Vous êtes sur la page 1sur 12

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/282975092

Correlation between process parameters, grain


size, and hardness of friction stir welded Cu-Zn
alloys
Article in Rare Metals October 2015
Impact Factor: 1.01 DOI: 10.1007/s12598-016-0704-9

READS

38

2 authors:
Akbar Heidarzadeh

Tohid Saeid

Sahand University of Technology

Sahand University of Technology

37 PUBLICATIONS 247 CITATIONS

26 PUBLICATIONS 437 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

SEE PROFILE

Available from: Akbar Heidarzadeh


Retrieved on: 15 April 2016

RARE METALS

Rare Met.
DOI 10.1007/s12598-016-0704-9

www.editorialmanager.com/rmet

Correlation between process parameters, grain size and


hardness of friction-stir-welded CuZn alloys
Akbar Heidarzadeh, Tohid Saeid*

Received: 13 September 2015 / Revised: 10 October 2015 / Accepted: 2 March 2016


The Nonferrous Metals Society of China and Springer-Verlag Berlin Heidelberg 2016

Abstract In this study, the effects of tool rotational


speed, tool traverse speed, and Zn content on the grain size
and hardness of the friction-stir-welded (FSWed) CuZn
alloy joints were investigated. The microstructures of the
joints were examined using optical microscope (OM) and
scanning transmission electron microscope (STEM).
Vickers hardness test was conducted to evaluate the hardness of the joints. In addition, the relationships between the
process parameters, grain size, and hardness of the joints
were established. The results show that the developed
relationships predict the grain size and hardness of the
joints accurately. The Zn content of the alloys is the most
effective parameter on the grain size and hardness, where
the tool traverse speed has the minimum effect. The relationship between the hardness and grain size of the joints
has a deviation from the HallPetch equation due to formation of high dislocation density inside the grains. At
higher Zn amounts, the dislocation tangles with high density form instead of dislocation cells, and hence, lower
conformity with the HallPetch relationship is observed.
Keywords Friction stir welding; Grain size; Hardness;
CuZn alloy

1 Introduction
Copper and brasses (CuZn alloys) have vast industrial
applications because of their special characteristics such as
high electrical and thermal conductivities, good combinations
A. Heidarzadeh, T. Saeid*
Faculty of Materials Engineering, Sahand University of
Technology, Tabriz 51335/1996, Iran
e-mail: saeid@sut.ac.ir

of strength and ductility, and excellent resistance to corrosion.


Therefore, there is a large demand for welding of these types
of alloys. On the other hand, a high heat input condition is
needed during conventional fusion welding of the copper
alloys due to their high thermal conductivity. Consequently,
the higher heat input conditions dispose the joints to distortion,
solidification cracking, and high oxidation rate [13]. Also,
low boiling temperature of zinc leads to its evaporation during
fusion welding processes of brasses, which results in color
change and formation of a porous and weak layer of copper or
copper oxide. Moreover, it is notable that the zinc vapor is
toxic and can be harmful to the health of welding operators [4].
It seems that friction stir welding (FSW), as a solid-state
process, could be a good alternative to overcome the problems
arising from conventional fusion welding of copper and
CuZn alloys [5].
Despite variant investigations into the pure copper FSW
[614], the studies in the case of brass alloys are rather
limited [1518]. For example, Cam et al. [15] revealed that
the best combination of the strength and ductility in the
case of Cu10 wt% Zn and Cu30 wt% Zn brass alloys
could be achieved at a traverse speed of 210 mmmin-1
and a rotational speed of 1600 rmin-1. Emami and Saeid
[16] showed that the higher traverse speeds or lower
rotational speeds caused finer grain size of the Cu30 wt%
Zn joints and hence higher hardness values. Xie et al. [18]
demonstrated that the higher rotational speeds caused
coarser grain sizes in the joints of 5-mm-thick brass plates.
Furthermore, they found that increasing the rotational
speed had no noticeable effect on the tensile and yield
strengths, but it increased the elongation.
According to the above literatures, the FSW parameters
have a considerable effect on the microstructure and
mechanical properties of the brass alloy joints. Thus, the

