Sei sulla pagina 1di 40

Accepted Manuscript

Title: Friction stir welding of similar and dissimilar AA7075


and AA5083

Author: M.M.Z. Ahmed Sabbah Ataya M.M. El-Sayed


Seleman H.R. Ammar Essam Ahmed

PII: S0924-0136(16)30409-5
DOI: http://dx.doi.org/doi:10.1016/j.jmatprotec.2016.11.024
Reference: PROTEC 15025

To appear in: Journal of Materials Processing Technology

Received date: 17-9-2016


Revised date: 19-11-2016
Accepted date: 21-11-2016

Please cite this article as: Ahmed, M.M.Z., Ataya, Sabbah, El-Sayed Seleman,
M.M., Ammar, H.R., Ahmed, Essam, Friction stir welding of similar and
dissimilar AA7075 and AA5083.Journal of Materials Processing Technology
http://dx.doi.org/10.1016/j.jmatprotec.2016.11.024

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Friction stir welding of similar and dissimilar AA7075 and AA5083

M.M.Z. Ahmeda,b,c*, Sabbah Atayaa,b,d, M.M. El-Sayed Selemana,b, H.R. Ammara,b,e, Essam

Ahmeda,b

a
Suez and Sinai Metallurgical and Materials Research Center of Scientific Excellence (SSMMR-CSE), Suez

University, Suez, Egypt

b
Metallurgical and Materials Engineering Department, Faculty of Petroleum and Mining Engineering, Suez

University, 43721 Suez, Egypt.

c
Mechanical Engineering Department, The British University in Egypt, El-Sherouk 11837, Cairo, Egypt.

d
Department of Mechanical Engineering, College of Engineering, Al Imam Mohammad Ibn Saud Islamic University,

Riyadh 11432, KSA

e
Mechanical Engineering Department, College of Engineering, Qassim University, Buraydah, 51452, KSA

* mohamed.zaky@suezuniv.edu.eg ; mohamed.zaky@bue.edu.eg

1
Abstract
Aluminum alloys AA7075-T6 and AA5083-H111 were friction stir welded at constant
rotation rate of 300 rpm and different traverse speeds of 50,100, 150, and 200mm/min in
similar and dissimilar joints. The microstructures and crystallographic textures of base
materials (BM) and the welds were investigated using electron backscatter diffraction
(EBSD) technique. The mechanical properties were evaluated using hardness and tensile
testing. The fracture surface of the tensile tested samples was examined using scanning
electron microscope (SEM). Microstructural examination of AA7075 and AA5083 BMs
showed significant difference in the grain structure of the two alloys with average grain
size of 40µm that was free of substructure in case of AA7075. While AA5083 BM revealed
an average grain size of 25µm with high density of substructure. In addition the type
precipitates also are different in the two alloys. Although, the two alloys welded using the
same FSW parameters the two alloys exhibited different response in terms of the
recrystallized fine grains after FSW. Significant grain refining occurred in the NG zone of
AA7075 with average grain size of 6µm at 50mm/min welding speed that was reduced to
2 µm by increasing the welding speed to 200mm/min. A relatively recrystallized coarser
grain structure obtained in AA5083 joints NG zone with average grain size of 9 µm at 50
mm/min that is reduced to 3µm at 200mm/min. This implies that the starting material
characteristics have a significant impact on the final grain structure after FSW. The
crystallographic texture in the nugget zone revealed a simple shear texture without a
significant influence of varying the welding speed. The dissimilar joints exhibited ultimate
tensile strength ranged between 245 and 267 MPa and fracture strain ranged between 3
and 5.6 %.A brittle/ductile fracture mode was dominating in the examined fracture surface
of the dissimilar joints.

Key Words: Friction stir welding; Aluminum alloys; Grain structure; crystallographic
texture; mechanical properties; EBSD

1. Introduction
FSW of dissimilar alloys is controlled by several process parameters. Numerous studies
were carried out on the influence of these variables on the properties of the FSWed joints
of dissimilar Al-alloys. Sivachidambaram et al. (2015) investigated the influence of FSW
variables of tool rotation rates (700, 800, and 900 rpm) and transverse speeds (40, 60,
and 80 mm/min) on joining of dissimilar Al alloys AA 5383 and AA7075. They reported

2
that the 700 rpm and 40 mm/min welding condition yields the best combination of strength
and hardness of the weldment. Shojaeefard et al. (2013) developed an artificial neural
network (ANN) model to simulate the correlation between the applied FSW variables and
the obtained mechanical properties of the FSW of AA7075-O to AA5083-O Al alloys. They
reported that defects such as tunnel and pin hole were formed in dissimilar FSWed
AA5083-O and AA7075-O Al alloys of 6mm thickness at rotational speeds from 500
to1600rpm and traverse speeds from 20 to 96mm/min. They Also found that a good
correlation exists between the predicted data obtained from an ANN model and the
measured results.

Palanivelet et al. (2012) studied the influence of tool rotational speed and pin profile on
the strength and microstructure of FSWed AA5083-H111and AA6351-T6 Al alloys. They
reported that, the rotational speed and pin profile was found to be significantly affect the
joints strength and microstructure and the difference in tensile strength of the dissimilar
welds was related to the behavior of material flow, elimination of cold working effect in the
heat affected zone (HAZ) of AA5083, dissolution of second phase precipitates in AA6351
alloy and the imperfections formation in the weld zone. Gan et al. (2008) studied hardness
and tensile properties of FSWed 5083-H18 and 6111-T4 Al alloys. It was found that the
weld zones (WZs) in AA6111 alloy displayed a comparable strength to the base material
(BM), while, the WZs in the AA5083-H18 alloy was noticed to be softer than the BM. This
softening is related to recovery of the strain hardened structure as a result of the heat
generated from FSW process in dissimilar FSWed of Al alloys of 2024-T3 and 7075-T6
(Khodir and Shibayanagi, 2008). The increase in the welding speed was observed to form
porosity and kissing bond, in particular, when the position of 2024 Al alloy on the retreating
side. It was also reported that patterns in the form of onion ring were obtained and
described by bands of various equiaxed grain sizes and inhomogeneous scattering of
alloying elements in the stir zone regardless of welding speed and fixed locations of BM

Dilip et al. (2010) investigated the microstructures, hardness, and tensile properties of
dissimilar friction stir welded joints between AA2219-T87 and AA5083-H321. The lowest
hardness was found in the HAZ on the side of 5083 alloy where fracture was observed to

3
occur. Tensile test results showed that a joint efficiency of 90% could be achieved which
is markedly higher than that obtained with traditional welding techniques.

