in: Margolin (Ed.), Recrystallization, Grain Growth and Tex-

tures, ASM, Metals Park, OH, 1966, p. 434.19] L.E. Murr, Interfacial Phenomena in Metals and Alloys, Addison-Wesley, Read-

ing, 1975.20] J.K. Mackenzie, Biometrika 45 (1958) 229.21] W.W. Mullins, J. Vinals, Acta Metall. 37 (1989) 991.22] L. Zhang, T. Bartel, M.T. Lusk, Comput. Mater. Sci. 48 (2010) 790.23] A.E. Lobkovsky, A. Karma, M.I. Mendelev, M. Haataja, D.J. Srolovitz, Acta Mater.

52 (2004) 285.24] J.W. Cahn, Acta Metall. 10 (1962) 789.25] K. Lücke, K. Detert, Acta Metall. 5 (1957) 628.26] K. Lücke, H. Stüwe, Acta Metall. 19 (1971) 1087.27] P. Gordon, R.A. Vandermeer, Trans. AIME 224 (1962) 917.28] D.A. Molodov, B.B. Straumal, L.S. Shvinderlerman, Sov. Phys. Solid State 26

(1984) 207.29] L.F. Mondolfo, Aluminium Alloys, Butterworth, London, 1976.

30] C.J. Smithells, Smithells Metals Reference Book, Butter-Heinemann, Oxford,

1992.31] A.P. Sutton, R.W. Balluffi, Interfaces in Crystalline Materials, Oxford University

Press, Oxford, 1995.32] A. Lens, C. Maurice, J.H. Driver, Mater. Sci. Eng. A 403 (2005) 144.