123

A. Heidarzadeh, T. Saeid

correlation between FSW parameters, microstructure, and


mechanical properties of the brass joints can be very useful
for scientific and industrial applications. In this regard, one
of the applicable methods is response surface methodology
(RSM). Some workers have proved that the RSM can be
applied successfully for FSW of different metals and alloys
[1921]. For instance, Rajakumar et al. [19] established
mathematical models to evaluate the effect of different
FSW parameters on the grain size, ultimate tensile strength
(UTS), and hardness of the AA6061-T6 aluminum alloy
joints using RSM. They showed that the rotational speed of
1100 rmin-1, welding speed of 80 mmmin-1, axial force
of 8 kN, shoulder diameter of 15 mm, pin diameter of
5 mm, and tool hardness of HRC 45 resulted in the best
mechanical properties. Palanivel et al. [20] used RSM to
correlate the FSW parameters (tool rotational speed, traverse speed, and axial force) and UTS of AA5083-H111
aluminum alloy joints. They disclosed that the joint welded
at tool rotational speed of 1000 rmin-1, traverse speed of
69 mmmin-1, and axial force of 13.3 kN had higher UTS.
In addition to developing empirical relationships
between the FSW parameters and joint performances,
correlating the microstructural features and mechanical
properties is a key issue. One of the general methods to
determine the relationship between microstructure and
strength or hardness of the materials is HallPetch (HP)
equation. The HP equation in terms of hardness can be
expressed as follows [22]:
H H0 kd 1=2

where H is the hardness, d is the average grain size, and H0


and k are suitable constants associated with hardness measurements, respectively. Moreover, k is the slope of HP
equation, which indicates the relative strengthening contribution of grain boundaries. It has been revealed that the HP
equation needs to be modified in the case of severe plastic
deformation processes of the metals due to formation of the
substructures [22]. Thus, in the case of friction-stir-welded
alloys which undergo a severe plastic deformation, the HP
equation can be deviated from its linear relationship

(Eq. (1)). For example, Park et al. [23] studied the effect of
microstructure on HallPetch relationship in the case of the
friction-stir-welded (FSWed) thixomolded AZ91D Mg
alloy. They showed that the substructures affect the HP
relationship as well as the grain size of the joints.
Although some workers investigated the effect of FSW
on the brass joint properties, an investigation into the
correlation between FSW parameters, microstructure, and
hardness of the brass joints with different amounts of Zn
seems necessary. Furthermore, the HP equation for the
FSWed pure copper and brass joints has rarely been discussed according to both the grain size and substructure
effects. Therefore, in present study, three kinds of alloys
including pure copper, Cu30 wt% Zn, and Cu37 wt% Zn
brasses were friction-stir-welded under different tool rotational and traverse speeds. The relationships between FSW
parameters and joint features (grain size and hardness)
were established using RSM. Moreover, the hardness and
microstructure of the joints were correlated based on HP
relationship.

2 Experimental
The CuZn plates with different contents of Zn (0 wt%,
30 wt%, 37 wt%) were used as base metals (BMs) with
dimensions of 100 mm 9 100 mm 9 2 mm. The plates
were annealed at 500 C for 1 h. In order to produce a doublephase structure, the Cu37 wt% Zn BM was heated at 810 C
for 70 min and then was quenched in water at room temperature. Then, the plates were stress-relieved at 250 C for 1 h.
The microstructures of the different BMs are shown in Fig. 1.
The Design Expert software was used to design the
experiments and establish mathematical models. The analysis of variance (ANOVA) was performed to validate the
developed models. The considered parameters with their
levels and units and the experimental design matrix used in
this study are summarized in Tables 1 and 2. The plates were
friction-stir-welded at different rotational and traverse
speeds according to Table 2. In all of the experiments, a tool

Fig. 1 OM images of BM in alloys with different Zn contents: a 0 wt% Zn (pure copper), b 30 wt% Zn (single-phase brass), and c 37 wt% Zn
(double-phase brass)

123

Rare Met.

Correlation between process parameters, grain size and hardness of FSWed CuZn alloys

-1

quadratic, and interaction constant coefficients,


correspondingly.
Considering A, B, and C parameters, Eq. (3) can be
stated as Eq. (4):

Traverse speed (A)/(mmmin-1)

100

200

300

Rotational speed (B)/(rmin-1)

Y b0 b1 A b2 B b3 C b11 A2 b22 B2

450

700

900

30

37

Table 1 Coded and actual values of parameters


Parameters

Levels

Zn content (C)/(wt%)

with a cylindrical shoulder (12.0 mm in diameter) and a


simple cylindrical pin (3.0 mm in diameter and 1.7 mm in
length) made of H13 hot work tool steel were used. Also, the
tilt angle of the tool relative to the normal direction of the
plate surface was set constant at 2.5.
After FSW, the microstructures of the joints were studied
using optical microscope (OM, Olympus 100). The metallographic samples were cut from the joints transverse to the
welding direction and then polished and etched with a
solution of 20 ml nitric acid and 10 ml acetic acid. Clemex
image analysis software was applied to measure the average
grain size of the different joints. For deep physical study of
the joints, scanning transmission electron microscopy
(STEM, Hitachi S-4800) was used. The STEM samples were
thin-polished and then double-jet electro-polished using a
solution of HPO4:CH4O:H2O = 1:1:2 (volume ratio). The
Vickers hardness test was performed for hardness measurement in the center of the joints using load of 0.5 N for 10 s.

b33 C 2 b12 AB b13 AC b23 BC

The coefficients of Eqs. (3) or (4) were calculated by


Design Expert software using the following formulas [24
28]:
X
XX
b0 0:142857
Y  0:035714
Xii Y
5
X

Xi Y
6
bi 0:041667
X
XX
bii 0:03125
Xii Y 0:00372
Xii Y
X 
7
0:035714
Y
X
Xij Y
8
bij 0:0625
Finally, after coefficient calculations, the mathematical
models for Dav and hardness (H) of SZs have been
established as Eqs. (9) and (10), respectively,
Dav 8:80  1:13A 3:05B  3:34C  0:84AB
0:53AC  2:03BC  0:32A2 0:087B2  2:64C 2 9
H 117:90 7:47A  10:03B 16:39C 1:93AB
0:96AC 3:13BC 0:29A2 2:36B2  5:39C 2 10

3 Results and discussion


3.1 Empirical relationships
According to Eq. (2), the response parameter (Y), i.e.,
mean grain size (Dav) of the stir zone (SZ) or SZ hardness
is a function of input parameters, i.e., tool traverse speed
(A), tool rotational speed (B), and alloy type (here is Zn
content) (C). Moreover, Table 3 shows that the Design
Expert software suggests the quadratic model for both of
the responses. Therefore, in this study, the empirical relationships were developed by means of a second-order
polynomial regression model including the main and
interaction effects of the input parameters as indicated in
Eq. (3):
Yf A; B; C
Y b0

k
X
i1

2
bi k i

k
X
i1

bii Xi2

X
i\j

bij Xi Xj

where Xi and Xj are independent variables, b0 stands for the


mean value of responses, and bi, bii, and bij are linear,

Rare Met.

The experimental and predicted values (by Eqs. (9),


(10)) of Dav and hardness are summarized in Table 2. Also,
the normal plots of residuals and the predicted response
versus actual response plots are, respectively, illustrated in
Fig. 2ad, for the response Dav and hardness. The normal
probability plot indicates whether the residuals follow a
normal distribution or not, in which case the points will
follow a straight line. If the points do not follow a straight
line, it means that a transformation of the response may
provide a better analysis [29]. Figure 2a, b demonstrates
that errors are extended normally because the residuals
follow a straight line. Figure 2c, d reveals that the
predicted response values are in good agreement with the
actual ones within the ranges of the process parameters,
because the data points are split evenly by the 45 line. In
other words, there is a strong correlation between the
models predictions and its actual results.
The significance of the models and their coefficients can
be determined according to the ANOVA results as shown
in Tables 4 and 5. In ANOVA results, the F value, P value,
R2 and adjusted R2 can be used to recognize the significance of the models and coefficients, where F value is a
test for comparing curvature variance with residual (error)

123

A. Heidarzadeh, T. Saeid
Table 2 Design layout including experimental and predicted values
Nos.

Run

Dav/lm

Coded values of parameters


A

Experimental

Hardness (HV)
Predicted

Experimental

Predicted

15

-1

-1

-1

5.70

5.02

106.0

107.34

11

-1

-1

4.80

4.52

113.0

111.63

21

-1

-1

4.30

3.38

118.0

116.52
89.90

-1

-1

9.70

10.86

87.0

17

-1

9.50

9.51

96.1

96.12

16

-1

8.30

7.53

100.4

102.93

7
8

22
4

-1
0

1
1

-1
-1

17.10
14.20

16.86
14.67

80.3
88.8

77.18
85.33

-1

10.60

11.85

91.4

94.06

10

20

-1

-1

4.71

5.82

126.0

125.03

11

24

-1

4.66

5.85

131.0

130.29

12

-1

4.10

5.24

134.0

136.13

13

-1

9.34

9.62

113.0

110.72

14

13

9.16

8.80

119.0

117.90

15

7.40

7.36

128.0

125.66

16

23

-1

15.30

13.59

97.0

101.13

17

26

13.50

11.94

105.0

110.23

18

12

9.70

9.65

124.0

119.92

19

10

-1

-1

2.70

1.33

132.0

131.95

20

-1

2.30

1.89

139.0

138.16

21

27

-1

1.60

1.82

143.0

144.96

22
23

18
5

-1
0

0
0

1
1

3.10
2.60

3.10
2.82

120.0
126.0

120.76
128.90

24

25

2.40

1.91

141.0

137.62

25

-1

3.60

5.04

117.0

114.29

26

14

3.20

3.92

125.0

124.35

27

19

2.50

2.17

133.0

135.00

variance, P value is the probability of seeing the observed


F value if the null hypothesis is true, and R2 or R-squared is
a measure of the amount of variation around the mean
value explained by the model. In summary, larger F value,
R2 and adjusted R2, and lower P value disclose the more

significant model and coefficients [30]. According to the


ANOVA results for Dav (Table 4) and hardness (Table 5),
the F value, P value, R2 and adjusted R2 for the predicted
models of Dav, and hardness are 47.3300, \0.0001, 0.9617,
0.9414 and 91.8900, \0.0001, 0.9799, 0.9692,

Table 3 Results of different conducted models for responses of grain size and hardness, where 2FI being two-factor interaction model
Parameters
Grain size

Hardness

123

Source

P value

Linear

R2

Adjusted R2

Condition
Suggested

\0.0001

0.6581

0.7284

2FI

0.0028

0.7770

0.8429

Quadratic

0.0002

0.8900

0.9414

Cubic

0.0008

0.9459

0.9878

Linear

\0.0001

0.9079

0.9247

2FI

0.0507

0.9216

0.9408

Quadratic

0.0026

0.9482

0.9692

Cubic

0.8179

0.8636

0.9611

Suggested

Rare Met.

Correlation between process parameters, grain size and hardness of FSWed CuZn alloys

Fig. 2 Normal plots of residuals a, c and predicted response versus actual response plots b, d for responses: a, b Dav, and c, d hardness
Table 4 ANOVA data for response Dav
Source

Sum of squares

Model

Degree of freedom

Mean square

F value

P value

Condition
Significant

495.370

55.040

47.370

\0.0001

23.010

23.010

19.800

0.0004

167.020

167.020

143.740

\0.0001

201.340

201.340

173.270

\0.0001

AB

8.480

8.480

7.300

0.0151

AC

3.410

3.410

2.940

0.1047

BC

49.610

49.610

42.700

\0.0001

A2

0.600

0.600

0.520

0.4823

B2

0.046

0.046

0.039

0.8452

41.850

41.850

36.020

\0.0001

19.75
0.9617

17

1.160

Residual
R2
Adjusted R2

0.9414

correspondingly. Thus, the results show that the developed


models predict the responses adequately.
Furthermore, in ANOVA tables, P values of smaller
than 0.0500 confirm that a coefficient is significant,
and P values of larger than 0.1000 validate that a
coefficient is not significant [31]. Consequently, A, B, C,
AB, BC, and C2 are significant terms for both of the
predicted models. Therefore, by considering the
Rare Met.

significant terms only, the final relationships can be


stated as follows:
Dav 8:65  1:13A 3:05B  3:34C  0:84AB
2:03BC  2:64C 2

11

H 119:67 7:47A  10:03B 16:39C


1:92AB 3:12BC  5:39C 2

12

123

A. Heidarzadeh, T. Saeid
Table 5 ANOVA data for response hardness
Source

Sum of squares

Degree of freedom

Mean square

F value

P value

Condition
Significant

Model

8030.64

892.29

91.89

\0.0001

1005.01

1005.01

103.50

\0.0001

1810.01

1810.01

186.40

\0.0001

4834.72

4834.72

497.89

\0.0001

AB

44.47

44.47

4.58

0.0472

AC

11.02

11.02

1.13

0.3016

BC

117.19

117.19

12.07

0.0029

A2

0.52

0.52

0.05

0.8197

B2

33.45

33.45

3.44

0.0809

174.24

174.24

17.94

0.0006

17

9.71

Residual
R2
Adjusted R2

165.08
0.9799
0.9692

Fig. 3 Counters at different conditions ac and perturbation plots d for response Dav. Numbers in ac being values of Dav

The F values demonstrate that the order of more


significant terms in the relationships developed for Dav and
hardness (Eqs. (11), (12)) are as follows, respectively:
C [ B [ A [ BC [ C2 [ AB and C [ B [ A [ C2 [
BC [ AB.

123

3.2 Effect of parameters on Dav and hardness


The contour and perturbation plots for Dav are illustrated in
Fig. 3ad. As well, the microstructures of the joints welded
at different welding conditions are shown in Fig. 4.

Rare Met.

Correlation between process parameters, grain size and hardness of FSWed CuZn alloys

Fig. 4 OM images of joints welded under different welding conditions corresponding to experimental Nos. in Table 2: lower heat input, a No. 1
(0 wt% Zn), b No. 10 (30 wt% Zn), and c No. 19 (37 wt% Zn); higher heat input, d No. 7 (0 wt% Zn), e No. 16 (30 wt% Zn), and f No. 25
(37 wt% Zn)

According to Figs. 3 and 4, larger traverse speeds and


lower rotational speeds cause smaller Dav. Based on the
fine and equiaxed grains in the SZ of the joints, it can be
concluded that dynamic recrystallization (DRX) occurs
during FSW. Since FSW is a hot deformation process due
to the existence of heat and deformation, the grain size of
the joints will be controlled by thermomechanical parameters such as strain rate and temperature [32]. The strain
rate and temperature during thermomechanical processes
can be related by ZenerHolloman as follows [33]:
Z e_ Q=RT

13

where Z is ZenerHolloman parameter, e_ is strain rate, T is


temperature, Q is activation energy, and R is the gas
constant. e_ and T can be calculated using following
equations, respectively [33]:
e_ Rm  2pre =Le
 2 a
T
x
2
k
Tm
104  t

14
15

where Rm, re, and Le, respectively, denote the half of tool
rotational speed, the effective radius, and depth of the
dynamically recrystallized zone; k and a are constants
between 0.040.06 and 0.650.75, respectively; x is tool
rotational speed; t is tool traverse speed; and Tm is the
melting point of the alloy [34]. In addition, it has been
demonstrated that the grain size during thermomechanical
processes has an inverse relationship with Z. Therefore,

Rare Met.

according to the Eqs. ((13)(15)), larger e_ causes smaller


Dav, where higher T leads to bigger Dav. It can be concluded that e_ and T are challenging in determining the final
Dav during the thermomechanical processes of the alloys.
In this study, according to Figs. 3 and 4, with the rotational
_ and the traverse
speeds increasing (higher T and lower e)
speeds decreasing (higher T), Dav increases continuously.
As a result, in present investigation, the dominant factor
which controls the final Dav is T.
Furthermore, Figs. 3 and 4 reveal that the higher Zn
content of the alloys causes smaller Dav under the same
process condition. This can be explained by the effect of Zn
on the stacking fault energy (SFE) of the different alloys. The
effect of SFE on the steady state minimum grain size (dmin) of
the severely plastic deformed (SPDed) materials can be
stated using the model developed by Mohamed [35]:
c q
dmin
SFE
a
16
b
Gb
where b is Burgers vector, cSFE stands for amount of SFE,
G is shear modulus, and a and q are constants. The dmin
value of the materials processed by different SPD methods
agrees with Mohamed model [3639]. Recently, Morishige
et al. [40] have investigated the effect of Mg content on
dmin value of the friction-stir-processed (FSPed) 5052 and
5058 aluminum alloys. They showed that with the Mg
content increasing, the dmin value of the FSPed alloys
decreased due to lower SFE. According to the relationship
reported by Gallagher [41], the effect of Zn content on the

123

A. Heidarzadeh, T. Saeid

Fig. 5 Counters at different conditions ac and perturbation plots d for response hardness. Numbers in ac being values of hardness

SFE of the CuZn alloys (cCuZn ) could be expressed


through the following equation:
( 
2 )
xZn =xZn
cCuZn c0 : exp kc
17
1 xZn =xZn
where c0 stands for SFE of the pure copper, xZn is Zn
concentration, xZn denotes the solubility limit of Zn at high
temperature, and kc is a dimensionless constant. Equations (16) and (17) reveal that the grain refinement can be
achieved with Zn content increasing in CuZn alloys during FSW. Similar results have been reported for the CuZn
alloys processed by high-pressure torsion (HPT) and by
HPT followed by cold-rolling [42]. It seems that this effect
of Zn in CuZn alloys is very similar to that of Mg in Al
Mg alloys [40], which is attributed to the reduction in
dislocation mobility and hence recovery rate in these types
of solid solution alloys.
3.3 Hardness-grain size correlation
The effects of FSW parameters and Zn content on the
hardness of the joints are illustrated in Fig. 5 using counter
and perturbation plots. According to Fig. 5, lower

123

rotational speeds and higher traverse speeds cause higher


hardness values, which is due to smaller grain size as
explained in Sect. 3.2 (Figs. 3, 4). Furthermore, at constant
FSW parameters, the hardness values increase with Zn
content of the alloys increasing. This can be due to the
smaller grain size (Figs. 3, 4) and solid solution strengthening effect of Zn in alloys with higher amounts of Zn.
For correlation between hardness and grain size of the
joints, the HP relationships were estimated according to
the data in Table 2. As shown in Fig. 6, the HP relationships for the joints of different alloys can be stated as
follows:
H0 wt% Zn 46:6 145:9d1=2

18

H30 wt% Zn 74:3 124:1d

1=2

19

H37 wt% Zn 65:3 104:9d

1=2

20

where d refers to mean grain size of the joints. According


to the R2 values (0.94, 0.75, and 0.74, respectively, for
alloys with 0 wt%, 30 wt%, 37 wt% Zn), the HP equations have deviation from their linear relationships. The
reason of this deviation is the fact that in the HP relationship, only the high angle grain boundaries are considered as obstacles to the dislocation movement [43]. Thus,

Rare Met.

Correlation between process parameters, grain size and hardness of FSWed CuZn alloys

Fig. 6 Plots for HP relationships of alloys with different Zn contents: a 0 wt% Zn (pure copper), b 30 wt% Zn (single-phase brass), and
c 37 wt% Zn (double-phase brass)

Fig. 7 STEM images of joints for alloys with different Zn contents: a 0 wt% Zn (pure copper), b 30 wt% Zn (single-phase brass), and c 37 wt%
Zn (double-phase brass)

the existence of substructures such as different dislocation


structures, precipitates, and second phase particles can
affect the HP relationship [4447]. These microstructural
features prevent the dislocation movement and pin them at
an interval smaller than the grain size and hence reduce the
effect of grain size on the hardness.
Some investigations have shown that the FSW causes
formation of fine and equiaxed grains with high density of
dislocations [2, 5]. The STEM images of the joints for
different alloys are shown in Fig. 7, revealing that the grain
interiors contain a high density of dislocations. According
to Fig. 7, the high density of dislocations, especially in the
case of Cu30 wt% Zn and Cu37 wt% Zn alloys, causes
deviation from linear HP relationship. Furthermore, from
R2 values and slope of the HP relationships shown in
Fig. 6 (146, 124, and 105 are slop values, respectively, for
alloys with 0 wt%, 30 wt%, and 37 wt% Zn), with the Zn
content increasing (lower slope and R2), the effect of grain
size decreases in the HP equation. The lower SFE of the
alloys containing higher Zn amounts prevents the dislocation mobility and annihilation and hence causes formation
of high density of dislocation with tangle structures
(Fig. 7b, c). In contrast, higher SFE of the pure copper
results in more mobility and annihilation of dislocations

Rare Met.

and therefore lower dislocation density with cell structures


(Fig. 7a). It seems that the tangle structure of the dislocations (Fig. 7b, c) pins the mobile dislocations more than the
annihilated and cell structures (Fig. 7a).

4 Conclusion
In this study, the effects of FSW parameters and Zn content
on the Dav and hardness of the CuZn alloy joints were
investigated, and the relationships between parameters and
responses were correlated. RSM was used to correlate the
process parameters (tool traverse speed, tool rotational
speed, and Zn content of the alloys) and the responses (Dav
and hardness). The ANOVA data show that the developed
relationships can predict the responses accurately. The order
of the more significant and effective parameters on the Dav
and hardness of the joints are as follows: Zn content [ tool
rotational speed [ tool traverse speed. The effect of Zn is
due to its influence on the SFE of the CuZn alloys, where the
effects of tool traverse and rotational speeds result from their
influence on the ZenerHolloman parameter. In addition, the
relationships between hardness and Dav of the joints show a
deviation from HP equation. The origin of this deviation is

123

A. Heidarzadeh, T. Saeid

the formation of dislocations with high density in grain


interiors. According to the slope and R2 of HP equation, the
pure copper (0 wt% Zn) joints have more conformity with
the HP linear relationship compared to those of the brass
(30 wt% and 37 wt% Zn) joints. This behavior is due to the
effect of Zn content on the SFE and hence its influence on the
formation of different dislocation structures, i.e., dislocation
cells and tangles in the case of pure copper and brass joints,
respectively.

References
[1] Nandan R, DebRoy T, Bhadeshia HKDH. Recent advances in
friction-stir welding-process, weldment structure and properties.
Prog Mater Sci. 2008;53(6):980.
[2] Heidarzadeh A, Saeid T. A comparative study of microstructure
and mechanical properties between friction stir welded single
and double phase brass alloys. Mater Sci Eng A. 2016;649:349.
[3] Cam G. Friction stir welded structural materials: beyond Alalloys. Int Mater Rev. 2011;56(1):1.
[4] Heidarzadeh A, Saeid T. On the effect of b phase on the
microstructure and mechanical properties of friction stir welded
commercial brass alloys. Data in Brief. 2015;5:1022.
[5] Heidarzadeh A, Kazemi-Choobi K, Hanifian H, Asadi P. 3-Microstructural Evolution. In: Givi MKB, Asadi P, editors.
Advances in Friction-Stir Welding and Processing. Cambridge:
Woodhead Publishing; 2014. 65.
[6] Ambroziak A. Hydrogen damage in friction welded copper
joints. Mater Des. 2010;31(8):3869.
[7] Farrokhi H, Heidarzadeh A, Saeid T. Frictions stir welding of
copper under different welding parameters and media. Sci
Technol Weld Join. 2013;18(8):697.
[8] Galvao I, Leal RM, Rodrigues DM, Loureiro A. Influence of
tool shoulder geometry on properties of friction stir welds in thin
copper sheets. J Mater Process Technol. 2013;213(2):129.
[9] Heidarzadeh A, Jabbari M, Esmaily M. Prediction of grain size
and mechanical properties in friction stir welded pure copper
joints using a thermal model. Int J Adv Manuf Technol.
2015;77(912):1819.
[10] Hwang YM, Fan PL, Lin CH. Experimental study on friction stir
welding of copper metals. J Mater Process Technol. 2010;210(12):1667.
[11] Leal RM, Sakharova N, Vilaca P, Rodrigues DM, Loureiro A.
Effect of shoulder cavity and welding parameters on friction stir
welding of thin copper sheets. Sci Technol Weld Joining.
2011;16(2):146.
[12] Shen JJ, Liu HJ, Cui F. Effect of welding speed on
microstructure and mechanical properties of friction stir welded
copper. Mater Des. 2010;31(8):3937.
[13] Sun YF, Fujii H. Investigation of the welding parameter
dependent microstructure and mechanical properties of friction
stir welded pure copper. Mater Sci Eng A. 2010;527(26):6879.
[14] Xu N, Ueji R, Morisada Y, Fujii H. Modification of mechanical
properties of friction stir welded Cu joint by additional liquid
CO2 cooling. Mater Des. 2014;56:20.
[15] Cam G, Serindag HT, Cakan A, Mistikoglu S, Yavuz H. The
effect of weld parameters on friction stir welding of brass plates.
Materialwiss Werkstofftech. 2008;39(6):394.
[16] Emami S, Saeid T. Effects of welding and rotational speeds on
the microstructure and hardness of friction stir welded singlephase brass. Acta Metall Sin (Engl Lett). 2015;28(6):766.

123

[17] Sun YF, Xu N, Fujii H. The microstructure and mechanical


properties of friction stir welded Cu30Zn brass alloys. Mater
Sci Eng A. 2014;589:228.
[18] Xie GM, Ma ZY, Geng L. Effects of friction stir welding
parameters on microstructures and mechanical properties of
brass joints. Mater Trans. 2008;49(7):1698.
[19] Rajakumar S, Muralidharan C, Balasubramanian V. Predicting
tensile strength, hardness and corrosion rate of friction stir welded
AA6061T6 aluminium alloy joints. Mater Des. 2011;32(5):2878.
[20] Palanivel R, Koshy Mathews P. Prediction and optimization of
process parameter of friction stir welded AA5083-H111 aluminum alloy using response surface methodology. J Cent South
Univ. 2012;19(1):1.
[21] Rajakumar S, Balasubramanian V. Establishing relationships
between mechanical properties of aluminium alloys and optimised
friction stir welding process parameters. Mater Des. 2012;40:17.
[22] Ghassemali E, Tan M-J, Wah CB, Lim SCV, Jarfors AEW.
Effect of cold-work on the HallPetch breakdown in copper
based micro-components. Mech Mater. 2015;80:124.
[23] Park S, Sato Y, Kokawa H. Microstructural evolution and its
effect on HallPetch relationship in friction stir welding of
thixomolded Mg alloy AZ91D. J Mater Sci. 2003;38(21):4379.
[24] Azadbeh M, Mohammadzadeh A, Danninger H. Modeling the
response of physical and mechanical properties of CrMo prealloyed sintered steels to key manufacturing parameters. Mater
Des. 2014;55:633.
[25] Azadbeh M, Mohammadzadeh A, Danninger H, Gierl-Mayer C.
On the densification and elastic modulus of sintered CrMo
steels. Metall Mater Trans B. 2015;46(3):1471.
[26] Mohammadzadeh A, Azadbeh M, Danninger H. New concept in
analysis of supersolidus liquid phase sintering of alpha brass.
Powder Metall. 2015;58(2):123.
[27] Mohammadzadeh A, Azadbeh M, Danninger H. Microstructural
coarsening during supersolidus liquid phase sintering of alpha
brass. Powder Metall. 2015;58(4):300.
[28] Mohammadzadeh A, Azadbeh M, Namini SA. Densification and
volumetric change during supersolidus liquid phase sintering of
prealloyed brass Cu28Zn powder: modeling and optimization.
Sci Sinter. 2014;46(1):23.
[29] Mozammel M, Mohammadzadeh A. The influence of pre-oxidation and leaching parameters on Iranian ilmenite concentrate
leaching efficiency: optimization and measurement. Measurement. 2015;66:184.
[30] Heidarzadeh A, Barenji R, Esmaily M, Ilkhichi A. Tensile
properties of friction stir welds of AA 7020 aluminum alloy.
Trans Indian Inst Met. 2015;68(5):757.
[31] Heidarzadeh A, Saeid T, Khodaverdizadeh H, Mahmoudi A,
Nazari E. Establishing a mathematical model to predict the
tensile strength of friction stir welded pure copper joints. Metall
Mater Trans B. 2013;44(1):175.
[32] McNelley TR, Swaminathan S, Su JQ. Recrystallization mechanisms during friction stir welding/processing of aluminum
alloys. Scr Mater. 2008;58(5):349.
[33] Commin L, Dumont M, Masse JE, Barrallier L. Friction stir
welding of AZ31 magnesium alloy rolled sheets: influence of
processing parameters. Acta Mater. 2009;57(2):326.
[34] Arora A, De A, DebRoy T. Toward optimum friction stir
welding tool shoulder diameter. Scr Mater. 2011;64(1):9.
[35] Mohamed FA. A dislocation model for the minimum grain size
obtainable by milling. Acta Mater. 2003;51(14):4107.
[36] Parvin H, Kazeminezhad M. Dependency modeling of steady
state grain size on the stacking fault energy through severe
plastic deformation. Mater Lett. 2015;159:410.
[37] Qu S, An XH, Yang HJ, Huang CX, Yang G, Zang QS, Wang
ZG, Wu SD, Zhang ZF. Microstructural evolution and

Rare Met.

Correlation between process parameters, grain size and hardness of FSWed CuZn alloys

[38]

[39]

[40]

[41]

[42]

mechanical properties of CuAl alloys subjected to equal


channel angular pressing. Acta Mater. 2009;57(5):1586.
Huang CX, Hu W, Yang G, Zhang ZF, Wu SD, Wang QY,
Gottstein G. The effect of stacking fault energy on equilibrium
grain size and tensile properties of nanostructured copper and
copperaluminum alloys processed by equal channel angular
pressing. Mater Sci Eng A. 2012;556:638.
Cai B, Tao J, Wang W, Yang X, Gong Y, Cheng L, Zhu X. The
effect of stacking fault energy on equilibrium grain size and
tensile properties of ultrafine-grained CuAlZn alloys processed by rolling. J Alloys Compd. 2014;610:224.
Morishige T, Hirata T, Uesugi T, Takigawa Y, Tsujikawa M,
Higashi K. Effect of Mg content on the minimum grain size of
AlMg alloys obtained by friction stir processing. Scr Mater.
2011;64(4):355.
Gallagher PCJ. The influence of alloying, temperature, and
related effects on the stacking fault energy. Metall Mater Trans
B. 1970;1(9):2429.
Zhao YH, Zhu YT, Liao XZ, Horita Z, Langdon TG. Influence
of stacking fault energy on the minimum grain size achieved in
severe plastic deformation. Mater Sci Eng A. 2007;463(12):22.

Rare Met.

[43] Du D, Fu R, Li Y, Jing L, Wang J, Ren Y, Yang K. Modification


of the HallPetch equation for friction-stir-processing
microstructures of high-nitrogen steel. Mater Sci Eng A.
2015;640:190.
[44] Gao C, Zhu Z, Han J, Li H. Correlation of microstructure and
mechanical properties in friction stir welded 2198-T8 AlLi
alloy. Mater Sci Eng A. 2015;639:489.
[45] Sato Y, Park S, Kokawa H. Microstructural factors governing
hardness in friction-stir welds of solid-solution-hardened Al
alloys. Metall Mater Trans A. 2001;32(12):3033.
[46] Jones MJ, Heurtier P, Desrayaud C, Montheillet F, Allehaux D,
Driver JH. Correlation between microstructure and microhardness in a friction stir welded 2024 aluminium alloy. Scr Mater.
2005;52(8):693.
[47] Sato YS, Urata M, Kokawa H, Ikeda K. HallPetch relationship
in friction stir welds of equal channel angular-pressed aluminium alloys. Mater Sci Eng A. 2003;354(12):298.

123

Vous aimerez peut-être aussi