Gungor et al. (2014) studied the microstructure and mechanical properties of FSWed
5083-H111 and 6082-T651 Al alloys. They concluded that the joints with improved fatigue
life and reasonable strength could be attained by applying FSW at low speed. It was
observed from the microstructural analysis that a defect-free structure was obtained in
the weld zone of the processed alloys. Rodriguez et al. (2015) investigated the
microstructure and mechanical properties of FSW of AA6061toAA7050. They observed
that irrespective of tool rotational speed, microhardness values displayed an irregular
hardness profile across the weld nugget, as a result of the discrete mechanical properties
of the two alloys. Furthermore, it was concluded that the rotational speed displayed a
direct effect on the level of materials intermixing at the stir zone,

The effect of shoulder diameter to pin diameter ratio (D/d ratio) on mechanical properties
and microstructures of the dissimilar FSW Al alloy butt welds (AA2014-T6/AA7075-T6)
was studied by Saravanan et al. (2015), it was showed that the D/d ratio greatly influences
the mechanical properties of the welds. Kasman and Yenier (2014) examined the effect
of FSW parameters on the mechanical properties of dissimilar FSWed AA5754-to-
AA7075 alloys. They reported that increasing welding speed or tool rotational rate
resulted in decreasing the strength of the joints. The distribution profile of hardness values
of the weld zones was observed to vary according to the welding speed and the rotational
rate.

Although there are few studies on the similar and dissimilar joining of the AA5083 and
AA7075 alloys, there is still a need for more understating the relationship between the
microstructural changes and the obtained texture through FSW process and the
mechanical properties at a wide range of welding traverse speeds.

The present study investigated the influence of FSW traverse speeds on the
microstructure and mechanical properties of similar and dissimilar AA5083 and AA7075
joints. EBSD is used as a main technique to evaluate the microstructure and

4
crystallographic texture in the NG zone of the joints. The mechanical properties were
investigated and interrelated with the microstructural features.

2. Experimental Procedure
2.1. Friction Stir Welding procedure
Rolled plates of AA7075-T6 and AA5083-H111 of 5 and 6 mm thickness respectively were
friction stir welded in similar and dissimilar butt joints (AA7075/AA7075, AA5083/AA5083,
and AA7075/AA5083).The plate dimensions were 100mm X 200mm to produce a butt
joint with the dimensions of 200mm X 200mm. The chemical composition of the alloys is
listed in Table 1. Four welding speeds of 50, 100, 150 and 200 mm/min were applied. The
tool rotational rate and the tool tilt angle were fixed at 300 rpm and 3ᴼ, respectively. A tool
with 18 mm diameter concave shoulder and 4.8 mm long unthreaded taper cylindrical pin
was used. The top and bottom diameter of the pin was 3 mm and 7 mm, respectively. The
tool was machined from H13 steel and heat treated to achieve 58 HRC. In dissimilar
joints, the position of the AA5083-H111 alloy was at the advancing side (AS), while, the
location of the AA7075-T6 alloy was at the retreating side (RS).The tool was rotated in
the clock wise direction. For all joints, the unthreaded tool was plunged at the center line
until the shoulder touches the surface of the plates and then the tool was moving towards
so that weld could be accomplished. In case of the dissimilar joints the tool was plunged
until the shoulder touches the 5mm thickness plate and then traverse along the joint line
with some excess material swept around due to the 1mm thickness difference between
the two plates.

The resulting butt joints (200 mm x 200 mm) were cut into strips with sufficient width
perpendicular to the welding line to produce: three tensile test specimens, sample for the
metallographic investigation and sample for hardness testing.

2.2. Microstructural Characterization

5
The microstructure of the base and welded materials were characterized using optical
microscopy (OM) and scanning electron microscopy (SEM). The electron backscatter
diffraction (EBSD) technique in a Quanta FEG 250 SEM equipped with Hikari EDAX-
EBSD camera controlled by orientation image microscopy data collection software (OIM
DC 7.2) was used to investigate the grain structure and crystallographic texture. The
investigated specimens cut from the transverse cross section of each weld were ground
and polished according to the standard procedures and then etched with Keller's reagent
for OM. For EBSD the samples after mechanical polishing were further electropolished
for 60 s at minus 15°C and 14 volt. The electrolyte solution contained 30 vol.% nitric acid
and 70 vol.% ethanol. Fracture surface analysis of the tension tested samples was carried
out using SEM. This analysis aimed at examining the nature of the fracture of the welded
materials at the selected FSW parameters.

