You are on page 1of 31

Accepted Manuscript

Title: Characteristics of Aluminum-to-Steel Joint Made by


Friction Stir Welding: A Review
Author: Sadiq Aziz Hussein Hadzley Abd Salam
PII:
DOI:
Reference:

S2352-4928(15)30033-7
http://dx.doi.org/doi:10.1016/j.mtcomm.2015.09.004
MTCOMM 67

To appear in:
Received date:
Accepted date:

20-9-2015
23-9-2015

Please cite this article as: Sadiq Aziz Hussein, Abd Salam Hadzley, Characteristics of
Aluminum-to-Steel Joint Made by Friction Stir Welding: A Review, Materials Today
Communications http://dx.doi.org/10.1016/j.mtcomm.2015.09.004
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.

Characteristics of Aluminum-to-Steel Joint Madeby Friction Stir Welding: A Review


Sadiq Aziz Hussein, Abd Salam Md Tahir, MdHadzley B.A. Bakar
Faculty of Mechanical Engineering
University Teknikal Malaysia Melaka (UTeM), Malacca, Malaysia
salam@utem.edu.my, sadiq.utem@gmail.com
Tel: +6 016 6506784

Abstract
Friction stir welding exploits its solid state process behaviour to join aluminum to steel, which differs in thermal and
mechanical properties, and where combination of these metallic alloys by fusion welding prompts a deleterious
reaction as a result of the melting and resolidification phases. This review investigates the distinction and
characterization of aluminum to steel joints made by this welding method. An attempt to address sub-techniques into
three categories i.e., diffusion, plunging, and annealing, is made. Steel fragments spattered at the weld zone, weld
defects, sharp difference in grain size distribution, and the formed phases of the intermetallic layer and its thickness
are discussed, these factors and the process welding parameters are significantly influenced the joints strength of
this welding method.
Keywords: Friction stir welding, FSW, Aluminum to steel joint, Joint efficiency, Intermetallic layer.
1. Introduction
The rapid development of transport applications suggests adaptive and hybrid structures. These new sophisticated
structural designs can achieve low-weight-products when lightweight materials, aluminum, replace some steel parts
(Figure 1). Chemical plants and cryogenic applications also require these dissimilar materials [1]. For example, in
the mass production of vehicles by Honda Motor Co. Ltd.a hybrid structure of aluminum and steel was produced in
the sub-frame of its new Accord product (as shown in Figure 1-a). The components weight decrease of 25% should
contribute to lower fuel consumption[2]. Therefore, employment of lightweight materials in steel structures, along
with a new potential joining method, is of great interest.
The difficulty of joining aluminum to steel (Al-to-steel) using fusion welding methodsis related to the advanced
differences in melting temperature, thermal properties, and cooling rate after welding of both metallic alloys.
Therefore, the joint strength affected by dissolution of precipitates, shrinkage, and different microstructure of the
diverse materials at the weld zone (WZ)[3]. Furthermore, fusion welding methods are not recommended to weld
thick plates of this joint. Thick,intermetallic compound (IMC),layer is formed caused by high associated heat input
during the fusion process that resulted ina weak joint[4].Dehghani et al. [5] stated that using fusion welding methods
are impossible, or at least, difficult to perform with the low iron solubility tendency inaluminum matrix. This
phenomenon leads to a brittle intermetallic layer formation at the interface of the Al-to-steel joint. Haghshenas et al.
[6] emphasized that the selection of an effective welding method to combine the properties of aluminum and steel is
difficult. However, as mentioned by Ramachandran et al. [7]joining Al-to-steel still in progress to find a suitable
method and welding processs parameters.
Various joining methods and techniques, such as friction stir knead welding [8],friction welding [9, 10], surface
activated bonding [11], abrasion circle friction spot welding [12], cold metal transfer [13], laser penetration welding
[14], and barrel nitriding process[15],have been used to join Al-to-steel. Furthermore, the development of
conventional mechanical joining methods of Al-to-steel, such as lock seam, is on-going (as shown in Figure 1-b).
However, many limitations and difficulties arise when some of these methods are utilized such as cost, setup, and
the equipment required.
Regardless of recent developments in joining methods, and over the past decade, the green technologyFriction Stir
Welding (FSW) has become a potential industrial joining method that has been efficiently used for similar and
dissimilar materials combination.Continuous and spotFSW is more efficient in cost and time, effectively weld
various plate thicknesseswith different joint configurations, and the process generates lower heat input compare to
fusion welding methods which produce thinner IMC at the Al-to-steel interface[4, 16-19]. Moreover,its processis
1

free of high pressure or vacuumchamber needs.Therefore, itcan be easily done inside the assembly plant. Some of
these advantages are not provided by the other solid state welding methods such as diffusion, explosion, magnetic
pulse, ultrasonic,rolling, and forging.That is, sound weld of Al-to-steel is obtained by FSW despite the differences in
thermal properties, mechanical properties, and materials flow of these metallic alloys.Honda make a full useful of
this joint, they developed a new robotic FSW machine for this purpose. This new technology is successfully use in
the continuous welding of the mass production of their vehicle [2].In the present study, a trial has been conducted to
introduce a systematic review on FSW of Al-to-steel. The microstructure evolution, IMC formation, and resulting
strengths of this joint have been addressed.

Aluminum
Outer door panel
(steel Aluminum)

Joined using the


FSW method
Steel
(a)
Inner door
panel (steel)

Conventional

New form

Sealant

(b)

Adhesive agent

Sealant

Pool of adhesive agent

Figure 1.Trends in auto light weight product by Honda; (a)Continuous FSW instead of bolt joints in new sub-frame Al-to-steel
of Honda Accord 2013[2],(b)Newly developed joining method (3D Lock Seam) of Al-to-steel of Honda [20].

2. Research questions
The research questions are explained in Table 1. Furthermore, some sub-questions need for answers which helped
the authors of this study to be guided by these questions during the peer-review process.
Table (1):Research questions and the related sub-questions.
Question
No.
1.

2.

3.

4.

Main question

Sub-questions

Is the Al-to-steel joint made by FSW


efficient and free of defect?

What are the main effective process parameters on the joint efficiency?

What are the hybrid sub-techniques that


used in this type of joining?

Are the hybrid techniques really efficient?

What is the main feature of the resulted


grain sizes distribution?

Is there any effect on the mechanical properties?

Is the IMC (Al xFey) produced in this solidstate welding method as well?

What is the effect of this formation?

What are the effective parameters on the defect formation?

What are the limitations of these techniques?

Where the recrystallization process is mainly occurred?

What is the formation mechanism of this compound?

2 the main affected parameters on this IMC formation process?


What are
5.

Can we categorize the Al-to-steel joining


techniques of this welding method?

What are the main features of these techniques?

3. Materials and methods


The selected welded materials were limited to be aluminum and its alloys and different steel types. An automated
literature search was conducted on friction stir welding, friction stir lap welding (FSLW), friction stir butt welding
(FSBW), and friction stir spot welding (FSSW) of the selected materials. Yet, other FSW configurations such as Tjoint are not available in the literature for Al-to-steel. Recently,however, efficient joints have been made by using
this welding method for different T- geometries (T-lap joint, T-butt joint, and T-butt-lap joint)but for other
materials[18, 19].
The main methodology that followed in this study is shown in Figure2.

Figure 2.Flow diagram of the study and process for


studies inclusion.

4. Joint strength and main affected process parameters


3

Several experimental investigations have been conducted on FSW to join steel and Al Alloys in different
configurations.Figure 3shows a schematic diagram of FSBW and FSLW while the FSSW represents the plunge
stage of the FSLW. Existing studies on FSW joint strength of Al-to-steel exposed different findings, as shown
inTable 2for FSBW, Table 3 for FSLW and Table 4for FSSW. As evident in Tables 2, 3, and 4,numerous
parameters can affect the joint strength for both butt and lap Al-to-steel FSW. The main affected process parameters
are; rotation speed, welding speed (v), tilt angle, plunge depth, and tool design [5, 6, 42, 46].In addition, the material
flow and weld line represent crucial key on the joint strength [52,53]. The material flow of steel and aluminum has
been investigated separately by Morisada et al. [54]. The investigation was implemented during the FSW
process,and using real-time imaging 3D visualized X-ray transmission system. Thus, steel material flow during the
FSW was dimensionally visualizedalong the welding direction (WD) and the transverse direction (TD). It was
uniform in case of low rotational speed (400 rpm) compared to aluminum (Figure 4). Formation of defects in the stir
zone related to the stagnation of steel flow on the advancing side under stir condition. They related the differences in
material flow to the dissimilarity of deformation resistance for both steel and aluminum. However, this limitation
interprets a difficulty of Al-to-steel FSW when optimal steel flow implies low rotational speed [54].

Figure 4.Effect of rotation speed on material flow of steel


and aluminum[54].

It is worthy to mention that to quantify the joint strength of FSBW,both terms;ultimate tensile strength (UTS)[3, 21]
or joint efficiency of weldhas been used. Joint efficiency is defined as the ultimate tensile strength of weldment over
the lower ultimate tensile strength of the parent materials[5, 22].However, for the FSLW, Chen et al. [42] defined
4

the joint strength as F/w ratio which is the ultimate applied shear load (F) over the width of the tested samples (w).
However, in other studies, tensile shear stress was used, determined by dividing the fracture load over the crosssectional area of the lower plate thickness [36, 46, 55]. Furthermore, Haghshenas et al. [6] adopting the failure stress
which resulted from dividing fracture load by the actual weld surface area that measured after fracture.

Table (2): Summary on the used parameters and joint strength of FSBW (Al-to-steel).
Not; The symbols in this table are explained in Figure 3.
Metallic alloys to be jointed
Al alloy

Steel

ta
(mm)

ts
(mm)

Tools material

Ds (mm)

Dp (mm)

N (rpm)

v
(mm/min)

6061

AISI 1018

tool steel H13

25

5.5

914

140

5083

SS400

tool steel SKH57

15

250

25

6013-T4

X5CrNi18-10
stainless steel

800

5083

SS400

tool steel

15

6056

304

7075-T6

mild steel

1050-H16

ASTM A284

Al with
5% Si
(AlSi)

Low carbon steel


(European grade:
DC01)
HC260LA HSS
DP600

6181-T4

()

Efficiency - (UTS)

Reference

38.6% (117 MPa)

[21]

86% - (240 MPa)

[22]

80

Fatigue
strengthefficiency is
70%

[23]

250

25

86% - (240 MPa)

[24]

800

80

[25]

12

500

100

/ - (333 MPa)

[3]

20

6.5

900

100

59% - (81 MPa)

[26]

13

Conical Dr=
6.5

400

300

77% - (115 MPa)

[27]

1.5

1.5

Tungsten
RheniumWRe25

13

1600

480

74% - (2008 MPa)


80% - (2112 MPa)

[28]

73% - (146 MPa)

[29]

/
/

3003-H18

mild steel (St 52)

18

Conical

450

12

5186

mild steel (St52)

18

M3

355

56

90% - (245 MPa)

[5]

6061

stainless steel 304

tungsten carbide

15

3.5

710

30

63% - (254 MPa)

[30]

pure Al

IF steel

tungsten carbide
pin, high speed
steel shoulder

25

600

100

2.5

86% - (123 MPa)

[31]

6082

mild steel

300

/ - (160 MPa)

[32]

6061T6511

TRIP
780/800 steel

1.5

1.4

5052H32

HSLA steel
(IRSM42-93)

AISI 4140 tool


steel
tungsten carbide
with 10% cobalt
content
tungsten carbide

Conical Dr
=5,Dt =2.5

25

12.7

Conical
Dt=2.5

1800

90

85% - (240 MPa)

[33]

20

Conical Dr
=5, Dt = 4

500

45

91% - (188 MPa)

[7]

Table (3): Summary on the welding parameters and joint strength of FSLW of (Al-to-steel ).
Not; In all the cited references Al Alloys were on the top while the steel alloys were on the bottom side.
Not; All the symbols of this table are explained in Figure 3.
Metallic alloys to be jointed

ta (mm)

ts (mm)

Tools material

Ds
(mm)

Dp (mm)

PLD
(mm)

N
(rpm)

v
(mm/min)

()

Results

Reference

Al alloy

Steel

1100H24

Low carbon
steel (SPCC)

1.2

Tool steel
SKD61

10

0.1

2502

252

Peel test; fracture


load=1.024 kN

[34]

5083

SS400

Tool steel JISSKH57

20

0.1

225

43.8

F/w = 559 N/mm (186 MPa)

[35]

AC4C
cast
aluminum

Low carbon
steel Zinc
coated

0.8

Tool steel
SKD61

15

-0.1

1500

100

331.8 MPa

[36]

AC4C
cast
aluminum

Low carbon
steel Zinc
coated

0.8

Tool steel
SKD61

15

-0.2

800

80

F/w = 251 N/mm (314 MPa) without


removing Zinc

[37]

6181-T4

HC340LA

1.5

1.5

UHB Marax
ESR(1.6358)
alloy

13

M5

0.1

1600

360

F/w = 345 N/mm (230 MPa)

[38]

5083

St-12

Tool steel H13

16

Conical Dr
= 4.5 and
Dt = 3

0.1 to
0.2

1125

150

F/w = 304 N/mm (304 MPa)

[39]

0.3
6060-T5

Mild steel

Tool steel H13

20

1400

20

2.5

F/w = 299 N/mm


(115 MPa)
F/w = 435 N/mm
(218 MPa)

[40]

3003

SUS304

15

12

Tungsten
carbide

19

M7

900

300

1.5

76.8
MPa (perpendicular
to the sliced
specimens)

[41]

6060-T5

mild steel

Tool steel H13

25

threaded
(6mm)

1400

20

2.5

F/w = 435 N/mm


(218 MPa)

[42]

6063

Zinc coated
HIF steel
(galvannealed)

Tool steel
SKD61

11

0.1

700

30

F/w = 205 N/mm


(205 MPa)

[43]

5083

St-12

Tool steel
H13

16

Conical Dr=
4.5 and
Dt=3

0.2-0.3

750

230

F/w 328 N/mm (328 MPa) using


annealing after
welding

[44]

1060

stainless steel
(1Cr18Ni9Ti)

Tungsten
carbide

25

10 with
cutting
edges

0.2

950

190

89 Mpa

[45]

DP600 steel zinc


coating
5754

6061

22MnB5 alloy
steel coated by
Al12Si alloy
Zinc coated HIF
steel
(galvannealed)

Fracture stress 50.5


MPa

3
2.1

Tool steel
H13

12

Tool steel
SKD61

11

-0.1

1800

16

[6]
Fracture stress 19.3
MPa

1.5

M3

0.5

1000

50

F/w 249 N/mm


(268 MPa)

[46, 47]

Table (4): Summary on the used welding parameters and the results of FSSW (Al-to-steel ).
Not; The symbols in this table are explained inFigure 3.
Metallic alloys to be
jointed

ta
(mm)

ts
(mm)

Tools
material

Ds
(mm)

Dp
(mm)

PLD
(mm)

N
(rpm)

Plunge
speed
(mm/min)

Dwell
time
(sec)

Description

Results

Reference

Al Alloy

Steel

6061

IF-Steel

1.2

Tungsten
rhenium
alloy
(W25Re)

1.7

3000

The tool used seems to


be conventional one, its
dimensions was not
reported

F 4.5 kN

[48]

6022

Coated
steel

1.1

1.2

SKH51
steel

10

-0.2

3000

Four different coating


materials used to cover
the steel plate. Low
melting temperature of
zinc and its alloy
(<430C) enhanced the
joint strength.

Modified cross
tension
test (M-CTT,
hereafter). Maximum
F 1.6 kN

[49]

6111-T4

DC04 low
carbon

Tungsten
carbide

11
Scrolled
profile

Tapered
Dr= 3
mm

0.1

800

The tool move along an


orbit path with 5 mm
diameter, the optimum
moving speed is 1000
mm/min

F =3.5 kN

[12]

6061-T6

low carbon
steel

10
scrolled
groove

No pin

Shoulder
plunge is
1 mm

3000

A comparison between
different joints had
been made, the Al-tosteel joint exhibited
higher strength
compare to Al/Al joint

Static stress was


approximately 63
MPa while the
fatigue strength was
nearly to be same of
that for Al/Al joint

[50]

6061-T6

Mild steel

12 and
15

The back plate


containing round dent,
the process
accomplished within
two steps, the first step
was a conventional
FSSW, while in the
second step a flat tool
(without pin) used to
eliminate the key hole

F =3.607 kN for
final flat weld and
3.189 kN for the
conventional method

[51]

The formed defects caused by unsuitable process parameters, such as tunnel defect, are highly affecting the joint
strength of FSBW [5]. These defects were the main target to be eliminated by Dehghani et al. [5], using different
parameters. They claimed that a free defect joint with strength efficiency of 90% was obtained. They used criteria of
equation 1 that had been developed by Zhang et al. [56] in their investigation to avoid the tunnel defect.
.

2.

(1)

where V is the volume of the defect, k is a proportional factor, P is the applied pressure (Pa), N is the tool rotational
speed (rpm), v is the welding speed (mm/min), D s and Dpare shoulder and pin diameters respectively (mm), (
seeFigure 3).
Another defect can also affect the tensile strength, as stated by Springer et al.Kirkendall porosity formation on the
Al-to-steel interface declined the joint strength between pure Al and steel while joint of AlSi5 to steel affected by
the thickness of the Al5Fe2 phase which formed in case of both used Al alloys [27]. Watanabe et al.referred to the
deleterious effect of the spattered steel fragment, oxide formation, and voids at the weld zone on the tensile strength
[24]. Using of not optimal tool material such as tungsten- and molybdenum-based alloys is significantly affect the
weld structure because of the incoming tool wear and steel fragment when pin is made from steel alloys [1]. In
different circumstances, the incomplete recovery/recrystallization process and the associated properties of the strong
microstructure gradient at the weld zone can weaken the joint further, as it will discuss later [28].
However, as concluded by Kimapong and Watanabe,FSBW of Al-to-steel could be easily and successfully to
achieve. Adjusting the rotating pin position, distance e in Figure 3, (only + 0.2 mm inserted in the steel plate) is
important when it is well matching with the rotational and welding speeds, 250 rpm and 25 mm/min respectively, to
achieve tensile strength efficiency of 86%. Increasing this e distance (+ 0.6, + 1, and + 2 mm) or decreasing (0 and 0.2 mm) caused drastically decrement in joint strength[22]. Ramachandran et al.examined this distance effect using
inclined weld line, started and ended with different e for each weld pass. They investigate the pin profile effect as
well, tapered pin with an angle of 10 showed the optimum strength compared to 30, 20, and 0. For e = + 2 mm,
the pin taper angle of 10 is fully contacts the steel thickness length, hence fully weld incidence was occurred along
that thickness. The joint efficiency was the maximum (91%) with these two values of e and taper angle [7].
In FSLW, pin plunging of steel plate is very important [35, 45]. Though, the pin is not reaching the plate surface or
penetrates through, joint can be produced [6, 42].The defects at the weld region of top plate (Al Alloy) resulted from
not adequate selected parameters values such as rotational speed, welding speed, and tool tilt angle which can
drastically decrease the shear strength. Plug fracture mode at the stir zone (of Al Alloy) with relatively low shear
strength was occurred [45] as shown in Figure 5 which might be affected by the weld defect of the Al- plate.
Another type of defect may occur at the bottom plate (steel). The hook defect formation at the steel surface adds
another deleterious effect on the joint strength of the FSLW when the pin penetrates the steel plate. This defect
could serve the crack propagation incident under the loading process[40]. The steel surface coating condition shown
in Table 2is also influenced the joint strength of this welding configuration (FSLW) [6, 36, 37, 43],Chen and
Nakata[37]reported that Al/zinc-coated steel joint exhibited higher failure load as compared to that with no-zinccoat steel, zinc coating assisted better diffusion at Al-to-steel interface.
The absence of melting in friction welding methods exploits the superior joint strength in Tables 2, 3, and 4. In
fusion welding methods, differences in melting and resolidification points of both materials, and low iron solubility
tendency in aluminum lead to brittle IMC and defect formations[5]. Su et al.stated that fusion-based welding
brazing methods provide lower joint strengthas compared to solid-state friction welding. Their lap joint made by
alternate-current double-pulse gas metal arc welding of both 1 mm thick AA5052 and galvanized mild steel, the
ultimate tensile shear strength showed 210 MPa. Thick-brittle IMC layer (Al5Fe2 andAl3Fe phases) is formed at the
Al-to-steel interfaces after resolidification incidence [57]. Qin et al.used hybrid method consisted laser and MIG arc
brazing-fusion welding technique to increase the joint efficiency between 1 mm thick 6013-T4 and 2 mm galvanized
steel, the filler metal was AlSi5, the ultimate shear strength was 247.3 MPa, thin and brittle IMC layers were formed
[58]. Laser welding-brazing technique without filler metal was used to lap weld 6 mm thick AA5083-H22 to 2 mm
thick high strength low alloy steel (XF350), the obtained ultimate tensile shear strength was 261 MPa (F/w = 522
N/mm), the brittle IMC was existedin this welds as well [59].Ma et al. used two pass laser welding, even though,the
10

results werefree-defect, the ultimate shear strength was 211 MPa (F/w was 158 N/mm)and the Al-rich compound
formed at the weld interfacebetween dual phase steel DP590 and AA6061-T6 [60].Butt welding of 3 mm thick 5A06
aluminum alloy and AISI 321 austenite stainless steel was made by Song et al.using TIG weldingbrazing
method.The UTS was 125.2 MPa despite the using of different filler metals containing different silicon additions to
prevent the formation of thick IMC [61]. The hybrid fusion techniques were also used for this joint configuration to
increase the joint strength. For example, poor UTS 96.7 MPa was obtained fromMIG welding made by Zhang et al.
to combine 3.5 mm thick AA7005 and stainless steel 321,they noticed an enhancement in the UTS to be 146.7
MPawhen the process was assisted by an auxiliary TIG arc to heat the lower thermal conductivity plate (steel)
[62].For spot welding, when the electrodes design of the resistance spot welding modified by Zhang et al. [63]to
solve the thick IMC formation problem, failure shear load 5.4 kN was obtained for joint of 1.5 thick AA6008-T66 to
1 mm thick H220YD galvanized steel. Perhaps the short time of resistance spot weld process and the function of the
newelectrode which prevented thick IMC formation can express this result.
From the aforementioned literature, it is evident that fusion welding methods give limited strength values of Al-tosteel joint. Therefore, Liu et al.suggested Tricladtransition jointwhich is made by solid state welding (explosion
welding) to avoid fusion welding of Al-to-steel directly. The Tricladconsisted ASTM A516 steel as backer plate,
AA5456 as flyer plate, and AA1100 as interlayer. In this case, it is easier to weld AA6061-T6 and AHSS steel to the
aluminum and steel sides respectively of the Triclad part. They used a hybrid laser- arc welding method to join the
aluminum-aluminum and steel-steel, the UTS was 220 MPa, the joint efficiency was 73% of the AA6061-T6 base
alloy, and the fracture occurred at the heat affected zone of the AA6061-T6 side [4]. This joint is a combination of
three joints, Triclad, aluminum-aluminum, and steel-steel which could increase the cost and time.
However, though Schneider and Radzilowskibelieved that piece by piece substitution of high strength-weight
sophisticated structure in an application design still not optimum [64], FSW can attain robust joint of Al-to-steel (see
Tables 2, 3, and 4). It appears from the above literature concerning the FSW of Al-to-steel that to achieve an optimal
joint strength, the selected parameters should be well matching. Furthermore, the hybrid FSW technique may
enhance the joint strength further as it will be shown in the next paragraph.

Welded
region

Steel

Defects

Fractu
re

Al

Figure. 5.Fracture occurrences at the processing area of the


upper aluminum plate at the process/weld region [45].

5.Hybrid technique consequence


Recently, hybrid techniques have beenemployed in FSWfor several materials and alloys,particularly for steelsteel
joining [65, 66]. These methods are generallyaimed to pre-heat the steel plate materials. Such action would soften
the steel material and enable better flow. When high heat is produced by these techniques, the mechanical properties
of mostlightweight alloys are affectedby a new modification on the microstructure grain size, which is related to the
low melting and annealing temperature. For instance,Luo et al.supported the experimental setup of the FSW by an
11

electric resistance heat source to separately weld (similar welding) magnesium alloy (AZ31B), Al Alloy(AA7075),
and steel. A 220-voltpower supply as input power and 0150 amperesadjustable output electrical current were used
to implement theirhybrid technique. The microstructural grain sizes of the welds were investigated for the three
different joints. They found that the AA7075 grain sizes of the weld zone increased when the applied current
increased. An increase was noticed at the heat-affected zone as the grain sizes of AZ31B were significantly refined
at the stir zone. Theyconcluded that this hybrid technique was more suitable for St/St welding [67]. However, to
increase the joint efficiency of Al-to-steel,some hybrid techniques were used toassist the FSW, such as the assisted
techniques to preheat the steel alloys only during the welding process, which were proposed byMerkleinand
Giera[68] andBang et al. [69].
Merkleinand Giera,presented a tailored hybrid welding method by adaptingthe spot laser assistance that
preheatedthe steel ahead of the FSW tool. This technique can enhance the deformationof the steel plate and
minimize wearof tool. The resulting joint efficiency of AA6016-T4 to DC04 steel FSBW can reach 80% of the
aluminum base alloy. The results showed that welding speed increment to 2000 mm/mincan demonstrate high weld
formability.Furthermore, the researchersstated that nointermetallic compounds have been observed [68].Bang et
al.used gas tungsten arc welding (GTAW) to assist the FSW of AA6061-T6 in stainless steel (STS304) butt
configuration. The resulting joint efficiency increased to 93% along with a significant increase in elongation. They
suggestedthat the optimized results are related to the partial annealing and enhancement of plastic flow of steel at the
leading edge of the rotated tool [69].
However, using hybrid techniques in the FSW process requires extra equipment,which contributes high cost.
Compared with other welding methods, FSWis less costly and easier to setupthan other methods. Therefore,
manipulating the in-process parameters of FSW to perform a robust joint of Al-to-steelismore beneficial.
6. Microstructure evolution
The microstructure of Al-to-steel dissimilar weld is investigatesat the stir zone (SZ), thermo-mechanically affected
zone (TMAZ), heat-affected zone (HAZ) and the base material (BM). The microstructure of FSLW in case of no pin
plunging in steel is same of that well discuss for similar welding of Al Alloys [70]. Hence, the dynamic
recrystallization (DRX) if occurred would be inthe upper plate only(Al Alloy)[6], this process represent friction stir
processing (FSP) [71].Nonetheless, Sharp gradient of aluminum material grain size below the shoulder was reported
by Haghshenas et al. [6], the results indicated a distinct boundary between SZ and TMAZ because of no
recrystallization occurrence at TMAZ. In general, the modified grains size caused by DRX for similar FSW/FSP can
follow the expression[70]:
=

+ ( ) (2)

Where D is the grain diameter, a and b are constitutive constants, and Z is Zener-Hollomon parameter:
= exp

(3)

Where is strain rate, Q is the activation energy, R is the gas constant, and T is temperature in kelvin.
When the pin plunge the steel plate (lower plate) in FSLW the SZ region will be a combination of two different
recrystallization grain size regions (mixed stir zone) [38]. Elrefaey et al. notice these regions, a very fine steel grain
sizes were arranged in a thin layer beneath the rotated pin when the pin plunge the steel plate. They also referred to
the grain size variety which indicated a distinct boundary between SZ and TMAZ for the Al Alloy plate [34], closed
results were exposed by Coelho et al. [38]as well.
More complicated microstructure is formed at the Al-to-steel FSBW interface and especially when steel fragment
spattered at the WZ as shown in Figure 6. Electron backscatter diffraction (EBSD) technique used by Coelho et al.
[28] exposed large differences in grain size distribution and shape at the BM, HAZ, TMAZ, and SZ as shown in
Figure (7).However, at the SZ region which is reported as slightly bigger [5] or smaller [42] than the pin diameter,
the small grain size produced through DRX mechanism [5]. Grain size affected by the stirring process and
temperature distribution, Kundu et al. [31] founded that increasing rotational speed was the main cause of re-modify
12

the grain size as shown in Figure (8-a). It is well known, the heat generation model is mainly affected by the
rotational speed [72]. Besides, usingof low welding speed can increase the heat cycle time duration and thereby
grain sizes were coarsened [34]. Therefore, it can be concluded that the characteristic of DRX is attributed to sever
plastic deformation and differences in thermo-mechanical cycles [28, 34].This is aided by Springer et al.results;
welding of different Al Alloys with different grain size may lead to same grain size at the SZ when same process
parameters have been used, they reported same grain size, 3 m, at the Al-to-steel interface when different of 20 and
120 m grain sizes of AA99.5 and AASi5 base material were weld to steel respectively, a relatively low rotational
(400 rpm) and high welding (300 mm/min) speeds were used[27].

Stainless
steel

HAZ

TMAZ

Advancing side

SZ

TMAZ

HAZ

AA6013

Retreating side

Figure 6.Standard cross-sectional image illustrates the weld zone [23].

At the TMAZ, no DRX occurred as reported by Dehghani et al. [5]. While grain size deformed and elongated due to
the rotation action at the TMAZ, no major change of the grain size at the HAZ compared to the BM [5, 23] see
Figure 9. The Al alloy base material shows elongated grains with a size of 70 m while the Al alloy at the SZ shows
characteristics of fine equiaxed grains 5 m in diameters (i.e. refined by ~ 93%) as shown in Figure 7 [28].
However, Kundu et al. reported the increase in the steel and aluminum grains number at the SZ, TMAZ, and HAZ as
shown in Figure (8-a), the maximum number of grains were at the SZ[31]. In FSSW, Sun et al.observed that Al
Alloys grains refined by ~ 84% at the SZ while the steel grains appeared low tendency to change [51].
However, distinct boundary between SZ, TMAZ, HAZ, and BM affect the mechanical properties of the joint [28,
31]. For example, the microhardness results along the WZ of Kundu et al. [31] investigation were influenced by this
grain sizes variety as shown in Figure (8-b).

Figure 7.Large difference in grain sizes and


distribution at 13
the weld zone [28].

H
ar
dn
es
s
Vi
ck
er
s

A
S
T
M
gr
ai
n
siz
e

FSW zones

Distance (mm)

(a)
(b)
Figure 8.An illustration for grain sizes distribution at the weld zone and microhardness for different rotational speeds, (a)
number of grains at each region of the weld zone, (b) the microhardness results [31].

HAZ
TMAZ

SZ

Figure 9.Distinct gradients of the grains distribution at the weld zone


between SZ, TMAZ, and HAZ [5].

The microstructure of Al-to-steel weld showed also small amount of steel fragments as observed by Uzun et al. [23]
and Watanabe et al. [24] at the SZ of FSBW where the stirring effect highest during welding. The wear of the
pincaused cluster of discrete pores dispersed along the faying line especially at the bottom of the weld as
investigated by Chen and Kovacevic [1], further steel fragments resulted from tool wear were also seen in this
investigation.On the other hand, an oxide film 5 m thickness was appeared at the Al-to-steel interface where the
distancee in Figure 3small [24],increasing this distance eliminated the oxide film while the Fe fragment at the SZ
14

increased due to the further steel rubbing.Furthermore, Energy Dispersive X-ray Spectroscopy (EDS), X-Ray
Diffraction (XRD), SEM (Scanning Electron Microscopy),and Transmission Electron Microscopy (TEM) tests
detected an IMC formation which represents a main complex point in the Al-to-steel FS weld investigations and
analyses.
6.1 Characteristics of intermetallic compound formation
In addition to the fusion welding methods, the intermetallic layers were observed at the aluminum/steel interface in
composites produced by melt infiltration or by solid-state hot-pressing, in Al-clad 430 stainless steel foils after solidstate homogenization heat treatments, and also in brazing of aluminum to stainless steel (see Roulin et al. [73]).
During the FSW process, stir incident and supplied pressure intends to break the oxide layers of aluminum and steel
surfaces. When sufficient heat is existed the atomic bonding between aluminum (Al) and iron (Fe) atoms is
produced [37]. The resulted atomic bonding is an IMC which approximately illustrated in Figure 10.

Aluminum

Intermetallic
compound

Steel

Figure 10.An illustration of IMC formation at the


Al/-to-Steel interface.

Most studies that evaluated the joint strength of IMC agreed that thick IMC (if formed) decreases the joint strength
of FSBW and FSLW of Al-to-steel. Movahedi et al.[39] and Ghosh et al.[30] agreed with this view because the
crack propagation region shifted to the stir zone instead of the IMC zone when the IMC was thin. Bozzi et al. [48]
found that IMC seems necessary to improve the weld strength of the Al-to-steel joint; however, the thick layer of
IMC is detrimental because it significantly deteriorated the bond strength.They believed that the effect of layer
thickness is the significant parameter that affects the weld strength. Chen et al. [42] agreed with this consideration as
well. In contrast, Sun et al. [51]reported that preventing IMC is important to improve the joint strengthof FSSW
while Chen et al. [12]produced successful joint without intermetallic presence for abrasion circle friction spot
welding.
Springer et al. used the annealing processes after welding to produce that reaction because the IMC formations could
not be observed. After annealed, the IMC details became attainable at the Al-to-steel interface. The annealing
process (for temperature 300 C) affected the mechanical properties of Al-weld regions, which resulted in low
tensile strength. For annealing temperature 500 C, particularly when the annealing time increased as well, the
IMC thickness increased and the failure locations shifted to the Al-to-steel interface[27].
In some other studies, IMC could not be observed. This absence can be related to the discontinuity of IMC layer
which may hamper showing of the IMC at the selected investigated sample along the welding line (transversely and
longitudinally) [10, 42, 51]. Furthermore, Girard et al. [26] exposed an important fact that SEM test cannot detect
the thin IMC, they used TEM test to observe very thin layer when the first test did not detect that layer. Different
temperature distribution along that welding line also can play a significant role on this absence.
6.1.1 IMCsforms, formationand growing process, andstrength
According to existing researches, the low solubility of iron in aluminum matrix promotes different constitutions of
IMCs in various forms at the Al-to-steel interface[28].Electron diffraction patterns using XRDtechnique performed
on several precipitates confirmed them to be in these different phases. In the binary phase diagram shown in Figure
11, AlFe3, AlFe, Al3Fe2, Al2Fe, Al5Fe2, and Al3Fe (Al13Fe4) are the most expected IMC forms consistent with the
15

findings of most studies[6, 21, 35, 42]. The atomic aluminum percentage that governs the resulting forms of IMC is
presented in Table 5.

Table (5):Atomic percents of the main


IMCs forms[10, 48,74].
IMCs forms

Al percents (at. %)

AlFe3

25

AlFe

50

Al 3Fe2

63

Al 2Fe

66-67

Al 5Fe2

69.7-73.2

Al 3Fe (Al13 Fe4)

74-76

All the IMC phases are possible to form,as shown in Figure 11; however, a formation mechanism is needed for
clarity. To explain the formation mechanism, Das et al. [43]cited the formation mechanism of IMCs from the
available literatures which has the followingthree stages:
Stage 1The first expected phase occurs because of metal-to-metal interaction. The first phase develops by the most
negative heat on the concentration of the lowest eutectic of the binary system.
Stage 2Phases react with each other to form a new phase with a composition between that of the interacting phases
and the lowest eutectic composition. Al13Fe4has the lowest effective free energy of formation;thus, it is
expected to form at this stage.
Stage 3Al13Fe4 and Fe phases react with each other to form more Fe-rich compound.
It can be assumed that the first stage has the feature of the amorphous phase which investigated by Ogura et al. [41]
while the third stage noticed byHaghshenas et al.when a transformation of Al5Fe2 to AlFe phases was occurred
promoted by the lower welding speed (16 mm/min instead of 45 mm/min) [6], lower welding speed promotes higher
heat cycle[46, 47]. Besides, Das et al.also noticed the formation of the Al13Fe4 followed by formation of AlFe3 when
a longer heat cycle with higher peak temperature was implemented through increasing rotational speed (1000 rpm
instead of 500 rpm) and decreasing welding speed (50 mm/min instead of 100 mm/min) [46, 47]. However, as
claimed by Kusuda, Honda Motor Co., Ltd.has obtained Al-rich phase, Al13Fe4, for their new product of Accord
2013, using continuous FSW of Al-to-steelsub-frame structure[2].
In general, the Al-Fe phase diagram shows two different phases groups:a) Fe-rich intermetallic (IM) phases that
include AlFe and AlFe3, and b) three Al-rich IM phasesAl2Fe, Al5Fe2, and Al13Fe4 [27].As mentioned above SEM,
TEM, EDS,and XRD have been used to characterize IMCs at the interfacial area of Al-to-steel.However, the crystal
morphology of these compounds arrange in various structures. Someof these structuresare shown in Figure
12,whereas some contrast may be emphasized in the literature especially for the microstructure of Al2Fe phase,
seeTable 6.

16

Atomic percent Aluminum

Te
mp
erat
ure
(C
)

Weight percent Aluminum


Figure 11.AlFe binary phase diagram [75].

Table (6):Crystal structure and hardness of the phased formed in the Al-Fe binary phase diagram.
Phases

Crystal structure

Hardness Vickers

[74]

[48]

[75, 74]

[48]

([74])

AlFe

BCC

470

400

(491-667)
(344-368)

AlFe3

DO3

330

Al3Fe2

Cubic

650

Al 2Fe

Triclinic

Rhombohedric

1000

(1058-1070)

Al5Fe2

Orthorhombic

Orthorhombic

1013

1100

(100-1017)

Al 3Fe

Monoclinic

Monoclinic

892

820

(772-1017)

( ) means; the available data are cited by reference[74]

The compound layer thickness is mainly influenced by the generated heat. So, after the formation of the IMCs, the
thickness continuing increases depending on heat time hold. Springer et al. [27] usedanexpressionto estimate the
thickness increasing rate as shown in Equation 4during their annealing effect investigation.
= . (4)
where d is the average thickness data from the annealing experiments for 450, 500, and 600C (m); is the growth
rate (m sec-1); and t is the annealing time (sec). The value is linearly related to process temperature as explained
by Su et al. [76].The IMC thickness resulted fromequation 4was well agreed the actual size in fusion welding as well
[76].
However, a constant acceptable thickness of the IMC layer, if formed, for various Al-to-steel joints remains difficult
to determine.For instance, when the AlSi5 was used instead of the pure Al, an increase in the IMC layer thickness of
more than 1.6 m could lead to site fracture at those faying surface locations. But, when the thickness of the IMC
17

layer increased approximately more than 8.5 m, the fracture locations occur at the faying surfaces between the steel
and Al Alloys, pure Al 99.5%, in spite of kirkendall porosity presence [27]. Thus, an upper allowable thickness limit
of the IMCs in the literature seems difficult to be adopted so far. Table 7summarizes the main IMCs forms and the
resulting thicknesses in some selected studies. It is worthily mentioned that thicknesses values in Table 7represent
the interfacial layer between aluminum and steel plates. Investigation thicknesses of each formed phase (AlxFey) are
difficult when this layer consisted of more than one phase [46] seeFigure 12.

(a)

(b)
Figure 12.TEM micrographs ofIMC tangle at the interface Al /St of weld. (a) Selected area diffraction
patterns of Al 2Fe (rhombohedric phase) and Al5Fe2 (orthorhombic phase), and (b) micrograph of Al3Fe
(monoclinic phase) [48].

Joint of AA5754/22MnB5-steel with Al5Fe2 IMCs formation exhibited lower failure strength compare to
AA5754/DP600 joint with AlFe IMCs [6]. This can be attributed to the Al-rich (Al5Fe2) IMCs which is hard and
brittle while Fe-rich (AlFe) phase showshigher ductility and strength [6,75,77].The hardness values of different
IMCs phases are shown in Table 6. That is, formations of brittle IMCs, Al-rich, are detrimental to the mechanical
properties of dissimilar Al-to-steel joint [38, 75]. Ozaki and Kutsuna, explained in Figure 13 the mechanical
properties of the IMCs forms; it is evident that Fe-rich phases accommodate relatively higher compression strength
and ductility [75]. An increase in layer thickness weakens the joint strength caused by the crack initiation through
the tangles of IMC [48].

18

Table (7):Summery on the IMC forms and thicknesses in the


literatures.
IMC phase
Al 13Fe4, Al5Fe2

Layer thickness
(m)
5

Reference

0.2 - 43

[35]

7.7 - 58.1

[36]

0.2

[38]

AlFe, Al3Fe
Al 5Fe2, Al 13Fe4
Al 5Fe2, Al 3Fe

[21]

1.3 (when
AA5083 and
AA1100 welded to
steel)
5 - 48

[48]

0 - 23

[27]

[39]

2.5 - 25

[31]

[2]

Al 2Fe, Al5Fe2

1-2.5

[42]

Al 2Fe, Al 5Fe2,
Al 13Fe4
Al 5Fe2, Al 3Fe

0.7 - 1.8

[30]

0.33 - 9.11

[44]

[33]

Al 3Fe
Al 5Fe2, Al 3Fe, Al2Fe
Al 5Fe2
Al 13Fe4, Al5Fe2
Al 3Fe
Al 13Fe4

AlFe, AlFe3
AlFe, Al5Fe2
Al 13Fe4, AlFe3
Al 5Fe2, Al 3Fe, AlFe

[3]

2-3

[6]

4.1-6.5

[46, 47]

<< 0.1 - 26

[7]

800
Co
mpr 600
essi
on
stre 400
ss
(MP
a)
200

0
Strain (-)
Figure 13. Stress-strain curves, compressive test of AlFe IMCs [75].

6.1.2Factors affected IMCs formation process


In some previous studies, ambiguous claims concerning the main influencing parameters on the IMC formations
required clarification.

19

Many process parameters can affect the IMC formation. For instance, Chen et al.,used arange of welding speed for
their experiments (60-2000 mm/min), the rotational speed was relatively slow (800 rpm), theydid not notice the IMC
layer,less than one second to pass an orbital welding path with 5 mm diameter as a time cycle was successfully used
to get a robust joint[12]. Tanaka et al.,did not clearly observe the IMC for AA7075-to-mild steel. The parameters
used were low rotational speed 500 rpm and relatively high welding speed 100 mm/min[3]. However, Bozzi et al.,
reported that the IMC layer formation depends mainly on the rotational speed. When the rotational speed increased
by 1000 rpm (from 2000 rpm to 3000 rpm), the IMC layer thickness increased approximately by 3 m. However, an
extra increase by 500 rpm (from 3000 rpm to 3500 rpm) caused a high increment (from 8 m to 42 m) in layer
thickness[48].Kundu et al.,investigated as shown in Figure 14 the effect of rotational speed on the intermetallic layer
thickness. However, direct relations between process parameters and IMC layer thickness were obtained in Figure
15, they explained the relationship between tool rotational speeds and IMC layer thicknessesfor a joint made of
commercially pure aluminum and interstitial free steel by FSBW [31].
Liu et al.examined the e offset distance in Figure 3 together with the effect of rotational speed and welding speed
(v). When the rotational speed was 1200 rpm and e = + 1.03 mm, the IMC thickness was linearly dependence of
1/v, the formed IMC phase was AlFe. Other testeddistance was e = + 1.63 mm, for different rotational speeds,
1200 and 1800 rpm linear dependence of 1/v0.5 was observed and the formed phase was AlFe3[33]. Girard et al.,
inserted the full pin surface in the Al Alloy plate (e = 0 mm). The SEM showed that an IMC layer of 0.5 3 m
thickness was formed when 20 mm/min was used as a welding speed. But when they used v = 100 mm/min, the
TEM test detected thin IMC layer which was not continuous and 50-100 nm in thickness only [26].The pin length is
another very important process parameter in FSLW. Enough pin length to penetrate the bottom plate (steel) by PLD
distance that shown in Figure 3 may be contributed in this IMC formation. Kimapong and Watanabe [35] observed
the effect of PLD distance (Figure 3) on the IMC formation phase and thickness. For PLD = + 0.1 mm, pin plungein steel, two compounds Al3Fe and AlFe were formed, when PLD increased to + 0.2, + 0.3, and + 0.4 mm only
Al3Fe phase was formed. The thickness increased by more than 3 m when + 0.4 mm used instead of + 0.1 mm. By
contrast, and while Lee et al. [78]stated that no variation was occurred in the formed IMC layer thickness when
different pin plunging depths were used,Sun et al.[51]noticed that no IMC could be formed even if different
plunging depths were used. Nonetheless,the used pin length may be shorter than the upper plate thickness, i.e. no pin
plunging into steel (PLD 0) is occurred, but the IMC is formed at the interface as well [6,36, 37].
Dehghani et al. [5, 29]revealed that the tool tilt angleand tool sizeare effective parameters on IMC evolution; these
parameters affect the generated peak temperature in FSBW. For example, when the used tilt angle was 5instead of
3,the heat input increased, and so,the IMC thickness changed from 0.9 to 3.9 m[5]. Ramachandran et
al.investigated the significant effect of the pin taper angle on the formed phase of IMC. This parameter has a
significant effect on the material flow during the welding process. For 0, 10, 20, and 30 as pin taper angles, the
formed phases were Al3Fe, Al3Fe, Al5Fe2, and AlFe respectively [7].

(a)

(b)

Figure 14. Effect of rotational speed on intermetallic layer thickness:(a)


600 rpm, (b) 1200 rpm [31].

20

I
M
C
thi
ck
ne
ss
(
m

Tool rotational speed (rpm)


Figure 15.Effect of rotational speed on IMC thicknessin
FSW[31].

The generated forces and energies during the welding process as a result of the solid-state behavior of FSW had an
essential role in studies made by Liu et al. [33] and Das et al. [43] respectively. However, the force parameter is
influenced by other process parameters, particularly the rotational and welding speeds, and it is difficult to control
along the welding line especially with different existence stages in FSW process (plunge, dwell time, tool moving,
and tool withdrawal stages) [42, 79].
Thermal cycle gives temperature history at each point along the welding line as shown in Figure 16, it was the
indicator factor on the IMC formation and growing in Chen et al. [42] and Das et al. [46, 47] studies. Lee et al. [78]
noticed the IMC dependence on the holding time along with specified temperature range.

Te
mp
era
tur
e
(
C)

Time (sec)
Figure 16.Heat cycle resulted from used parameters;
1400 rpm and 20 mm/min as rotational and welding
speedsrespectively [42].

In a similar way, Tanaka et al. [3] andDehghani et al. [5]used an approximate heat input models. The heat generation
model of the FSW process developed by Frigaard et al. [72] was adopted by Dehghani et al. [5] when it was
modified to consider the effect of welding speed. So, the model is divided by the welding speed and the resulting
term called heat input factor. Because of, the force and the friction coefficient factors couldnt be real measure
21

during that process, they were not considered in their studies [3, 5]. Furthermore, the activation energies that needed
to obtain the IMC represent another important indicator factor, for example, the activation energies to attain Al5Fe2,
AlFe, and AlFe3 layers were 141 kJ mol-1,180 kJ mol-1, and 260 kJ mol-1 respectively during a diffusion treatment
process (Fe-rich compound required high energy) [77]. In contrast, for FSW, among the resulted energies for 700
rpm, 1000 rpm, and 1500 rpm rotational speeds with 30 mm/min welding speed, the lower energy 379 kJ (for 700
rpm) promoted superior joint strength of Al-to-steel as compared to 421 kJ and 652 kJ energies (for 1000 rpm and
1500 rpm respectively) [43]. Thus, this factor is also difficult to adopt with the force changing tendency during the
FSW process along the welding line [42, 79].
With regard to the heat effect on the IMC formation and its thickness on the shear strength, Tanaka et al. [3] found
an experimental relationship among the joint strength, the formed layer thickness, and the heat input for FSBW of
1100, 5052, and 7075-T6 Al Alloys to mild steel, as shown in Figure 17.

Te
nsi
le
str
en
gth
(M
Pa
)

Te
nsi
le
str
en
gth
(M
Pa
)

IMC thickness (m)

Heat input parameter ((m2.K)/N)

Figure 17.Effect of heat input and IMC thickness on the resulting tensile strength of different Al Alloys to steel [3].

Small IMCs at the upper part of Al-to-steel WZ have been observed while no IMC formation at the faying surface
[22, 24]. According to some studies, this zone represented the higher temperature region because of the shoulder
heat generation on the top surface [72]. This declaration supports the view of Sun et al. [51] and Ogura et al. [41]
about the heat dominancy on the IMC formation that mentioned above. In same study of Kimapong and Watanabe
[22],the steel fragments that spattered at the Al-to-steel interface zone were surrounded with an IMC layer, this can
be related to the relatively excessive stirring of these fragments when rotated around the pin tool which induced
friction and thereby heat. However, the tendency of the higher heat regions to accommodate the IMC formation was
not mentioned in other studies [3, 25]. Moreover, Dehghani et al. [5] noticed that, in some cases,the IMC formation
located at the bottom part of the weld.
The percentage of atoms of aluminum and iron along the IMC layer thickness are vary, producing different phases
of IMCs along that thickness. In Haghshenas et al. [6] and Bozzi et al. [48] investigations, the selected points along
the IMC layer thickness were produced different formed phases along that thickness (see also [43, 46]), as shown in
Figures 12 and 18.This different formation is heat and time dependent. In specimens diffusion treated by Kobayashi
and Yakou [77] at temperature greater than 1000 C and for a long diffusion treatment time AlFe and AlFe3 layers
were formed between the first formed layer (Al5Fe2) and steel.

22

Figure 18.Different IMC phase along the interfacial layer


thickness[43].

As mentioned before, that an amorphous phase formed before the IMC layer formation. Ogura et al. [41] indicated
that amorphous phase, it is most likely a kneading stage, leads to the IMC formation when heat being enough. Sun et
al. [51] emphasized that the amorphous phase is caused mainly by mechanical alloying during the excessive plastic
deformation, and not because of a thermally driven mechanism. They are convinced that the formation of IMCs is a
thermally driven diffusion process. Thus, when the peak temperature is relatively low, the IMC could not be found.
Furthermore, Springer et al. stated that the pre-deformation of the parent materials that caused by the FSW process
behaves as the pronounced effect on the composition and growth of the IMC layer. They used an annealing process
after welding to obtain the IMC layer [27].
Some studies suggested using annealing process after welding to promote the IMC layer. The used temperature and
the duration time are the two main factors affected this layer formation. There was no IMC formation occurred in
weldment of Springer et al., [27] and Movahedi et al., [44]. Therefore, they used this technique to form that metallic
layer and study its effect on the final joint strength. Broadly discussion on this technique will be shown later.
Chen and Nakata investigated the joint of aluminum and zinc-coated steel FSLW. They noticed a new formation,
Al-Zn, which was defined as a eutectic structure product besides the main IMC formation. The formation process of
this compound has been explained as follow; pure zinc on the steel surface is melted due to the friction, stir, and tool
plunging that generate high temperature and pressure. The high pressure makes oxide rapture at the aluminum
surface which promotes liquid Al-Zn phase formation that converted to a solid eutectic structure after welding
(cooling). Furthermore, the surface condition, whether zinc-coated or brushed-surface steel, has been reported to
play a major role in the IMC formation and tensile strength. No IMC was observed for brushed steel surface,
whereas two different types of IMC, Al13Fe4 and Al5Fe2, were observed at the Al/zinc-coated steel interface [37].
Feng et al. investigated the effect of steel coated layer on the joint strength and morphology for FSSW. They used
separately four types of steel plated with pure zinc, zinc alloy, Al-Si alloy, and zinc alloy including Fe, the melting
temperatures of these layers were 420, 330, 640 and 880 C respectively while the process temperature was 430 C.
It was found that strength and interface morphology of Al-to-steel joint were highly affected by the melting point of
coating layer. The lower melting temperature (lower than process temperature) of plated layer enhanced the joint
strength. Thus, steel plated with pure zinc or zinc alloy resulted in high joint strength compared to steel plated with
the other two coated layers. Thin intermetallic was observed at the joint interface of the Al-to-steel coated with pure
zinc or zinc alloy [49]. In general,the joint of Al-to-steel exhibited superior joint strength in lap welding when the
used steel is zinc-coated. The reason has not been clearly stated, nonetheless, the aluminum affinity to zinc is larger
than zinc to steel. This phenomenon form an intermetallic phase of zinc and aluminum, the other aluminum could be
expected to form Fe-rich IMC [60].
6.1.3 Controlling the IMC thickness
23

Evidently, slower welding speed and/or higher rotational speed for same tool design can assist IMC growth kinetic
[3,12]. These two parameters influence the generated heat and plastic deformation which are the main reasons of the
IMC formation. Generated force andrubbing effect between the rotated tool and the steel plate material which
caused bye or PLD distancesadd further heat[5,22,33].The other process parameters such as tool design and tool tilt
angle affectplastic deformation andthe generated heat as well [5,72]. Thin IMC formed when the steel plate in
FSLW coated with zinc, but there is no evidence that zinc can control this thickness[37, 49].
As conclusion, thin IMC layer can be achieved by employing lower rotational and higher welding speeds.The
selected tool design is preferred to produce low heat and material flow. For the same reason, the tilt angle should be
small. However, it is very difficult to join Al-to-steel and obtain free defect weld at the WZ with these process
parameters levels. Different claims aboutPLD distance effect on this IMC growing process were stated while e
distance effect is conjugated the rotational speed value.
6.1.4 Common factors
The available studies referred to many contributed factors in IMCformation and growing processes, Figure
19explains most of these parameters.Adopting a common factor to indicate the IMC formation and growing is
importantfor easily controlling this process. In FSW, the heat is generated by the effect of welding process
parameters such as rotational speed, welding speed, generated forces, and other welding parameters. More generated
heat is produces by increasing rotational speed and decreasing welding speed. The heat increased further by
increasing the forces throughusing bigger tool size, higher plunge depth, larger pin offset in steel plate, and large
tool tilt angle[5, 29, 42, 47, 72, 79]. This heat is the main affected factor of amorphous phase transformation process
to the IMC phases form [41, 51]. Therefore, temperature measurement (thermal cycle) along the welding line using
thermocouple is the best method to indicate the IMC phase and thickness changing.The steel surface condition in
FSLW is another factor whileannealing may be used after welding to generate this compound. Consequently, the
contributedfactors in IMC formation that explained in Figure 19can be reduced to the expressed one in Figure 20.

Force

Welding speed

Tool design

Tilt angle

Force
Heat input
Plunging and/or
diffusion

Material
flow

Material
flow

Rotational
speed
Heat generated

Heat input per


unit time

Forces
IMCs formation
and thickness

Different flow of
materials

Annealing

Roughness

Different flow
of material

Coating

Joint configuration
(Butt and Lap)

Heat treatment
after weld

Different chemical
compositions

Steel surface
condition in case of
Lap welding

Al Alloys and steel


types

Figure 19. Cause and effect diagram of the main affected parameters on the IMC formation.

24

Primary plastic
deformation and
material flow

Primary plastic
deformation and
material flow

Heat cycle during


welding process

Heat cycle during


annealing process

Surface condition in
case of FSLW

During welding process


After welding process

Note;The pressure assumed to


be constant during the FSW
process.

Plunging or diffusion
used techniques

Note;The surface condition


effect in annealing technique is
not well investigated yet.

Figure 20. Main affected parameters on the IMC formation at the


Al-to-steel interface when pressure assumed to be constant.

7. Annealing, plunging, and diffusion techniques


From the aforementioned literature, various techniques to promote the IMC layer and weld Al-to-steel by FSW
method have been used. Generally, these techniques can be categorized under three main topics:

Plunging; in butt Al-to-steel FSW, the pin is usually not inserted in the center of the faying line. An offset
toward the Al Alloy (distance e inFigure 3) should be used to avoid wearing out the pin and to attain good joint
strength [22- 33]. However, plunging in FSLW of Al-to-steel is achieved when the pin is plunge-in, by a PLD
distance that shown in Figure 3(PLD > 0), into the bottom plate(steel plate). Steel fragments are spattered inside
the stirring zone (SZ) of plunged welds, as shown in Figure 21. The steel fragments and its surrounding IMC
formation can enhance the findings [34]. It is claimed that an aluminum matrix composite was formed at the
WZ [21, 34].Nonetheless, the destructive test of Kimapong and Watanabe [22] shows the tendency of fracture
occurrence at the SZ caused by steel fragment existence, as shown in Figure 22. Another defect caused by pin
plunging which is hook defect [40]. Furthermore, this technique generates relatively high forces, especially for
the lap welding when the pin should be plunged into the steel plate; thus, the tool is wear-out [80].

Al alloy
Fracture
line
Steel
fragments

Steel

Al Alloy
Figure 21.Steel fragments spattered in weld nugget
because of the plunging [21].

25

Steel

Figure 22.Fracture location and path because of


the steel fragment[24].

Diffusion; As mentioned, the IMC formation is a thermally driven diffusion bonding [51], at the same time it
can help the joining feature of Al-to-steel[42, 48]. Therefore, Haghshenas et al. [6] and Chen et al. [36] made
full useful of these IMC formation concepts when they approved that the joining of Al-to-steel by FSLW can be
produced without pin plunging. The pin length used is shorter than the upper plate (Al Alloy) thickness, that is,
the pin tip did not plunge the bottom steel plate i.e. the PLDdistance in Figure 3 is 0. By controlling the
welding speed only, a bonding between the two metallic alloys has been obtained [6, 36]. This technique
prevented pin wear and high supplied force that presented by steel plate plunging. However, process with
plunging technique improved shear strength of by 20% more than diffusion one in study made byWei et al., but
with different welding speed (190 mm/min for plunging process and 150 mm/min for diffusion process) [45].

Chen et al., reported the joint strength of three pin plunging cases; I) PLD = +0.3 mmi.e. pin is plunged-in steel,
II) PLD = 0 mm i.e. pin tip is just touching steel plate surface, and III) PLD = -0.3 mm i.e. the pin is not
plunged-in steel. Same rotational 1400 rpm and welding 20 mm/min speeds were used for all experiments. The
optimum obtained strength was in case II while cases I & III lower by 31% and 73% respectively. Highpercent
of reduction in joint strength for diffusion case (III) can be related to the relatively big distance between pin tip
and steel surface (PLD = -0.3 mm)[42]. As mentioned, different phase can be formed at the Al-to-steel interface
due to different heat cycles. Furthermore,the IMC layer is controlled by the diffusion of Fe atoms into the IMC
layers. This can be attributed to the higher diffusion coefficient of iron into aluminum matrix 5310 -4 m2 s-1
(793K 922K) as compare to that of aluminum into iron matrix 1.810-4 m2 s-1 (1003K 1673K), the growth
rate of the IMC layer thickness decreases with increasing carbon content in the steel substrate, and it is inhibited
by silicon atoms [77].
In FSBW, Kimapong and Watanabe investigated two different offset distances (e in Figure 3); I)e = 0,i.e. the
pin is fully inserted in the Al Alloy side and merely touch the steel plate, and II)e = -0.2 mm,i.e. the pin is fully
inserted in the Al Alloy, they obtainedvery low joint strength for both cases. Perhaps, the used rotational speed
(250 rpm) was not enough to generate an adequate heat to form thin IMC at the faying interface[22]. For higher
rotational speed 900 rpm and e = 0 mm Girard et al.[26] achieved 59% as a joint efficiency. However,this
technique imply more investigations for both FSBW and FSLW to achieve low cost and robust Al-to-steel joint.
Annealing; this technique was used by Springer et al. [27] and Movahedi et al. [44]for the butt and lap FSW
joint of Al-to-steel. The technique helped the formation of the IMCs after welding, which generally decreased
the joint efficiency of Springer et al. [27] for FSBW. Bycontrast, an adequate improvement has been made on
the finding of Movahedi et al. [44] for FSLW. In general, high annealed temperature over a long time period
increases the IMC thickness and therefore decreases the joint efficiency, whereas relatively low annealing
temperature can achieve optimized joint efficiency. However, annealing technique can affect the mechanical
properties of the parent materials [81, 82]. Mertin et al., measured the temperature at Al-to-steel interface during
FSW process; the recorded temperature was 425 C. Then, they investigate the effect of the annealing
temperature and duration time on the strength of AA6061-T4 (parent material) using different temperature
values, 250C, 350C, and 450 C. A reduction of ~ 20% of tensile strength was observed when the annealed
temperatures were 350C and 450 C regardless the duration time. Meanwhile, the annealed temperature, 250
C, slightly improved the strength (5.6%) when a long annealing time, 30 sec, was used [82].Therefore, this
technique need for a careful use to avoid any effect on the mechanical properties of the parent materials.

26

Limitations
There are some limitations in this review. Reviewing thereference lists method to search for the related articles was
used; this procedure can ensure included more relevant articles in our investigation, some of these articles we did not
find. Besides, as we confined to companion checked process to included articles there may be some studies bias
form this review process. However, an expectation that non-peer reviewed articles would change these findings is
unsure.
Summary points
What was already known on the topic?
FSW has been used to weld aluminum and steel.
FSW has been customarily used to manufacture hybrid structure of aluminum and steel which can achieve low
weight transport application.
What this study added to our knowledge?
An overview on the main affected parameters of this joining process and the obtained joint efficiencies.
Despite the solid-state behaviour of this FSW method, the intermetallic compound AlxFey has been promoted
different following procedure of that in fusion welding, the fundamental effective factors areexplained.
Yet, the utilized strategy to obtain Al-to-steel joint arranged in three classifications; plunging, diffusion, and
annealing.
8. Conclusion
Friction stir welding is a relatively efficient method of welding aluminum to steel (Al-to-steel), using various
techniques, with comparably easy setup and low cost. A trial has been carried-out to present a systematic review of
the FSW of Al-to-steel. It is clear that high joint strength is attained when the weld zone promotes low defects and
steel fragments, a smooth gradient of grain size, and thin Fe-rich formed intermetallic layer. Similar to conventional
welding methods, the intermetallic compound formation represents the main problem issue in FSW. A cause and
effect diagram exposes many of the parameters that affect IMC formation and thickness. Heat cycle has a significant
effect on IMC thickening, phase transformation, and the elimination of the amorphous phase.
The hybrid techniques utilized in FSW for Al-to-steel joints reveal high joint strength in conjunction with high costs
of setup and equipment. These techniques modify the steel flow around the rotated pin by adding an assisted heat
source. This heat source shouldnt significantly affect the microstructure of the Al alloys.
Without using hybrid techniques, three major techniques are used in the FSW of Al-to-steel. First, the annealing
technique may affect the mechanical properties of the parent materials; particularly the Al alloys. Second, the
plunging technique result in fragments spattered at the weld zone matrix for both lap and butt welding
configurations and induces tool wear-out. Finally, the diffusion technique represents the minimum cost method to
join Al-to-steel. This technique requires further investigations to increase the joint strength and avoid the undesired
plunging and annealing outcomes. To the best of our knowledge, this technique (diffusion) has not been investigated
sufficiently in the existing studies of FSBW; such studies can add further important data.
References
References
[1]
[2]
[3]

C. . Chen and R. Kovacevic, Joining of Al 6061 alloy to AISI 1018 steel by combined effects of fusion and solid state welding, Int. J.
Mach. Tools Manuf., vol. 44, no. 11, pp. 12051214, Sep. 2004.
Y. Kusuda, Honda develops robotized FSW technology to weld steel and aluminum and applied it to a massproduction vehicle,Ind.
Robot An Int. J., vol. 40, no. 3, pp. 208212, Apr. 2013.
T. Tanaka, T. Morishige, and T. Hirata, Comprehensive analysis of joint strength for dissimilar friction stir welds of mild steel to
aluminum alloys, Scr. Mater., vol. 61, no. 7, pp. 756759, Oct. 2009.

27

[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]

W. Liu, J. Ma, M. Mazar Atabaki, and R. Kovacevic, Joining of advanced high-strength steel to AA 6061 alloy by using Fe/Al structural
transition joint, Mater. Des., vol. 68, pp. 146157, Mar. 2015.
M. Dehghani, a. Amadeh, and S. a. a. Akbari Mousavi, Investigations on the effects of friction stir welding parameters on intermetallic and
defect formation in joining aluminum alloy to mild steel, Mater. Des., vol. 49, pp. 433441, Aug. 2013.
M. Haghshenas, a. Abdel-Gwad, a. M. Omran, B. Gke, S. Sahraeinejad, and a. P. Gerlich, Friction stir weld assisted diffusion bonding
of 5754 aluminum alloy to coated high strength steels, Mater. Des., vol. 55, pp. 442449, Mar. 2014.
K. K. Ramachandran, N. Murugan, and S. Shashi Kumar, Effect of tool axis offset and geometry of tool pin profile on the characteristics
of friction stir welded dissimilar joints of aluminum alloy AA5052 and HSLA steel, Mater. Sci. Eng. A, vol. 639, pp. 219233, Jul. 2015.
M. Geiger, F. Micari, M. Merklein, L. Fratini, D. Contorno, a. Giera, and D. Staud, Friction Stir Knead Welding of steel aluminium butt
joints, Int. J. Mach. Tools Manuf., vol. 48, no. 5, pp. 515521, Apr. 2008.
M. Sahin, Joining of stainless-steel and aluminium materials by friction welding, Int. J. Adv. Manuf. Technol., vol. 41, no. 56, pp. 487
497, Apr. 2009.
E. Taban, J. E. Gould, and J. C. Lippold, Dissimilar friction welding of 6061-T6 aluminum and AISI 1018 steel: Properties and
microstructural characterization, Mater. Des., vol. 31, no. 5, pp. 23052311, May 2010.
M. M. R. Howlader, T. Kaga, and T. Suga, Investigation of bonding strength and sealing behavior of aluminum/stainless steel bonded at
room temperature, Vacuum, vol. 84, no. 11, pp. 13341340, Jun. 2010.
Y. C. Chen, a. Gholinia, and P. B. Prangnell, Interface structure and bonding in abrasion circle friction stir spot welding: A novel approach
for rapid welding aluminium alloy to steel automotive sheet, Mater. Chem. Phys., vol. 134, no. 1, pp. 459463, May 2012.
R. Cao, G. Yu, J. H. Chen, and P.-C. Wang, Cold metal transfer joining aluminum alloys-to-galvanized mild steel, J. Mater. Process.
Technol., vol. 213, no. 10, pp. 17531763, Oct. 2013.
S. Chen, J. