Sie sind auf Seite 1von 10

MAT E RI A LS CH A R A CTE R IZ A TI O N 7 4 (2 0 1 2) 11 0

Available online at www.sciencedirect.com

www.elsevier.com/locate/matchar

Thermal stability of friction stir processed ultrafine grained


Al\Mg\Sc alloy
N. Kumar1 , R.S. Mishra
Center for Friction Stir Processing and Department of Materials Science and Engineering, Missouri University of Science & Technology, Rolla,
MO 65409, USA

AR TIC LE D ATA

ABSTR ACT

Article history:

Two different aspects of the thermal stability of grains were studied (i) thermal stability

Received 23 May 2012

of grain structure evolution during friction stir processing (FSP), and (ii) thermal stability of

Received in revised form

grains after FSP. It was concluded that Al3(Sc,Zr) dispersoids are effective in limiting grain

30 August 2012

growth to ultrafine grained regime (UFG) by stabilizing the microstructure during FSP. The

Accepted 1 September 2012

presence of these dispersoids before processing was more effective in refining the grain size
than the case in which these dispersoids precipitated during FSP. The mean grain sizes for

Keywords:

these cases were 0.40 0.17 m and 0.50 0.25 m, respectively. Isothermal grain growth

EBSD

study revealed that UFG grain structure was maintained up to 723 K even after 16 h of

Thermal stability

annealing whereas most of the UFG alloys reported in the literature show extensive grain

Friction stir processing

growth above 473 K. To understand the role of Al3(Sc,Zr) dispersoids in this alloy,

Grain refinement

5086Al\H32 was friction stir processed and subjected to identical annealing conditions.

Grain growth

Annealing of FSP 5086Al\H32 at 623 K for 1 h resulted into abnormal grain growth (AGG).
The onset of AGG was observed for UFG sample only after annealing at 823 K for 1 h. This
was rationalized using Humphreys' model for AGG [F.J. Humphreys, Acta Mater., 45 (1997)
5031].
2012 Elsevier Inc. All rights reserved.

1.

Introduction

In the past two decades several grain refinement techniques


[15] have been developed to take advantage of HallPetch
strengthening [6,7]. Polycrystalline materials having grain size
below 1 m (ultrafine grained (UFG) and nanocrystalline) have
opened up entirely new domain of research due to their
difference in physical, chemical, and mechanical behaviors as
compared to coarse grained (CG) materials [8,9]. But, as grain
size decreases there is an increase in grain boundary area per
unit volume. Assuming a cube shaped grain it can be shown
that 50% reduction of grain leads to an increase in surface area
by ~ 16% and a reduction of 99% can cause surface area to

increase by ~ 3267% [10]. If grain boundary energy per unit


area of a material is assumed to be independent of its grain
size, there will be corresponding increase in total grain
boundary energy per unit volume. Hence, grain refinement
will always lead to an increase in the free energy of the system
thereby making the microstructure unstable. Such grains
when subjected to thermal cycle have tendency to grow to
minimize their energy by reducing the grain boundary area
per unit volume.
Thermal instability of the nanocrystalline [1115] and UFG
[1624] materials have been an issue and studied by many
researchers. Thermal stability of equal channel angular pressing (ECAP) processed UFG Al\3wt.%Mg was studied by Wang et

Corresponding author at: Center for Friction Stir Processing and Department of Materials Science and Engineering, University of North
Texas, Denton, TX 76203, USA. Tel.: + 1 940 565 2316; fax: + 1 940 565 4824.
E-mail addresses: Rajiv.Mishra@unt.edu (N. Kumar), Nilesh.Kumar@unt.edu (R.S. Mishra).
1
Present address: Center for Friction Stir Processing and Department of Materials Science and Engineering, University of North Texas,
Denton, TX 76203, USA.
1044-5803/$ see front matter 2012 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.matchar.2012.09.003

MAT ER I A LS CH A R A CTE R IZ A TI O N 7 4 ( 2 0 12 ) 11 0

al. [25]. They studied the thermal stability of this alloy in the
range 443803 K. The UFG structure was found to be quite
unstable. The recrystallization and grain growth resulted in
elimination of UFG microstructure above 550 K. Stability of UFG
microstructure up to 700 K was reported by Furukawa et al. [26]
in a similar work on ECAP processed Al-5.5 Mg\2.2Li\0.12Zr
(wt%). Such a high resistance to grain growth was attributed to
very fine dispersion of Al3Zr dispersoids. In a similar work on
ECAP processed UFG Al\0.2 wt% Zr with grain size 0.7 m, UFG
microstructure was retained only up to 573 K [16]. In another
study by Park et al. [18] on accumulative roll bonding (ARB)
processed UFG 6061Al, thermal stability was studied for 1 h in
the temperature range 373773 K. UFG microstructure was
maintained only up to 473 K. Roy et al. [21] have investigated
the thermal stability of a cryomilled UFG 5083Al alloy. Similar to
Furukawa et al. [26] the stability of UFG microstructure was
reported up to 673 K. However, such a high level of stability of
grains was attributed to the dispersion of particles such as
oxides, carbides, and nitrides on the grain boundaries. This brief
survey on the thermal stability of UFG alloys indicates that the
thermal stability is not very high and use of dispersoids pushes
the stability envelop towards the higher temperature range.
The rate of increase of grain boundary area per unit
volume is strongly influenced by the mode of deformation.
Gil Sevillano et al. [27] showed that rate of generation of grain
boundary area per unit volume was maximum for the sample
deformed under compression than those deformed either by
rolling, wire drawing or torsion for the same level of equivalent
strain. It indicates towards the fact that different processing
techniques introduce different types of microstructure. Hence,
UFG materials processed via different routes will have different
microstructures which would result into different response to
thermal treatment. Therefore, thermal stability of UFG alloys
processed via friction stir processing (FSP) is of interest in spite
of similar studies done on UFG materials processed by other
techniques [1624].
In the present work two different aspects of thermal stability
will be addressed. FSP has potential of refining grains to as small
as 25 nm [2830]. But commonly observed grains are larger than
1 m [31]. Zener pinning mechanism is known to inhibit grain
growth. This is realized by the use of precipitates or dispersoids
in the microstructure. Precipitates are generally not very
effective grain growth inhibitor due to either rapid dissolution
or excessive coarsening during FSP. In the present work the role
of dispersoids for grain stabilization to obtain UFG alloy will be
addressed. Another aspect is grain growth of UFG alloy after
processing.

2.

Experimental Procedure

2.1.

Processing of the Alloys

A twin-roll cast (TRC) Al-Mg-Sc sheet (nominal composition:


Al\4 Mg\0.8Sc\0.08Zr, wt%) of 3.75 mm thickness was
subjected to FSP in as-received (here afterwards referred to
as TRC-AR) and aged (here afterwards referred to as
TRC-Aged) conditions. The aging was carried out at 563 K for
22 h. It resulted into precipitation of Al3(Sc,Zr) dispersoids. To
evaluate the effectiveness of these dispersoids in inhibiting

grain growth during FSP, work hardenable 5086Al\H32 was


friction stir processed under same processing condition as
TRC alloy. A tool rotation rate of 325 revolutions per minute
(rpm) and a tool traverse speed of 8 ipm (203 mm/min) were
used to process these alloys. Tool tilt angle was 2.5 and
plunge depth was 0.097 (2.5 mm). The tool material was tool
steel and tool had concave shoulder and stepped spiral pin
profile. The diameter of the shoulder was 12.0 mm. The pin
diameter at root and the tip were 6.0 mm and 3.5 mm,
respectively. The height of the pin was 2.2 mm measured
from the root of the pin.

2.2.

Annealing Heat Treatment for the Grain Growth Study

For thermal stability study, FSP TRC-Aged and 5086Al-H32


samples were subjected to isothermal annealing heat treatment. FSP TRC-Aged alloy was annealed at 523, 623, 723, and
823 K. Time at each temperature varied from 116 h (at 823 K
only 1 h). FSP 5086Al-H32 was annealed at 523 K for 116 h
and at 623 K for 1 h.

2.3.

Characterization of the Microstructure

Characterization of the microstructure was done by electron


backscattering diffraction (EBSD) microscopy using HKL EBSD
system fitted on FEI Helios NanoLab 600 FIB/FESEM. Each sample
was mechanically polished using water based diamond suspension up to 1 m grit size and then polished using 0.02 m
colloidal silica suspension. EBSD was carried out in as-polished
condition. Centre of the FSP nugget on transverse cross-section
was chosen for EBSD analysis. A Tecnai G2 F20 S-Twin 200 keV
field-emission Scanning Transmission Electron Microscope
(S/TEM) was used for imaging nano-sized Al3(Sc,Zr) dispersoids.
TEM foil was prepared using twin-jet electropolishing technique.
The samples were mechanically polished down to 150 m
thickness. A 2 mm disk was punched out and was further
polished down to 6080 m. It was followed by twin-jet electropolishing at 30 C in CH3OH\20 vol.%HNO3 electrolyte.

3.

Results

3.1.

Thermal Stability During FSP

The microstructure after FSP is shown in Fig. 1 for two different


starting conditions. The accompanying inverse pole figure (IPF)
illustrates the various crystallographic axes associated with
different colors. The IPF has been plotted with respect to FSP
direction which is perpendicular to the plane of the paper.
Hence, these crystallographic directions basically represent
direction vector, in each grain, oriented parallel to FSP direction.
Fig. 1a corresponds to TRC alloy processed in AR condition at
325 rpm and 8 ipm (203 mm/min). EBSD micrographs corresponding to TRC alloy processed in aged condition has been
shown in Fig. 1b. It can be noticed that FSP of TRC-Aged alloy
resulted in a finer microstructure. Quantitative analysis of grain
size showed a mean grain size of 0.50 0.25 m and 0.40
0.16 m for FSP TRC-AR and FSP TRC-Aged, respectively. The
grain size distribution (GSD) histograms for these two conditions have been shown in Fig. 2. In both the cases GSD can be

MAT E RI A LS CH A R A CTE R IZ A TI O N 7 4 (2 0 1 2) 11 0

Fig. 1 EBSD micrographs showing grain structure in FSP Al-Mg-Sc alloy, (a) AR+ FSP and (b) Aged (563 K, 22 h) + FSP.
represented by theoretical Rayleigh or Louat's distribution [32]
and has been shown in the same figure. It is evident from the
GSD plot that some of the grains for FSP TRC-AR were larger
than 1 m. But, almost all the grains were less than 1 m for FSP
TRC-Aged. The GSD histograms show that in both the cases
minimum grain size was same. It is an artifact of step size
selection during EBSD data acquisition. In both the cases
step-size was 0.10 m and minimum grain size was found to
be 0.12 m. A smaller step size may have some impact on the
distribution of grain size towards the smaller grain sizes. But it
is not expected to change the average grain size and overall GSD
considerably.
TEM images for FSP TRC-Aged alloy have been shown in
Fig. 3. Fig. 3a shows Al3(Sc,Zr) particles distributed inside
matrix as well as at grain or subgrain boundaries. In Fig. 3b, a
high resolution TEM (HRTEM) image depicts a nano-sized
Al3(Sc,Zr) particle of approximately 56 nm radius.
FSP 5086-H32 alloys showed an average grain size of 1.1
0.81 m. The cumulative grain size distribution plot for this
alloy has been shown in Fig. 4 along with the cumulative
distribution of TRC Al-Mg-Sc alloy. From this figure, the effect
of dispersoid on the stability of grains during FSP can be

Fig. 2 Histogram showing the grain size distribution of FSP


TRC alloy in as-received (light colored bars) and aged (563 K,
22 h) conditions. Also shown are the theoretical Rayleigh or
Louat's distributions to represent the grain size distribution
of grain size cross-sectional area measurement.

noted. FSP 5086Al-H32 alloy had ~ 40% grains larger than 1 m


whereas FSP TRC alloy had almost all the grains smaller than
1 m (for FSP TRC-AR a very small fraction of grains were
>1 m). Moreover, TRC-Aged alloy showed about 80% grains
below 0.50 m. It was about 60% for TRC-AR alloy.

3.2.

Grain Growth Study

EBSD micrographs corresponding to annealing of FSP 5086Al\


H32 along with as-processed (AP) EBSD micrograph are shown
in Fig. 5. It can be noted that annealing at 523 K from 1 h (Fig. 5b)
to 16 h (Fig. 5c) only resulted in recovery and annihilation of
substructure developed during FSP. In Fig. 5a thin black lines
represent substructure developed in the grains during processing. Due to annealing at 523 K for 1 h grain interiors appear to be
cleaner (Fig. 5b) than that of as-processed condition (Fig. 5a).
When sample was annealed at 623 K for 1 h, abnormal grain
growth (AGG) was observed. Very large grains (~100 m) can be
seen in Fig. 5d along with very fine grains at the centre of the
image. The corresponding cumulative grain size distributions
for these four conditions are shown in Fig. 6. In AP condition
~60% grains were smaller than 1 m. On annealing at 523 K for
1 h, larger grains grew at the expense of smaller grains as
evident from reduction of percentage of grains below 1 m
(38%). But fraction of grains below 1 m marginally changed
after annealing the sample at 523 K for 16 h. Cumulative
frequency curves related to these two conditions also indicate
that there was no significant change in the grain size
distribution on annealing for these two times. In the specimen
annealed at 623 K, the GSD distribution curve up to 34 m
follows GSD distribution curves related to samples annealed at
lower temperature very closely and afterwards become flat up
to 100 m. It is indicative of the presence of very large grains
due to AGG.
The EBSD images corresponding to annealing of FSP
TRC-Aged alloy are shown in Fig. 7. It can be inferred that up
to 623 K for 16 h there was a marginal increase in the grain
size. The EBSD micrograph corresponding to 723 K shows
significant grain growth. Fig. 7d which represents microstructure annealed at 823 K for 1 h shows considerable grain
growth.
Fig. 8a shows the variation of average grain size as a
function of time and temperature. It can be noted that

MAT ER I A LS CH A R A CTE R IZ A TI O N 7 4 ( 2 0 12 ) 11 0

Fig. 3 Transmission electron microscope (TEM) images of Al3(Sc,Zr) dispersoid in FSP TRC-Aged (a) a low magnification image
showing presence of Al3(Sc,Zr) dispersoids (b) a high resolution TEM image showing Al3(Sc,Zr) dispersoid.

annealing of samples at 523 K and 623 K up to 16 h did not


change the average grain size at all. Although, grain growth
was observed at 723 K when annealed between 1 to 16 h the
grain growth kinetics was not very fast. The sample annealed
at 823 K for 1 h showed a substantial grain growth and very
high grain growth kinetics as evident from the slope of the
curve. In Fig. 8b, thermal stability of FSP TRC-Aged alloy
annealed at different temperature for 1 h has been compared
with the thermal stability of UFG alloys processed by various
other routes. Comparison of the present data shows that most
of the UFG alloys have very poor thermal stability and above
~ 473 K UFG microstructure is lost. The FSP TRC-Aged alloy
maintained its UFG microstructure up to 723 K.

4.

Discussion

As mentioned before FSP has the potential for producing


nanocrystalline grains. But, this advantage is lost due to
excessive grain growth of refined microstructure during FSP
because of high temperature involved during the process. The

thermal cycle a processed volume goes through during FSP is


shown schematically in Fig. 9. The final grain size depends
largely on how much time a processed volume spends above
Tmin, minimum temperature to induce grain growth. One way
of cutting down time spent above Tmin is to increase the rate
of cooling as shown by red (curve 1), blue (curve 2), and black
(curve 3) lines representing cooling rates in increasing order.
These different levels of cooling rates can be obtained by
using external cooling media such as liquid nitrogen, copper
backing plate, forced chilled air cooling, etc. Another method
of reducing the time above Tmin would be to reduce the peak
temperature (shown by curve 4). The peak temperature (Tpeak)
can be manipulated by controlling the tool rotation rate and/
or tool traverse speed. These two strategies have been widely
used to stabilize the processed microstructure during FSP.
Here, use of a third strategy has been shown to refine the
microstructure during FSP as outlined next.
In the present study, two different scenarios were investigated (a) precipitation of dispersoids during FSP, i.e., FSP
of as-received Al\Mg\Sc which contained Sc in solid solution
and (b) FSP in aged condition. The effectiveness of case (a) in
grain refinement was lower than that of case (b). Since, in case
(a) dispersoids have to precipitate out during FSP, their
effectiveness in acting as Zener pinning agent will depend
on the kinetics of stable precipitate nuclei formation and also
the volume fraction of the dispersoids. For coherent stable
particles or dispersoids containing alloys, limiting grain size is
given as
2r
D
3F V

Fig. 4 Comparison of grain size distribution of FSP TRC-AR


and TRC-Aged alloys with FSP 5086Al-H32 alloys.

r and FV are average grain size in FSP condition,


where D,
radius and volume fraction of precipitates or dispersoids,
respectively [10]. Ignoring the amount of Zr, the present alloy
system can be treated as a binary system (Al,Mg) \Sc.
Al\Sc literature shows that depending upon the rate of
cooling solid solubility of Sc in Al can vary anywhere from
0.35 wt.% (equilibrium cooling) to 0.6 wt.% (at a cooling rate of
100 K/s) [33]. In the twin-roll casting a cooling rate of 100 K/s is
easily obtained. Hence, 0.6 wt.% of the Sc can be easily

MAT E RI A LS CH A R A CTE R IZ A TI O N 7 4 (2 0 1 2) 11 0

Fig. 5 EBSD micrographs show the grain growth behavior of FSP 5086Al-H32 alloy in different conditions; (a) as-processed, (b) 523 K,
1 h, (c) 523 K, 16 h, and (d) 623 K, 1 h.

assumed to be present in Al\Mg\Sc alloy. If all the Sc in solid


solution precipitates out in the form of Al3(Sc,Zr) dispersoids,
applying the lever rule it can be shown that it will correspond
to 0.0168 volume fraction of these dispersoids. Based on the
present TEM work (and also supported by the work of Lee et al.
[34] and Marquis et al. [35] on Al\Mg\Sc system), the average
size of Al3(Sc,Zr) dispersoids are expected to be ~ 5 nm in size.

Fig. 6 Cumulative grain size distribution of FSP


5086Al-H32 alloy in various thermo-mechanical conditions
(AP: as-processed).

Substituting these values (FV = 0.0168 and r = 5 nm) in Eq. (1),


is ~ 200 nm. The experimentally observed
limiting grain size D
mean grain size in the case (b) was also 400 nm. The
discrepancy between the predicted and experimentally measured values is quite expected. Eq. (1) has been derived for the
case where Zener pinning force is maximum. We know, in
reality, not all the particles will exert maximum pinning force
on grain boundaries. Hence, the predicted values would be
smaller than the experimentally measured one, as is the case
here. The measurable coarsening of Al3(Sc,Zr) dispersoids
takes place only above 573 K, if heated for considerable
amount of time. In the case (a), a material volume in the
wake of the FSP tool is not expected to be above 573 K for a
considerable amount of time. Hence, the dispersoid size
should be about 23 nm. The works of Fuller et al. [36],
Marquis and Seidman [37], and Knipling et al. [38] support this
assumption. The mean grain size found in the case (a) was
0.5 m. Substituting these values (limiting grain size equal to
mean grain size and dispersoid size equal to 2 nm) in Eq. (1),
the volume fraction of the dispersoids was found to be 0.0026.
The estimated volume fraction in this case is lower than that
of case (b). This is an expected result. Given the very short
duration of FSP, complete precipitation of the dispersoids is
unlikely. So, the volume fraction of Al3(Sc,Zr) dispersoids
would be lower. Hence, in spite of having smaller dispersoid
size in the case (a), lower volume fraction of the dispersoids
resulted into a higher limiting grain size. Other noticeable

MAT ER I A LS CH A R A CTE R IZ A TI O N 7 4 ( 2 0 12 ) 11 0

Fig. 7 EBSD micrographs show the grain growth behavior of FSP TRC-aged alloy in different conditions; (a) 523 K, 16 h, (b) 623 K,
16 h, (c) 723 K, 16 h, and (d) 823 K, 1 h.

difference due to the presence of Al3(Sc,Zr) dispersoids before


FSP was higher frequency of smaller grain sizes.
Grain growth study revealed that presence of stable
dispersoids like Al3(Sc,Zr) can stabilize very fine grain size at
temperatures as high as 723 K. Such a high level of thermal
stability of grains is hardly observed in any of the previously
reported UFG materials except for a report by Ma et al. [39] for
Al\4Mg\1Zr alloy. The highly thermal resistant Al3Zr dispersoids were attributed for such a high thermal stability of this
alloy system. The UFG microstructure was maintained up to

698 K (mean grain size 0.98 m at this temperature) in this


case. Most of the UFG Al alloys lose their UFG structure above
473 K523 K.
The case of AGG for particle containing alloys have been
dealt with by many researchers, but only the model proposed
by Humphreys' [40] will be discussed here. The condition for
AGG is given as

 dR R dR > 0
R
dt
dt

Fig. 8 (a) Grain size as a function of time and temperature for FSP TRC-aged alloy. (b) Thermal stability comparison plot for FSP
UFG TRC-aged alloy with literature data [1618,20,21].

MAT E RI A LS CH A R A CTE R IZ A TI O N 7 4 (2 0 1 2) 11 0

Fig. 9 A schematic illustration of the thermal cycle experienced by a particular volume of material during FSP and its effect on
grain refinement process.

 and R are mean grain size and size of any grain,


where R
respectively, in a polycrystalline material.


 Z
dR

M  
dt
R R R

 M


dR
 0:25Z
dt
R

where for incoherent particles



Z incoh 3F V R=2r

and for coherent particles



Z coh 3F V R=r


 , M, Mand
Here, ,
Z are average grain boundary energy,
grain boundary energy, grain boundary mobility, average
grain boundary mobility, and dimensionless particle pinning
parameter, respectively. For the present case substituting
 1 and
Eqs. (3) and (4) into expression (2) and assuming =

M=M 1, expression (2) can be rewritten as
4Z1X 2 41Z 4 > 0


for Z < 0.25. For Z 0.25, dR=dt
0. Hence, AGG can take place

only if RdR=dt > 0, i.e.,


 Z


 dR RM
 >0
R

dt
R R R

 , the Eq. (8) leads to the


Again under the assumption
following inequality,
X > 1=1Z


where X R=R.
 with Fv/r
The variation of X with Z and mean grain size (R)
has been shown in Fig. 10 for coherent and incoherent
particles. It depicts the regions of AGG and no grain growth.
When Z reaches unity, the requirement for minimum grain
size to initiate AGG becomes so large (as Z 0, R from
Eq. (8)) that at the dispersion level higher than this grain
growth seizes completely. In Fig. 10a, it is represented by
curve Xmin and in Fig. 10b curve representing Z = 1 (coh) and
Z = 1 (incoh). In Fig. 10b words coh represents coherent and

incoh incoherent particles. Fig. 10a has been plotted using


Eqs. (7) and (9) and Fig. 10b using Eqs. (5) and (6). As per Jones
and Humphreys, coherency is lost at ~ 20 nm. Hence, in the
present analysis in plotting Fig. 10a and b both aspects have
been considered. Parameter Z is a function of temperature. For
the present case it was calculated for all the annealing
temperatures. To calculate Z, volume fraction of dispersoids,
mean grain size (radius), and dispersoid radius were needed.
The volume fraction calculation was done at each temperature based on the solvus curve related to Al-Sc reported by
Ryset and Ryum [33]. The procedure of volume fraction
calculation at each temperature was same as outlined in the
discussion on thermal stability during FSP. The mean grain
size radius was calculated from the EBSD analysis. Since in
present work TEM analysis provides dispersoid size information for FSP TRC-Aged in as-processed condition only, the
dispersoid radius was taken from the work of Jones and
Humphreys [41] on Al3Sc dispersoids coarsening as a function
of temperature. The calculated Z values are listed in Table 1.
The calculated Z values were superimposed on the grain
stability map shown in Fig. 10a. From the values of Z, it is clear
that it decreases with increase in temperature. A similar
calculation by Charit and Mishra [42] for precipitates/dispersoids in FSP UFG Al-Zn-Mg-Sc alloy showed a decrease in Z
with the increase in temperature.
The Z values at 523 K and 623 K fall into no growth region
of the grain stability map. The grain size histograms corresponding to annealing at 523 K and 623 K for 1 h are shown in
Fig. 10c. Clearly, both the histograms superimpose each other
completely. Also, mean grain size in these two conditions
were same as in as processed condition. The Z value for the
sample annealed at 723 K for 1 h also lies in the no growth
region but this time it is close to AGG region. Ideally, there
should not be any grain growth in this sample too since it is a
part of no growth region. The mean grain size in this case was
0.74 m. It can be attributed to lowering of Zener pinning
pressure due to coarsening of Al3(Sc,Zr) dispersoids. The grain
size distribution histogram is shown in Fig. 10d. An increase in
relative frequency of grains in the range 1.5 to 2.0 m can be
noted. But still grain size distribution is same as those
annealed at lower temperatures. If the sample is annealed at
723 K for longer period of time, it will cause further coarsening
of Al3(Sc,Zr) dispersoids. That may result in reduction of Z value
further (and reduction of Fv/r too) and consequential shift in

MAT ER I A LS CH A R A CTE R IZ A TI O N 7 4 ( 2 0 12 ) 11 0

Fig. 10 Grain stability maps (a) X Vs. Z and (b) mean grain size Vs. Fv/r; (cf) comparison of the predictions made by this map
with the experimentally observed grain growth behavior.

the position of datum point towards the AGG region and may
eventually become part of AGG region as shown by a blue
colored open circle in Fig. 10a (and in Fig. 10b also). In such a
scenario the sample should show AGG. The analysis of grain
size distribution histogram of the sample annealed at 723 K
for 16 h shows a bimodal grain size distribution as shown in
Fig. 10e. Apart from a peak in UFG regime, the presence of

another peak at ~ 1 m can also be noted. The bimodal grain


size can be considered as an onset of AGG. The sample
annealed at 823 K for 1 h showed AGG. It is evident not only
from the EBSD micrographs shown in Fig. 7d but also from the
grain size distribution histogram shown in Fig. 10f. Apart from
the presence of multiple peaks, some grains larger than 10 m
can also be noted in Fig. 10f. It should be noted that the Z

Table 1 Values of different variables used in the calculation of particle pinning parameter, Z.

Dispersoid size (radius, nm)


Solubility of Sc in Al (wt%)
Dispersoid volume fraction
Mean grain size (radius, m)
Particle pinning parameter, Z
Fv/r, m1

523 K

623 K

723 K

823 K

5
Negligible
0.0168
0.26
2.6
3.36

5
Negligible
0.0168
0.26
2.6
3.36

9
0.034
0.0159
0.26
1.4
1.8

47
0.125
0.0133
0.26
0.11
0.28

MAT E RI A LS CH A R A CTE R IZ A TI O N 7 4 (2 0 1 2) 11 0

value corresponding to the sample annealed at 823 K falls in


AGG regime as shown in Fig. 10a. Hence, it satisfies the
condition for AGG which is evident from Fig. 10e. Similar
conclusions can be drawn based on Fig. 10b too.

5.

Conclusions

The following conclusions can be made based on foregoing


discussion:
I. Dispersoids like Al3(Sc,Zr) are effective means of
inhibiting the grain growth during FSP.
II. The presence of the dispersoids before FSP is much
more effective as Zener pinning agent than their precipitation during FSP.
III. Post-FSP annealing heat treatment showed a very high
thermal stability of UFG microstructure. UFG microstructure was maintained up to 723 K.
IV. Humphreys model for abnormal grain growth can be
utilized to explain the thermal stability of the present
alloy system.

Acknowledgment
Authors would like to acknowledge the financial assistance
from The Boeing Company, St. Louis and Intelligent System
Centre, Missouri S&T, Rolla to carry out this work. Authors are
also grateful to Dr. Saumya Nag at University of North Texas
for carrying out high resolution transmission electron microscopy for the present work.

REFERENCES

[1] Valiev RZ, Islamgaliev RK, Alexandrov IV. Bulk


nanostructured materials from severe plastic deformation.
Prog Mater Sci 2000;45:103-89.
[2] Valiev RZ, Langdon TG. Principles of equal-channel angular
pressing as a processing tool for grain refinement. Prog Mater
Sci 2006;51:881-981.
[3] Zhilyaev AP, Langdon TG. Using high-pressure torsion for
metal processing: fundamentals and applications. Prog Mater
Sci 2008;53:893-979.
[4] Saito Y, Utsunomiya H, Tsuji N, Sakai T. Novel ultra-high
straining process for bulk materials development of the
accumulative roll-bonding (ARB) process. Acta Mater 1999;47:
579-83.
[5] Tsuji N, Saito Y, Lee S-H, Minamino Y. ARB (accumulative
roll-bonding) and other new techniques to produce bulk
ultrafine grained materials. Adv Eng Mater 2003;5:338-44.
[6] Hall EO. The deformation and ageing of mild steel: III
Discussion of results. Proc Phys Soc B 1952;64:747-53.
[7] Petch NJ. The cleavage strength of polycrystals. J Iron Steel
Res Inst 1953;174:25-8.
[8] Gleiter H. Nanocrystalline materials. Prog Mater Sci 1989;33:
223-315.
[9] Gleiter H. Materials with ultrafine microstructures: retrospectives and perspectives. Nanostruct Mater 1992;1:119.
[10] Humphreys FJ, Hatherly M. Recrystallization and related
annealing phenomena. Oxford: Pergamon; 1995.

[11] Malow TR, Koch CC. Grain growth in nanocrystalline iron


prepared by mechanical attrition. Acta Mater 1997;45:2177-86.
[12] Lu L, Wang LB, Ding BZ, Lu K. Comparison of the thermal
stability between electro-deposited and cold-rolled
nanocrystalline copper samples. Mater Sci Eng A 2000;286:
125-9.
[13] Andrievski RA. Stability of nanostructured materials. J Mater
Sci 2003;38:1367-75.
[14] Liu F, Yang G, Wang H, Chen Z, Zhou Y. Nano-scale grain
growth kinetics. Thermochim Acta 2006;443:212-6.
[15] Chauhan M, Mohamed FA. Investigation of low temperature
thermal stability in bulk nanocrystalline Ni. Mater Sci Eng A
2006;427:715.
[16] Hasegawa H, Komura S, Utsunomiya A, Horita Z, Furukawa
M, Nemoto M, et al. Thermal stability of ultrafine-grained
aluminum in the presence of Mg and Zr additions. Mater Sci
Eng A 1999;265:188-96.
[17] Park KT, Kim YS, Lee JG, Shin DH. Thermal stability and
mechanical properties of ultrafine grained low carbon steel.
Mater Sci Eng A 2000;293:165-72.
[18] Park KT, Kwon HJ, Kim WJ, Kim YS. Microstructural and
thermal stability of ultrafine grained 6061 Al alloy fabricated
by accumulative roll bonding process. Mater Sci Eng A
2001;316:145-52.
[19] Zhilyaev A, Nurislamova G, Valiev R, Baro M, Langdon
T. Thermal stability and microstructural evolution in
ultrafine-grained nickel after equal-channel angular pressing
(ECAP). Metall Mater Trans A 2002;33:1865-8.
[20] Geng H, Kang S, He S. Microstructural evolution and thermal
stability of ultra-fine grained Al-4Mg alloy by equal channel
angular pressing. J Mater Sci Technol 2004;20:315-8.
[21] Roy I, Chauhan M, Lavernia EJ, Mohamed FA. Thermal
stability of bulk cryomilled ultrafine-grained 5083 Al alloy.
Metall Mater Trans A 2006;37:721-30.
[22] Homola P, Slmov M, Oenek V, Uhl J, Cieslar M. Annealing
response of Al\0.22Sc\0.13Zr alloy processed by accumulative
roll bonding. Mater Sci Forum 2008;584586:899-904.
[23] Lugo N, Llorca N, Suol JJ, Cabrera JM. Thermal stability of
ultrafine grains size of pure copper obtained by
equal-channel angular pressing. J Mater Sci 2010;45:2264-73.
[24] Nikitina MA, Islamgaliev RK, Kamalov AF. Thermal stability
of the ultrafine-grained Al\Cu\Mg\Si aluminum alloy. Rev
Adv Mater Sci 2010;25:74-81.
[25] Wang J, Iwahashi Y, Horita Z, Furukawa M, Nemoto M, Valiev
RZ, et al. An investigation of microstructural stability in an
Al\Mg alloy with submicrometer grain size. Acta Mater
1996;44:2973-82.
[26] Furukawa M, Iwahashi Y, Horita Z, Nemoto M, Tsenev NK,
Valiev RZ, et al. Structural evolution and the HallPetch
relationship in an Al\Mg\Li\Zr alloy with ultra-fine grain
size. Acta Mater 1997;45:4751-7.
[27] Gil Sevillano J, van Houtte P, Aernoudt E. Large strain work
hardening and textures. Prog Mater Sci 1980;25:135-271.
[28] Su J-Q, Nelson TW, Sterling CJ. A new route to bulk
nanocrystalline materials. J Mater Res 2003;18:1757-60.
[29] Rhodes CG, Mahoney MW, Bingel WH, Calabrese M.
Fine-grain evolution in friction-stir processed 7050 aluminum.
Scr Mater 2003;48:1451-5.
[30] Mishra RS. Bulk nanomaterials from friction stir processing:
Features and properties. In: Zehetbauer ZMJ, Zhu YT, editors.
Bulk Nanostructured Materials. KGaA Weinheim: Wiley-VCH
Verlag GmbH & Co; 2009. p. 255-72.
[31] Mishra RS, Ma ZY. Friction stir welding and processing. Mater
Sci Eng R 2005;50:178.
[32] Anderson MP, Grest GS, Srolovitz DJ. Computer simulation of
normal grain growth in three dimensions. Philos Mag B
1989;59:293-329.
[33] Royset NRJ. Scandium in aluminum alloy. Int Mater Rev
2005;50:19-43.

10

MAT ER I A LS CH A R A CTE R IZ A TI O N 7 4 ( 2 0 12 ) 11 0

[34] Lee S, Utsunomiya A, Akamatsu H, Neishi K, Furukawa M,


Horita Z, et al. Influence of scandium and zirconium on grain
stability and superplastic ductilities in ultrafine-grained
Al-Mg alloys. Acta Mater 2002;50:553-64.
[35] Marquis EA, Seidman DN, Dunand DC. Effect of Mg addition
on the creep and yield behavior of an Al-Sc alloy. Acta Mater
2003;51:4751-60.
[36] Fuller CB, Seidman DN, Dunand DC. Mechanical properties of
Al(Sc, Zr) alloys at ambient and elevated temperatures. Acta
Mater 2003;51:4803-14.
[37] Marquis EA, Seidman DN. Nanoscale structural evolution of
Al3Sc precipitates in Al(Sc) alloys. Acta Mater 2001;49:1909-19.
[38] Knipling KE, Karnesky RA, Lee CP, Dunand DC, Seidman DN.
Precipitation evolution in Al-0.1Sc, Al-0.1Zr and
Al-0.1Sc-0.1Zr (at%) alloys during isochronal aging. Acta
Mater 2010;58:5184-95.

[39] Ma ZY, Liu FC, Mishra RS. Superplastic deformation


mechanism of an ultrafine-grained alloy produced by friction
stir processing. Acta Mater 2010;58:4693-704.
[40] Humphreys FJ. A unified theory of recovery, recrystallization
and grain growth, based on the stability and growth of
cellular microstructures - II. The effect of second-phase
particles. Acta Mater 1997;45:5031-9.
[41] Jones MJ, Humphreys FJ. Interaction of recrystallization and
precipitation: The effect of Al3Sc on the recrystallization
behavior of deformed aluminum. Acta Mater 2003;51:2149-59.
[42] Charit I, Mishra RS. Low temperature superplasticity in a
friction-stir-processed ultrafine grained Al-Zn-Mg-Sc alloy.
Acta Mater 2005;53:4211-23.

Das könnte Ihnen auch gefallen