2.3 Mechanical Testing


Tensile test specimens (5 mm thickness) for similar and dissimilar joints were machined
according to the DIN EN10002-1 2001 (D) standards. The tensile test was performed at
room temperature using an Instron 4210 testing machine with a cross-head speed of 0.06
mm/sec. Three samples were tested for each welding condition. For comparison, the as-
received alloys were also tested. The hardness test was carried out using Vickers
hardness tester, model HWDV-7S. The hardness profile for each sample was generated
using a load of 2 kgf with a dwell time of 15 s. Hardness values were measured on the
center line along the cross section perpendicular to the welding direction and the distance
between each indentation was set to 0.5 mm.
3. Results and Discussion
3.1 Macrographs
Figure 1 (a), (b) and (c) show the optical macrographs of the polished transverse cross
sections of the FSWed joints of similar AA7075-T6, similar AA5083-H111, and dissimilar
AA7075-AA5083 Al alloys, respectively. From these macrographs, the typical FSW
features in Al alloys such as sharp AS interface and diffusive RS interface can be
observed. Furthermore, it may be noted in Fig.1 that all alloys joints exhibited a defect-
free welds within the plunge distance at the rotational speed of 300 rpm and traverse

6
speeds of 50, 100, 150, and 200 mm/min. This can be attributed to the low heat input
combination at the selected FSW parameters that resulted in the appropriate tool/material
contact condition (sticking condition) for producing sound joints. On the other hand the
use of high heat input conditions such as high rotation rate and low traverse speed can
result in inappropriate tool/material contact conditions (slipping condition) which can
produce joints with some types of defects Leitao et al. (2012). For example Shojaeefard
et al.(2013) reported that defects such as tunnel and pin hole were formed in dissimilar
FSWed AA5083-O and AA7075-O Al alloys of 6mm thickness at rotational speeds from
500 to1600rpm and traverse speeds from 20 to 96mm/min. In addition, it should be
mentioned that the tool of a 4.8 mm length resulted in a full penetration in similar AA7075
alloy joints of 5 mm plate thickness, as may be seen in Fig. 1 (a),whereas, a lack of
penetration around 1mm depth in the similar AA5083 and the dissimilar joints, as may be
observed in Fig. 1 (b) and (c).This lack of penetration of the welded joints is ascribed to
thickness difference between the two alloys and the use of the same FSW tool for joining
in similar and dissimilar joints.

An example of the optical microstructure of AA7075 base material, at the AS and RS


interface and inside the NG zone of the joint produced at 200mm/min welding speed is
shown in Figure 2 (a-d). Significant grain refining can be noted in the NG zone relative to
the coarse grain structure of base material as can be observed from Figure 2a and c.
From Figure 2 b and d the diffusive interface at the RS and the sharp interface at the AS
can be observed as indicated by the dashed white lines.

3.2. Base materials grain structure and texture


The starting materials were investigated in terms of grain structure and texture. As
mentioned in section 2.1, the as-received base materials (BMs); AA7075 and AA5083 Al
alloys are chemically different. Figure 3 shows the inverse pole figure (IPF) coloring with
respect to the rolling direction and grain boundary (GB) maps of (a) AA7075-T6 and (b)
AA5083-H111 Al alloys. It can be seen, that the BMs are also different in terms of grain
structure and low angle grain boundaries (LAGBs) distributions. It can be observed that
AA7075 GB map is almost free of LAGBs, while AA5083 GB map displayed high density

7
of LAGBs which indicates that the material is in high deformed state. Furthermore, the
as-received AA7075 Al alloy showed an average grain size of 40 µm and the AA5083 Al
alloy revealed an average grain size of 25 µm, as can be observed in Figure4 (a). The
misorentation angle distribution shown in Figure4 (a) indicates the high density of LAGB
in the as-received AA5083 alloy. In terms of crystallographic texture both BMs are
dominated by typical rolling texture that is only of two times random in case of AA7075
alloy,Figure4 (b), and of six times random in case of AA5083 alloy, Figure 4 (c). This
means that the AA5083 alloy displays strong rolling texture, as compared to the AA7075
alloy.

3.3. Grain structure and texture of similar FSWed joints


The grain structures of similar FSWed joints were investigated using EBSD technique to
study the effect of the lowest (50 mm/min) and highest (200 mm/min) traverse speeds on
such structure. The EBSD maps acquired from the nugget (NG) zone of each weld at the
positions indicated by the rectangles on the optical macrographs in Figure 1 using 0.25
µm step size. All the EBSD data presented for the FSWed joints were rotated according
to the methodology described by Ahmed et al.(2008) to align the FSW reference frame
(Welding direction: WD, transverse direction: TD, and normal direction: ND) with the
simple shear reference frame (θ, z, r). Figure 5 (a) and (b) show the IPF coloring OIM
maps relative to (r)and the grain boundary maps with the HAGB >15o in black lines and
LAGBs from 5o to 15o in red lines for similar FSWed AA7075-T6 Al alloys at welding
speeds of50and 200 mm/min, respectively. Figure 6 illustrates the distribution of grain
size and misorentation angle of the data presented in Figure4, as well as, the
crystallographic texture in 100, 101 and 111 pole figures. The IPF maps are mainly
dominated by <111> blue orientation which means the alignment of the <111> directions
with the shear plain normal (r).

From Figure 6 (a), it can be observed that the FSW at the selected welding parameters
has resulted in a significant grain refining in the NG zone with an average grain size of 6
µm at welding speed of 50 mm/min. Furthermore, increasing the welding speed up to 200
mm/min resulted in more refining of the grain structure with an average grain size of 2

8
µm. In fact, increasing the welding speed implies low heat input during the FSW, as well
as, high cooling rate after accomplishing the process, these are the two contributing
factors in refining the grain structure when increasing the FSW speed. With regard to
grain boundaries, the examined areas dominated mainly by HAGBs >15o with a fraction
of 0.86 and around 0.09 of LAGBs between 5o to 15o with virtually similar distribution of
misorentation angle at the two welding speeds, as may be noticed in Figure 5 (a).

In terms of crystallographic texture, it is mainly dominated by a simple shear texture, as


it can be seen from the pole figures shown in Figure 5 (b) and (c).This is the typical type
of texture reported in the probe dominated NG zone of the FSWed Al. Similar results
were observed by Ahmed et at. (2016) and by Fonda and Bingert (2007). It can be
observed that the texture was not affected by the welding speed. Almost identical texture
components with the same intensity were obtained at both welding speeds of 50 and 200
mm/min.

Figure 7(a) and (b) show the IPF coloring OIM maps relative to (r)and the grain boundary
maps with the HAGBs >15oin black and LAGBs from 5 to 15o in red lines for the similar
FSWed AA5083-H111 Al alloys at welding speeds of 50 and 200 mm/min, respectively.
The grain size and misorentation angle distributions of these data, as well as, the texture
in 100, 101 and 111 pole figures were plotted in Figure 8. The IPF map of the low speed
weld was dominated by mixed <001> red, <101> green and <111> blue orientation while
the high speed weld was dominated entirely by <111> blue orientation.

From the OIM maps as shown in Figure 7, a significantly fine grain structure can be
observed in the NG zone. The grain size distribution shown in Figure 9 (a) indicates the
possibility to obtain an average size of 9 µm for the similar AA5083-H111 joints at a
constant tool rotation speed of 300 rpm and a low welding speed of 50 mm/min. The grain

9
size values decreased to an average value of 3 µm for the same joints FSWed at a high
welding speed of 200 mm/min.

This behavior is similar to that observed in the similar FSW AA7075 Al alloys at different
welding speeds. The decrease in the grain size, as a result of increasing the welding
speed may be attributed to the reduced thermal cycle during the high speed FSW. Ma et
al. (2002) showed that the low thermal input during FSW is favorable for the grain size
reduction in the NG zone of FSWed Al alloys. It was reported also by Sato et al. (2002).
Figure 8(a) shows the large number of LAGBs and HAGBs for the high speed weld map,
as an indication for the fine grain structure in the high speed weld relative to the low
speed.

The visual comparison of the EBSD maps grain structure at the same welding speed
illustrates that the AA5083 joint has a coarser grain size than that appeared in AA7075
joint, as listed in Table 2 and shown in Figures 5 and 7. The grain size values are ( 9 and
6µm at 50 mm/min) and ( 3 and 2 at 200 mm/min ) for AA5083 and AA7075 joints,
respectively. It should be noted that, the two material welds were performed under the
same welding conditions, however, they are different in the starting grain structure and
chemical composition. It was reported by Leitao et al. (2012) that the intrinsic BM
characteristics affects the contact conditions between the FSW tool and the material.
Consequently, this will affect the recrystallization process which governs the fine grain
structure formation during the FSW process.

Due to the high stacking fault energy of Al that results in high rates of dynamic recovery,
both continuous dynamic recrystallization (CDRX) and geometric dynamic
recrystallization (GDRX) are reported to be responsible for the new grain structure
formation during hot plastic deformation of Al. CDRX involves the transformation of
LAGBs into HAGBs; and GDRX includes the fragmentation of the initial grains, as
reported by Gourdet and Montheillet (2000). The difference in the grain size of the two
alloys can be attributed either to the difference in the BM characteristics or the
precipitation behavior of the two alloys during FSW. In terms of the first approach, the

10
starting BMs are different in terms of grain size and density of the LAGBs. AA7075 BM is
almost free of LAGBs, as compared to high density in AA5083 BM. Furthermore, AA7075
BM displayed coarser grain size than that of AA5083 BM., According to the rules of
recrystallization which states that the larger the original grain size, the greater the amount
of deformation required to produce an equivalent amount of recrystallization as reported
by Humphreys and Hatherly (2004). Based on this concept and the precipitates difference
between the two alloys, AA5083 alloy was expected to proceed in the recrystallization
process and display coarse grain size which matches with the results obtained in the
present study.

In terms of the precipitation difference approach, AA7075 is mainly precipitation


hardenable alloy with the precipitates expected to dissolve/coarsen in the NG zone during
FSW depending on the thermal cycle experienced, while AA5083 is mainly cold work
hardenable alloy and contains stable Al6Mn particles that can become nucleation sites for
recrystallization during FSW. In terms of AA7075-T651 precipitation behavior during FSW
Su et al (2003) investigated the precipitates distribution in the NG zone of FSWed
AA7050-T651 and reported that the strengthening precipitates’ (Mg (Zn, Cu, Al) 2) and
 (MgZn2 and /or Mg3Zn3Al2) have gone into solution during FSW and re-precipitated
during the weld thermal cycle, with the precipitate redistribution strongly dependent on
the dislocation structure. Also they reported that in the grains exhibiting a recovered
structure, precipitation occurred on dislocation pile-ups or subgrain boundaries, providing
evidence that dislocations are the preferred sites for nucleation and growth of precipitates
during the on-cooling weld thermal cycle. These precipitation behavior can retard the
grain growth upon the thermal cycle and hence results in finer grain size.

3.4. Grain structure and texture of dissimilar FSWed joints


The grain structure of the dissimilar joint between AA7075 and AA5083 Al alloys at low
and high welding speeds was examined at the interface inside the NG zone using the
same step size of 0.25 µm. Figure 9 (a) and (b) show the IPF coloring maps and grain
boundary maps of the low and high speed dissimilar welds, respectively. The upper part

11
of the maps corresponding to AA7075 alloy while the lower part corresponds to AA5083
alloy.
From Figure 9, it can be observed that when applying the same welding speed, both
welds displayed different grain size where the AA7075 alloy in the upper part showed
smaller grain size than that of AA5083 in the lower part. These findings are in full
agreement with the aforementioned observations related to the FSW of similar materials.
It should be noted also that the effect of welding speed on the grain size is not obvious in
case of dissimilar welding, as may be noticed in Figure 9.This implies that the existence
of the two alloys together in the same joint conceals the significant effect of the welding
speed on the grain size, as my be seen in Table 2. Accordingly, the grain size and the
misorentation angle distributions illustrated in Figure 10 (a) for the two datasets confirm
the approximate matching in the grain structure at the two welding speeds. In terms of
texture, the PFs shown in Figure 10(b) and (c) for the low and high welding speed maps
are dominated by the simple shear texture with almost the same texture intensity of
around 3 times random.

3.5. Hardness test results


Figure 11 illustrates the hardness profiles of the similar and dissimilar FSWed AA7075
and AA5083Al alloys at a constant rotation speed of 300 rpm, and different traverse
welding speeds of 50, 100, 150, and 200 mm/min. For all the investigated joints, vertical
dotted lines were plotted in Figure 11 at the heat affected zone (HAZ) with the lowest
hardness values. In general, the hardness values of the all welded joints at various
welding speeds displayed a wide scattering band as shown in Figure 11.
Figure 11 (a) shows the hardness profile of the similar FSWedAA7075 Al plates which
represents a typical hardness profile of FSWed hardened material. The Vickers hardness
value of AA7075 base alloy was observed to decrease from 180 Hv to 159 Hv at the SZ
at the highest welding speed of 200 mm/min. On the other hand, the minimum hardness
value of 117 Hv was found in the TMAZ for a welding speed of 100 mm/min.

12
From Figure 11 (b), it can be observed that increasing the welding speed results in an
increase in the hardness values of the similar FSWedAA5083 joints. Regarding the
AA5083 BM, an average Vickers hardness value of 80Hv was obtained. The hardness
profile was reached a peak value of 90 Hv at the NG zone at the highest speed of 200
mm/min, while the minimum hardness value of 72 Hv was obtained at the lowest speed
of 50 mm/min. The softening at low welding speeds (50 and 100 mm/min) is attributed to
the high amount of generated heat during FSW, in addition to the relatively long time
required for complete heat dissipation from the NG zone. Furthermore, it can be noticed
that the hardness level of RS is relatively higher than that of the AS. Figure 11 (b) also
shows that the width of the region which reveals different hardness values, is 20 mm. This
region includes: NG zone, thermomechanical affected zone (TMAZ) and HAZ. From
Figure 11(a) and (b), it can be observed that the affected zone in case of FSWed joints
of similar AA7075 alloy is wider (28 mm) than that in the case of the joints of the AA5083
plates. The hardness profile of the dissimilar FSWedAA5083/AA7075 joints is presented
in Figure 11 (c).It can be noticed that the NG and the HAZ revealed higher hardness
values than AA5083 BM where a maximum value of 184 Hv was found at the highest
speed of 200 mm/min. From Figure 11 (c), it can be observed that the hardness profile
revealed three distinguishable zones: the softest corresponds to the advancing side
(AA5083); the hardest matches the retreating side (AA7075); and the weld nugget zone
(AA7075/AA5083) which displayed hardness values amongst the other two zones. In
addition, it may be observed that the width of the NG zone is virtually 8 mm which agrees
with the optical macrographs shown in Figure1.

Mishra and Ma (2005) reported that during the FSW process of Al alloys, high temperature
( 500C) can be generated in the stir zone (SZ).This value can be further increased up
to ( 550C) on FSW of AA5083-O Al alloy. The solidus temperatures of AA7075-T6 and
AA5083 Al alloys are 477°C and 591°C, respectively, as reported by Cole et al. (2014).
They have confirmed that generated heat depends on the tool rotational speed, pressure,
friction conditions and tool radius. On the other hand, heat dissipation depends on the
welding speed, material thickness, and the ambient conditions. The rate of heat
generation and heat dissipation determine the amount of heat causing temperature rise

13
and, consequently, structural changes in the NG zone and the surrounding materials. The
increased temperature in the SZ and the excess heat dissipated into the surrounding
material (TMAZ, HAZ, BM) can lead to softening due to dissolution of the precipitates
resulting in forming a supersaturated solid solution in the SZ and an over aging condition
in the surrounding material (Mishra and Ma 2005). These structural changes are the main
source of the reduced hardness values observed in the similar AA7075 joints and AA7075
side of the dissimilar joints.

3.6. Tensile test results


The average tensile test results of the tested three samples of each condition of the BMs,
AA7075 and AA5083 Al-alloys are presented in Table 3.

Figure 12 shows the stress strain curves of similar and dissimilar FSWed butt joints
welded at different travel speeds of 50, 100, 150 and 200 mm/min at a constant rotational
speed of 300 rpm. Figure 12 (a) represents the stress-strain curves of the similar FSWed
AA7075Al alloy joints. The AA7075 BM displayed an UTS of 577MPa and a fracture strain
of 11.5%. The similar AA7075 joints were observed to fracture at UTS lower than that of
the BM. The value of strain at fracture ranged between 0.75 and 1.5 %.Figure 12 (b)
shows the stress-strain curves of the FSWed AA5083 Al alloy joints. The AA5083 BM
showed an average UTS of 309MPa and high fracture strain of 20 %.The FSWed similar
AA5083 joints revealed a lower UTS than that of the BM and low fracture strain ranged
between2 and 3 %.Figure 12 (c) illustrates the stress-strain curves of the FSWed
dissimilar AA7075/AA5083 Al alloys. The dissimilar joints exhibited an average UTS lower
than that given by the both BMs with a fracture strain ranged between 3 and 5.6 %.

Several studies investigated the tensile properties of FSWed joints. Figure 13 shows a
comparison between the relative joint UTS of the Al alloys under the current study and
other Al alloys from the literature as a function of the welding speed. It should be
mentioned here, that the strength of the AA7075/AA5083 dissimilar joints were related to

14
the base alloy of lower strength; AA5083 BM. The joint strength of FSWed AA6082-T6
was reported to range between 216 and 360 MPa when processed under various welding
conditions as reported by Silva et al. (2015). This implies that the joint strength of AA6082-
T6 alloy ranged between 45 and 76% of the AA6082 BM strength, as seen in Figure 13.
The dissimilar FSWed joints of AA6061/AA5086, as reported by Ilangovan et al. (2015),
revealed up to 65% of the strength of the AA5086 base alloy. The joint strength of FSWed
AA2024 ranged between 76.7% and 94.6% of the AA2024 base alloy strength as reported
by Aydin et al. (2009).The joint strength of FSWed dissimilar AA6061-T6/AA7075-
T6alloys revealed 70% of the strength of the AA6061-T6 base alloy(Cole et al., 2014).

The tensile properties obtained in the present study are in an agreement with the results
reported in the literature, as previously discussed and presented in Figure 13. The UTS
of the FSWed dissimilar joints of AA7075 / AA5083 was observed to range between 77
and 87% of the strength of the base alloy AA5083 under various welding conditions.

3.7. Fracture surface analysis


Fracture surface analysis of tensile tested samples was carried out for examining the
nature of the fracture of the FSWed materials under study. Figure 14(a) and (b) show SEM
images of the fracture surface of the similar FSWed joints of AA7075 alloy at welding
speeds of 50 and 200 mm/min, respectively. These images reveal the brittle fracture
mode of AA7075 similar joints. The circled area in Figure 14(a) illustrates the brittle
separation at the poor joined regions. The poor joining may be attributed to the presence
of oxides-rich regions, as reported by Leitao et al. (2014).This poor local joining resulted
in a lower tensile strength, as compared to the strength of the AA7075 BM, as may be
seen in Figure 12 (a). Furthermore, decohesion and/or fracture of the hard precipitates
which indicated by arrows in Figure 14(b) are another feature of the brittle fracture of the
similar FSWed joints of AA7075 alloy. This led to poor ductility joint in terms of low strain
values ranged from 0.75 to 1.5 % at welding speeds of 50 and 200 mm/min, respectively
during the tension test, Figure 12(a).

15
Figure 14(c) and (d) show SEM images of the fracture surface of the similar FSWed joints
of AA5083 alloy at welding speeds of 50 and 200 mm/min, respectively. It can be
observed that the features of the fracture in this case indicates a mixed ductile-brittle
fracture mode, as confirmed by the observed dimple structure and the presence of some
smooth facets referred by arrows in Figure14(c).The refined dimple sizes (7-25 µm) may
be attributed to the high deformation during FSW process which led to the strengthening
of the NG zone in relative to the AA5083 BM. The joint strength of the similar FSWed
AA5083 material was observed lower than that of the AA5083 BM, as shown in Figure 12
(b) which may be attributed to the early fracture occurring at the interface between the
stirred zone and the base alloy.

Figure 15 shows SEM images of the fracture surface of the dissimilar


FSWedAA7075/AA5083 joints at the welding speeds of 50 and 200 mm/min; respectively.
Mixed fracture modes were observed in terms of dimple structure and grain boundary
cleavage in both figures.

3. Conclusions
In the present study FSW was used for developing butt joints between similar and
dissimilar AA7075-T6 and AA5083-H111Al alloys at a constant tool rotation speed of 300
rpm and various traverse welding speeds of 50, 100, 150, and 200mm/min. From the
microstructural, hardness and tensile investigations the following conclusions can be
drawn:-
1- The combination of FSW parameters used in this study have resulted in completely
defect-free joints within the plunge distance for the similar and dissimilar FSWed
AA7075-T651 and AA5083-H111 Al alloys.
2- The significant differences between the two investigated alloys in terms of grain
structure and type of precipitates have strong impact on the recrystallization

16
process that responsible for the formation of the final grain structure inside the NG
zones.
3- A significant grain refining was occurred in the NG zones of both similar and
dissimilar FSWed AA7075 and AA5083 with average grain size ranged between 9
µm and 2 µm depending on the FSW parameters and the alloy type.
4- Increasing the welding speed from 50mm/min to 200mm/min has resulted in a
great reduction in the average grain size for the similar joints from 6 µm to 2 µm in
case of AA7075 and from 9 µm to 3 µm in case of AA5083.
5- The effect of the welding speed almost was not significant in case of the dissimilar
joints with average grain size of 4 µm at both 50mm/min and 200mm/min welding
speeds.
6- The crystallographic texture in the NG zone is dominated by simple shear texture
in the similar and dissimilar joints and without obvious effect for the welding speed.
7- The hardness profile of the similar joints showed the typical behavior for the age
hardened Al alloys in case of AA7075 with hardness reduction in the NG zone and
also the typical behavior of the worked hardened Al alloys in case of AA5083 with
some increase in the hardness values in the NG zone.
8- The hardness profile of the dissimilar joints showed smooth transition in the
hardness between the two different hardness alloys.
9- The dissimilar joints exhibited ultimate tensile strength ranged between 245 and
267 MPa with joint efficiency ranged between 77 and 87% relative to the strength
of AA5083BM and fracture strain ranged between 3 and 5.6 %.
10- The fracture surface of FSWed joints of AA7075/AA5083 alloys displayed a mixed
fracture modes of ductile and brittle fractographic features such as dimples, grain
facets decohesion, and grain boundary cleavage.

Acknowledgements

17
The authors acknowledge the financial support rendered by the Science and Technology
Development fund (STDF), Egyptian State Ministry of Higher Education and Scientific
Research (Project Nos. 3926 and5304). The authors are also thankful to Eng. Hagar Amin
and Eng. Rana Gamal at SSMMR-CSE for their technical assistance through this work.

References

Ahmed, M.M.Z., Wynne, B.P., Rainforth, W.M., Threadgill, P.L., (2008) Quantifying
crystallographic texture in the probe-dominated region of thick-section friction-stir-
welded aluminium. Scripta Materialia 59, 507-510. Doi:
10.1016/j.scriptamat.2008.04.047.

Ahmed, M.M.Z., Wynne, B.P., Seleman M.M., Rainforth, W.M., 2016. A comparison of
crystallographic texture and grain structure development in aluminum generated by
friction stir welding and high strain torsion. Mater Design 103, 259-267. Doi:
10.1016/j.matdes.2016.04.056.

Aydın, H., Bayram, A., Uğuz, A., Akay, K. S., 2009. Tensile properties of friction stir
welded joints of 2024 aluminum alloys in different heat-treated-state. Materials and
Design 30, 2211–2221. Doi: 10.1016/j.matdes.2008.08.034.

Dilip, J., Koilraj, M., Sundareswaran, V., Janaki Ram, G.D., Koteswara, S.R., 2010.
Microstructural characterization of dissimilar friction stir welds between AA2219 and
AA5083. Transactions of The Indian Institute of Metals 63, 757-764. Doi:
10.1007/s12666-010-0116-8.

Fonda, R.W., Bingert, J.F., 2007. Texture variations in an aluminum friction stir weld.
ScriptaMaterialia 57, 1052-1055. Doi: 10.1016/j.scriptamat.2007.06.068.

Gan, W., Okamoto, K., Hirano, S., Chung, K., Kim, C., Wagoner, R.H., 2008. Properties
of Friction-Stir Welded Aluminum Alloys 6111 and 5083. Journal of Engineering
Materials and Technology 130, 1-15. doi: 10.1115/1.2931143.

18
Gourdet, S., Montheillet, F., 2000.An experimental study of the recrystallization
mechanism during hot deformation of aluminium.Materials Science and Engineering
283 A, 274-288.doi: 10.1016/S0921-5093(00)00733-4.

Gungor, B., Kaluc, E., Taban, E., Sik, A., 2014. Mechanical, fatigue and microstructural
properties of friction stir welded 5083-H111 and 6082-T651 aluminum alloys.
Materials and Design 56, 84-90. Doi: 10.1016/j.matdes.2013.10.090.

Humphreys, F.J. and Hatherly, M., 2004. Recrystallization and related annealing
phenomena, 2nd Edition, Elsevier Ltd.

Ilangovan, M., RajendraBoopathy, S., Balasubramanian, V., 2015. Effect of tool pin profile
on microstructure and tensile properties of friction stir welded dissimilar AA 6061/AA
5086 aluminium alloy joints, Defence Technology 11, 174-184, doi:
10.1016/j.dt.2015.01.004

Kasman, Ş., Yenier, Z., 2014.Analyzing dissimilar friction stir welding of AA5754/AA7075.
The International Journal of Advanced Manufacturing Technology 70, 145-156. Doi:
10.1007/s00170-013-5256-7.

Khodir, S.A., Shibayanagi, T., 2008. Friction stir welding of dissimilar AA2024 and
AA7075 aluminum alloys. Materials Science and Engineering B 148, 82-87. Doi:
10.1016/j.mseb.2007.09.024.

Leitao, C., Louro, R., Rodrigues, D.M., 2012. Analysis of high temperature plastic
behaviour and its relation with weldability in friction stir welding for aluminium alloys
AA5083-H111 and AA6082-T6. Materials and Design 37, 402-409.
doi:10.1016/j.matdes.2012.01.031.

Ma, Z.Y., Mishra, R.S., Mahoney, M.W., 2002. Superplastic deformation behaviour of
friction stir processed 7075Al alloy. Acta Materialia 50, 4419-4430.Doi:
10.1016/S1359-6454(02)00278-1

Mishra, R.S., Ma, Z.Y., 2005. Friction stir welding and processing, Mater. Sci. Eng. R 50,
1–78. doi: 10.1016/j.mser.2005.07.001.
19
Palanivel, R., Mathews, P.K., Murugan, N., Dinaharan, I., 2012. Effect of tool rotational
speed and pin profile on microstructure and tensile strength of dissimilar friction stir
welded AA5083-H111 and AA6351-T6 aluminum alloys. Materials and Design 40,
7-16. Doi: 10.1016/j.matdes.2012.03.027

Rodriguez, R.I., Jordon, J.B., Allison, P.G., Rushing, T., Garcia, L., 2015. Microstructure
and mechanical properties of dissimilar friction stir welding of 6061-to-7050
aluminum alloys. Materials and Design 83, 60-65. Doi:
10.1016/j.matdes.2015.05.074.

Saravanan, V., Banerjee, N., Amuthakkannan, R., Rajakumar, S., 2015. Microstructural
evolution and mechanical properties of friction stir welded dissimilar AA2014-T6 and
AA7075-T6 aluminum alloy joints. Metallography, Microstructure, and Analysis 4 (3),
178-187. Doi: 10.1007/s11664-016-4733-9

Sato, Y., Urata, M., Kokawa, H., 2002. Parameters controlling microstructure and
hardness during friction-stir welding of precipitation-hardenable aluminum alloy
6063. Metallurgical and Materials Transactions 33A, 625-635. Doi: 10.1007/s11661-
002-0124-3

Shojaeefard, M.H., Behnagh, R.A., Akbari, M., Givi, M.K.B., Farhani, F., 2013.Modelling
and Pareto optimization of mechanical properties of friction stir welded
AA7075/AA5083 butt joints using neural network and particle swarm algorithm.
Materials and Design 44, 190-198. Doi: 10.1016/j.matdes.2012.07.025.

Silva, Ana C. F, Braga, Daniel F. O, de Figueiredo, M A. Moreira, PM. G. P, 2015. Ultimate


tensile strength optimization of different FSW aluminium alloy joints. The
International Journal of Advanced Manufacturing Technology 79 (5), 805-814. Doi:
10.1007/s00170-015-6871-2

20
Sivachidambaram, S., Rajamurugan, G., Amirtharaj, D., 2015. Optimizing the parameters
for friction stir welding of dissimilar aluminium alloys AA 5383/AA 7075. ARPN
Journal of Engineering and Applied Sciences 10 (12), 5434-5437. SSN 1819-6608.

Su, J.-Q., Nelson, T.W., Mishra,R., Mahoney, M., (2003) “Microstructural investigation of
friction stir welded 7050- T651 aluminium” Acta Materialia 51, P. 713–729

21
Figure 1 Optical macrographs of the transverse cross sections FSWed joints: a) similar
AA7075-T6Al alloys; b) similar AA5083-H111Al alloys; and c) dissimilar AA7075-
AA5083Al alloys (AS: AA5083-H111, RS: AA7075-T6). Rectangles show the locations
for EBSD data acquired and the dashed lines show the locations of hardness
measurements.

22
Figure 2, optical microstructure of FSWed AA7075 at 200mm/min a) Base Material,
b) Diffusive interface between NG and TMAZ at the RS, c) The NG grain structure and
d) Sharp interface between NG and TMAZ at the AS.

23
Figure 3, Base materials inverse pole figure (IPF) coloring OIM maps relative to rolling
direction and the grain boundary maps with high angle grain boundaries (HAGBs)>15 in
black lines and low angle grain boundaries (LAGBs)b <15 in red lines acquired using 1µm
step size. a) AA7075 and b) AA5083.

24
Figure 4,(a) Grain size distribution and misorentation angle distribution of BMs AA7075
and AA5083. (b) and (c) are the calculated 100, 101, and 111 pole figures from EBSD
data of AA7075 and AA5083 BMs, respectively.

25
Figure 5 IPF coloring OIM maps relative to r (ND) and the grain boundary maps, with
HAGBs>15 in black and LAGBs <15 in red lines, acquired using 0.25µm step size of
FSWed AA7075-T6 at welding speeds of (a) 50mm/min and (b) 200 mm/min.

26
Figure 6,(a) Grain size and misorentation angle distributions of similar FSWed AA7075
Al alloys at 50mm/min and 200mm/min. (b) and (c) the calculated 100, 101, and 111
pole figures from EBSD data of the FSWed AA7075 at 50mm/min and 200mm/min,
respectively.

27
Figure 7, IPF coloring OIM maps relative r and the grain boundary maps, with HAGBs>15
in black and LAGBs <15 in red lines, acquired using 0.25µm step size of the similar
FSWed AA5083-131 Al alloys at different welding speeds of (a) 50mm/min and (b) 200
mm/min.

28
Figure 8,(a) Grain size and misorentation angle distributions of the similar FSWed
AA5083 Al alloy at 50mm/min and 200mm/min. (b) and (c) the calculated 100, 101, and
111 pole figures from EBSD data of the FSWed AA5083 at 50mm/min and 200mm/min,
respectively.

29
Figure 9. IPF coloring OIM maps relative r and the grain boundary maps, with HAGB
>15 in black and LAGBs <15 in red lines, acquired using 0.25µm step size of dissimilar
FSWed AA7075 and AA5083Al alloys at different welding speeds of (a) 50mm/min and
(b) 200 mm/min.

30
Figure 10,(a) Grain size and misorentation angle distributions of dissimilar FSWed
AA7075 and AA5083 Al alloys at 50mm/min and 200mm/min. (b) and (c) the calculated
100, 101, and 111 pole figures from EBSD data of dissimilar FSWed AA7075 and
AA5083 alloys at 50mm/min and 200mm/min, respectively.

31
Figure 11, Hardness profiles of similar and dissimilar FSWed joints. Dashed lines show
the NG zone width at the level of hardness measurement as indicated by the horizontal
lines on the optical macrographs in Figure 1.

32
600
(a)
Similar AA7075 joints
400 AA7075 BM
Welding speed 50
100
(mm / min) 150
200 200

0
450
(b)
Similar AA5083 joints
Stress, MPa

300
AA 5083 BM
Welding speed 50
150 100
(mm / min) 150
200

0
600
(c)
Dissimilar joints
400 AA7075/AA5083
AA7075 BM
50
100
200 Welding speed 150
(mm / min) 200
AA5083 BM

0
0 5 10 15 20

Strain, %

Figure 12, The stress strain curves of similar and dissimilar FSWed butt joints welded at
different welding speeds of 50, 100, 150 and 200 mm/min at a constant rotational speed
of 300 rpm.

33
1.5 5083 joints
7075 joints
Joint strength / Base alloy strength

7075/5083 joints
1.3
AA6082-T6, 1000 rpm, Silva et al. (2015)
AA2024, 2140 rpm, Aydin et al. (2009)
1.1 AA 6061 / AA 5086, 1100 rpm, Ilangovan et al. (2015)
6061-T6 / 7075-T6, 700 rpm, Cole et al. (2014)

0.9

0.7

0.5

0.3
0 50 100 150 200 250

Welding speed, mm/min

Figure 13, A comparison between the relative joint UTS of the Al alloys under study and
other Al alloys from the literature, as a function of the welding speed.

34
(a) (b)

(c) (d)

Figure14, SEM images of the fracture surfaces of the similar FSWed joints of AA7075
alloy at welding speeds of (a) 50 mm/min and (b) 200 mm/min and AA5083 alloy at
welding speeds of (c) 50 mm/min and (d) 200 mm/min.

35
(a) (b)

(c) (d)

Figure 15, SEM images of the fracture surface of the dissimilar FSWed joints of
AA7075/AA5083 alloys (a) at a welding speed of 50 mm/min; (b) high-magnification of
the marked area in (a);(c) at a welding speed of 200 mm/min; and (d) high- magnification
of the marked area in (c).

36
Table 1: Chemical composition (in wt.%) of the studied alloys.

Alloy Si Fe Cu Mn Mg Zn Ti Cr Al

AA5083-H111 0.04 0.15 0.02 0.56 4.75 0.04 0.05 0.05 Bal.

AA7075-T6 0.07 0.21 1.94 0.05 2.66 5.97 0.01 0.21 Bal.

37
Table 2. Average grain size of the BM and the NG zone of the different joints
Alloy / Joint condition Welding speed, Average grain size,
mm/min µm
AA7075-BM -- 40
AA5083-BM -- 25
Similar AA7075 50 6
200 2
Similar AA5083 50 9
200 3
Dissimilar AA7075-AA5083 50 3.6
200 3.7

38
Table 3. Tensile properties of the BMsAA7075 and AA5083 Al alloys
Yield strength, Ultimate tensile strength Fracture strain,
Alloy
(YS), MPa (UTS), MPa (Ef), %
AA7075 496 577 11.5
AA5083 156 309 20

39

Potrebbero piacerti anche