Sie sind auf Seite 1von 4

Eect of grain size renement and precipitation reactions

on strengthening in friction stir processed AlCu alloys


Xiuli Feng,
a,b
Huijie Liu
a,
and Sudarsanam Suresh Babu
b
a
State Key Laboratory of Advanced Welding and Joining, Harbin Institute of Technology, Harbin 150001, Peoples Republic of China
b
Department of Materials Science and Engineering, The Ohio State University, OH 43221, USA
Received 4 August 2011; revised 3 September 2011; accepted 6 September 2011
Available online 10 September 2011
Strengthening by grain size renement and precipitation in friction stir processing (FSP) of AlCu alloys was investigated. Inter-
estingly, even with signicant grain renement (<1 lm), there was a substantial reduction in microhardness. This is attributed to the
loss of precipitation hardening due to the formation of equilibrium h in lieu of metastable h
0
and h
00
precipitates. Theoretical calcu-
lations suggest that the average grain size in the FSP zone must be smaller than 120 nm to compensate for this loss.
2011 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Friction stir processing; Grain rening; Precipitation; Microhardness
Friction stir processing (FSP) was developed
based on friction stir welding (FSW), in which a rotating
tool is plunged into a xed workpiece and traversed
along the line of interest [1,2]. Since its development,
FSP has been applied to modify casting microstructures
[3], achieve superplasticity [4] and fabricate surface com-
posites [5]. Critical process parameters in FSP include the
tool geometry, rotation speed, traverse speed and exter-
nal cooling conditions. The grain size can be controlled
by changing these parameters. For example, researchers
have used FSP to obtain ultrane-grained and nanocrys-
talline aluminum alloys [610]. The grain size reported in
the literature ranges from 0.02 to 0.7 lm.
Grain renement is an eective means to improve the
mechanical properties, e.g. yield strength, of friction stir
processed material. This approach has been demon-
strated in FSP of 1050 Al alloy [10]. However, in age-
hardenable aluminum alloys, precipitation strengthening
is the dominant mechanism [11]. Dixit et al. [12] calcu-
lated the contribution of precipitation strengthening for
7050 and 7055 Al in peak-aged condition to be
100 MPa for both alloys. In contrast, the strengthening
contribution from a matrix with an average grain size
of 1.5 lm is only 69 MPa. However, during FSW/FSP,
coarsening, dissolution of strengthening precipitates, as
well as reprecipitation of metastable and/or equilibrium
phases are expected. These microstructure evolutions
may eventually lead to softening [1315]. The above re-
search also suggests that it is possible to oset the soften-
ing by greatly reducing the grain size to ner than
1.5 lm. Approaches to combine the strengthening due
to ultrane grain sizes and precipitation have not been
thoroughly explored. The goal of the current research
is to analyze the strengthening due to precipitation and
grain renement (<1.5 lm) in age-hardenable aluminum
alloys. To rene the grain size below 1.5 lm, FSP was
performed under water at dierent traverse speeds.
The 2.5 mm thick age-hardenable commercial AlCu
alloy (2219-T6 Al) plates were xed in a tank lled with
water at 298 K. A pinless tool with a shoulder diameter
of 22 mmwas utilized. FSP was performed at a tool rota-
tion speed (x) of 1000 rpm and a traverse speed (v) of
200400 mm min
1
. Displacement control was used to
maintain the plunge depth at 0.5 mm into the top sheet
surface. The tilt angle of the rotating tool with respect
to the Z-axis of the FSWmachine was 2.5. The directions
of the FSPgeometry are referred to LD, TDand ND(lon-
gitudinal, transverse and normal directions, respectively).
The friction stir processed plates were cross-sectioned
perpendicular to the LD, and then ground and polished
to 1 lm. Vickers microhardness was measured (load
50 g, dwell time 10 s, center-to-center spacing 120 lm)
using a Leco AMH-43 microhardness machine. Electron
back-scattered diraction (EBSD) analysis (step size of
0.2 lm) was performed using a Quanta 200 operated
at 20 kV to analyze the grain structure and texture
1359-6462/$ - see front matter 2011 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.scriptamat.2011.09.009

Corresponding author. e-mail addresses: liuhj@hit.edu.cn; alimse@


163.com
Available online at www.sciencedirect.com
Scripta Materialia 65 (2011) 10571060
www.elsevier.com/locate/scriptamat
development. For EBSD analyses, the samples were
polished using the same procedure as that for the
microhardness test, except that the nal polishing
consisted of vibratory polishing using a VibroMet

2
Vibratory Polisher with 0.02 lm colloidal silica.
Transmission electron microscopy (TEM) was used
to characterize the precipitates in the stirred zone.
TEM specimens were made by electropolishing using a
twinjet technique in a solution of 30% nitric acid in
methanol at 20 C. All TEM samples were examined
using a Philips CM200 microscope operated at 200 kV.
The microhardness proles measured in the processed
samples are shown in Figure 1. The microhardness maps
show that the stirred zones are softer (90105 HV) than
the base metal (138 HV). Interestingly, as v increased
from 200 to 400 mm min
1
, the peak microhardness in
stirred zone increased from 90 to 105 HV. It should be
noted that the rotation speed (1000 rpm) was kept con-
stant for all the above FSP conditions. Therefore, an in-
crease in hardness could be related to a reduction in grain
size due to the higher cooling rate behind the tool region.
The EBSD maps of grain orientations for the base
metal and stirred zone processed at v = 200 mm min
1
are shown in Figure 2a and b. The grain size of the base
metal varies from 10 to 30 lm (Fig. 2a) with an average
of 17 lm. A recrystallized microstructure (14 lm grain
size) was observed in the stirred zone (Fig. 2b). The
average grain size of the stirred zone was found to be
1.3 lma reduction in average grain size over an order
of magnitude compared to the base metal. Detailed
analyses of the EBSD data from stirred zone showed
that 15% of grains were smaller than 1 lm (Fig. 3a).
This suggests that FSP renes grains but may not yield
homogeneous grain size distribution throughout the
processed region. Based on the published literature, it
is hypothesized that these heterogeneities are caused
by non-uniform thermal, strain and strain-rate distribu-
tions during FSP [16].
The microstructures in the stirred zone are known to
exhibit ne grain due to dynamic recrystallization. Dier-
ent recrystallization mechanisms have been reported
[17,18], including dynamic recovery (DRV), continuous
dynamic recrystallization (CDRX), discontinuous
dynamic recrystallization (DDRX), geometric dynamic
recrystallization (GDRX) and particle-stimulated nucle-
ation (PSN). In order to rationalize the mechanism of
grain renement in this research, the texture of the base
metal and stirred zone was analyzed using TSL

EBSD
analysis software. The strength of textures (maximum
intensity 7.6) in the stirred zone is higher than that in
the base metal (maximum intensity 4.1). The results of
the texture component analysis for the base metal and
stirred zone are shown in Table 1. Recrystallization
texture (013)h331i with an area percentage of 11.9
and 10.3 was found in the base metal and the stirred zone,
respectively. In stirred zones, simple-shear textures [A
1
*
(111)h211i and C (001)h110i] with area percent-
ages of 12.4 and 11.4 were alsopresent. According to Fon-
da and Bingert [19], A
1
*
texture appeared at low strain
and C texture appeared at relatively high strain (35).
Another study by McNelley et al. revealed that retention
of a shear-type deformation texture is related to DRVand
GDRX[18]. In addition, randomtexture was observed in
FSP regions and this may suggest PSN as a plausible
mechanism. Further analyses of the EBSD data showed
that the area fraction of the high-angle grain boundary
(HAGB) for the stirred zone is 69% (Fig. 3b), which is
higher than that for the base metal (40%). This
phenomenon of increased misorientation spread is often
associated with CDRX. Based onthe above, it can be con-
Figure 1. Microhardness maps under FSP (x = 1000 rpm): (a)
v = 200 mm min
1
; (b) v = 300 mm min
1
; (c) v = 400 mm min
1
.
Figure 2. EBSD maps and pole gures for (a) base metal and (b) FSP
at v = 200 mm min
1
.
Figure 3. EBSD analysis: (a) grain size distribution; (b) misorientation
angle distribution.
Table 1. Texture component of the base metal and FSP zone
(x = 1000 rpm, v = 200 mm min
1
).
Base metal FSP zone
(hkl)huvwi % (hkl)huvwi %
(013)h331i 11.9 (013)h331i 10.3
(011)h111i 3.1 (011)h111i 0.8
(101)h010i 3.8 (111)h211i 12.4
(4411)h11118i 3.9 (001)h110i 11.4
(231)h124i 2.8 (112)h111i 6.3
(241)h112i 1.0 (123)h634i 1.1
1058 X. Feng et al. / Scripta Materialia 65 (2011) 10571060
cluded that DRV, GDRX, PSN and CDRX may all be
possible mechanisms during FSP, as a function of spatial
locations and transient conditions. It is recognized that
this conclusion is only applicable to samples processed
at a traverse speed of 200 mm min
1
.
For the samples processed at traverse speeds of 300
and 400 mm min
1
, the microstructures of the stirred
zone could not be distinguished by EBSD due to poor
diraction quality associated with high defect (disloca-
tion) density in the grains [20]. Higher defect densities
are expected due to incomplete recrystallization of these
samples as a result of fast cooling behind the tool. This
hypothesis is conrmed by TEM images that exhibit
microstructures with signicant dislocation networks in
the samples processed at 300 and 400 mm min
1
(Fig. 4).
For 2219-T6 aluminum alloy, the main strengthening
is due to plate-shaped coherent h
00
and semi-coherent h
0
precipitates [21]. However, TEM micrographs from all
the processed samples (Fig. 4ac) do not show plate-
shaped (h
00
or h
0
) precipitates. Instead, disc-shaped inter-
granular and intragranular precipitates are observed in
the thin foil regions. Cursory analyses also show that
the majority of these precipitates are associated with
the grain boundaries. Selected-area diraction was used
to analyze the precipitate (Fig. 4d) in the FSP sample
processed at v = 200 mm min
1
. These precipitates were
identied as equilibrium h phase using dark-eld imag-
ing and indexing of selected-area diraction patterns.
Qualitative analyses of the images (Fig. 4ac) also
showed that the interprecipitate spacing and precipitate
size decreased as the FSP traverse speed increased (Ta-
ble 2). In agreement with earlier hypothesis, the TEM
analyses clearly show that the matrix grains are less than
1 lm in the FSP samples processed at traverse speeds of
300400 mm min
1
(Fig. 4b and c).
Even with the presence of ultrane grains (<1 lm),
softened regions were observed in all the samples. An at-
tempt to strengthen these FSP regions by post-processing
aging (165 C, 18 h) did not improve the microhardness.
This result indicates that the matrix is not supersaturated
with alloying elements after FSP. TEM observations in
the as-processed region also indicate the presence of equi-
librium h phase. The lack of aging response and the pres-
ence of equilibrium h can be rationalized by two reasons.
First, the metastable precipitates (h
00
and h
0
) over-age to
equilibrium h phase by dwelling at a temperature above
480 C. Second, the metastable precipitates may dissolve
into the matrix upon heating above the solvus tempera-
ture (513 C) and reprecipitate as equilibrium h phase
upon slow cooling behind the tool. To determine which
of the above mechanisms is operating, computational
modeling of thermal cycles was conducted using the com-
mercial software FSWS supplied by Applied Optimiza-
tion. The software has been validated using data
available in the literature [22,23]. The modeling results
indicate that the peak temperature in the stirred zone is
520 C when v = 200 mm min
1
, and decreases to
500 Cwhen v increases to 300400 mm min
1
. This indi-
cates that the second mechanism may be operating when
v = 200 mm min
1
. Similarly, the rst mechanismmay be
operating when v is 300400 mm min
1
. The stable pre-
cipitates h have a much weaker strengthening eect than
the metastable (h
00
and h
0
) precipitates because of the loss
of coherency strains. This decreased strengthening is
probably the reason for the softening in the stirred zone.
The above hypothesis is evaluated quantitatively using
published strengthening models. The contribution to
yield strength of ne grain strengthening can be expressed
by the HallPetch relationship:
r
y
r
0

k

d
p ; 1
where r
y
is the increase in yield strength due to ne grain
strengthening, r
0
is the friction stress (20 MPa), k is a
constant (0.04 MPa m
1/2
) [24] and d is the grain size.
Using Eq. (1), the yield strength increment was calcu-
lated to be 55 MPa for a grain size of 1.3 lm, which
was 25 MPa higher than the grain renement strength-
ening contribution for the base metal.
The precipitation strengthening in the base metal is
due to the presence of plate-shaped metastable precipi-
tates. The strengthening eect is described by [25]:
Dr
0:85Gb
2p

1 m
p
C
D1 pCt=2D
ln
2D
pr
0

; 2
where Dr is the increase in yield strength due to precip-
itation strengthening, m is the Poissons ratio (0.3 for Al)
and G is the shear modulus (26 GPa for Al) [26].
C

fA
p
2=p p=2AfA; 3
where f is the volume fraction of precipitates, D is the
diameter of the precipitates, t is the thickness of precipi-
tates and r
0
is the dislocation core radius (6 10
10
m)
[27]. The value of f was calculated to be 0.12 fromthe fol-
lowing equation [28]:
x
b
Cu
fx
p
Cu
1 f x
m
Cu
; 4
Table 2. Yield strength increment due to precipitate strengthening in
FSP regions.
v (mm min
1
) k (nm) r (nm) Dr (MPa)
200 337 100 16
300 167 67 31
400 125 45 39
Figure 4. TEM images under FSP (x = 1000 rpm): (a)
v = 200 mm min
1
, (b) v = 300 mm min
1
; (c) v = 400 mm min
1
; (d)
dark-eld image and selected-area diraction of precipitate.
X. Feng et al. / Scripta Materialia 65 (2011) 10571060 1059
where the variables x
b
, x
p
and x
m
correspond to the con-
centration of Cu in bulk alloy, precipitates and matrix,
respectively. The values of D and t are measured to be
130 and 10 nm based on the publication by Chen et al.
[23]. The calculated Dr for the base metal is 123 MPa.
The precipitates in the stirred zone are stable disc-
shaped h, and the precipitation strengthening increment
under this condition can be calculated using the
OrowanAshby equation [29]:
Dr
0:13Gb
k
ln
r
b
; 5
where k is the interparticle spacing, r is the particle ra-
dius and b is the Burgers vector (2.84 10
10
m for
Al) [26]. k and r were measured from the TEM images.
The calculated Dr after FSP is summarized in Table 2.
As indicated in Table 2, the strengthening eect of pre-
cipitates in the stirred zone varies from 16 to 39 MPa.
Comparing this estimate with base metal (123 MPa)
strength, the decrease in yield strength in the stirred zone
is estimated to be 84107 MPa. This is much larger than
the estimated increase in strength caused by grain rene-
ment (25 MPa). To oset this decrease in yield strength,
the required reduction in grain size can be estimated
using Eq. (1). The calculations suggest that the average
grain size throughout the stirred zone should be less
than 120 nm. The feasibility of obtaining such a ne
grain size is discussed below.
Ultrane and nanograins have been attained in alu-
minum alloys by FSW/FSP and rapid cooling [610].
However, the challenge is to ensuring a condition that
will lead to uniform strain, strain rate and rapid cooling
rate to maintain the grain size and minimize equilibrium
precipitation strengthening throughout the stirred
zones. On the other hand, if the stirred zone temperature
is higher than solvus temperature and rapid cooling can
still be achieved, it will be possible to obtain a supersat-
urated matrix with ne grain size. Upon aging such a
sample after FSP, it is possible to obtain strengthening
by both grain boundary formation and precipitation.
Similarly, if the stirred zone temperature is maintained
below the solvus temperature and the dwell time is short,
transformation of metastable precipitates to equilibrium
h precipitate will also be controlled. Further research on
the eect of rapid cooling after FSP on the precipitation
kinetics should be performed.
In summary, underwater FSP was performed on the
surface of 2.5 mm thick 2219-T6 Al plates. Fine
(1.3 lm) and ultrane grain sizes (<1 lm) were obtained
by FSP using a constant tool rotating speed of 1000 rpm
and traverse speeds of 200 and 300400 mm min
1
,
respectively. The simple shear texture, random texture
and increased grain boundary misorientations in the stir-
red zone indicate that DRV, GDRX, PSN and CDRX
are all possible mechanisms of grain renement. Despite
signicant grain renement, the microhardness of the
samples decreased to 90105 HV after FSP. This was
attributed to the loss of metastable age-hardening precip-
itates (h
00
and h
0
) and the formation of equilibrium h pre-
cipitates during FSP, which reduced the precipitation-
hardening contribution. The calculated increase in yield
strength increase due to grain renement (25 MPa) was
determined to be smaller than the softening caused by
loss of precipitation hardening (84107 MPa). To oset
the yield strength decrease, the average grain size should
be smaller than 120 nm throughout the stirred zone.
The authors acknowledge the China Scholar-
ship Council (CSC, 2009612008) for nancial support;
Prof. John C. Lippold (metallography) and Dr. Henk
Colijn (electron microscopy) for providing access to
materials characterization laboratories; and Dr. Anil
Bhaskar Chaudhary and Dr. Matt Keller of Applied
Optimization for the modeling support.
[1] R.S. Mishra, M.W. Mahoney, S.X. McFadden, N.A.
Mara, A.K. Mukherjee, Scripta Mater. 42 (2000) 163.
[2] R.S. Mishra, M.W. Mahoney, Mater. Sci. Forum 357359
(2001) 507.
[3] Z.Y. Ma, S.R. Sharma, R.S. Mishra, M.W. Mahoney,
Metall. Mater. Trans. A 37 (2006) 3323.
[4] F.C. Liu, Z.Y. Ma, Scripta Mater. 59 (2008) 882.
[5] C.J. Lee, J.C. Huang, P.J. Hsieh, Scripta Mater. 54 (2006)
1415.
[6] C.G. Rhodes, M.W. Mahoney, W.H. Bingel, M. Cala-
brese, Scripta Mater. 48 (2003) 1451.
[7] J.Q. Su, T.W. Nelson, C.J. Sterling, J. Mater. Res. 18
(2003) 1757.
[8] J.Q. Su, T.W. Nelson, C.J. Sterling, Scripta Mater. 52
(2005) 135.
[9] D.C. Hofmann, K.S. Vecchio, Mater. Sci. Eng., A 402
(2005) 234.
[10] Y.J. Kwon, I. Shigematsu, N. Saito, Scripta Mater. 49
(2003) 785.
[11] M.J. Starink, A. Deschamps, S.C. Wang, Scripta Mater.
58 (2008) 377.
[12] M. Dixit, R.S. Mishra, K.K. Sankaran, Mater. Sci. Eng.,
A 478 (2008) 163.
[13] G. Liu, L.E. Murr, C.S. Niou, J.C. McClure, F.R. Vega,
Scripta Mater. 37 (1997) 355.
[14] Y.S. Sato, H. Kokawa, M. Enmoto, S. Jogan, Metall.
Mater. Trans. A 30 (1999) 2429.
[15] K.V. Jata, K.K. Sankaran, J.J. Ruschau, Metall. Mater.
Trans. A 31 (2000) 2181.
[16] R. Nandan, T. DebRoy, H.K.D.H. Bhadeshia, Prog.
Mater. Sci. 53 (2008) 980.
[17] J.Q. Su, T.W. Nelson, C.J. Sterling, Mater. Sci. Eng., A
405 (2005) 277.
[18] T.R. McNelley, S. Swaminathan, J.Q. Su, Scripta Mater.
58 (2008) 349.
[19] R.W. Fonda, J.F. Bingert, Scripta Mater. 57 (2007) 1052.
[20] A.W. Wilson, J.D. Madison, G. Spanos, Scripta Mater.
45 (2001) 1335.
[21] L.B. Ber, Mater. Sci. Eng., A 280 (2000) 91.
[22] P.A. Colgrove, H.R. Shercli, Sci. Technol. Weld. Join. 8
(2003) 360.
[23] Y.C. Chen, J.C. Feng, H.J. Liu, Mater. Character. 60
(2009) 476.
[24] N. Hansen, Scripta Mater. 51 (2004) 801.
[25] P.M. Kelly, Scripta Metall 6 (1972) 647.
[26] E. Hornbogen, E.A. Starke Jr., Acta Metall. Mater. 41
(1993) 1.
[27] S.D. Harkness, J.J. Hren, Met. Trans. 1 (1970) 43.
[28] A. Bigot, F. Danoix, P. Auger, D. Blavette, A. Reeves,
Mater. Sci. Forum 217222 (1996) 695.
[29] G.E. Dieter, Mechanical Metallurgy, third ed., McGraw-
Hill, New York, 1986.
1060 X. Feng et al. / Scripta Materialia 65 (2011) 10571060

Das könnte Ihnen auch gefallen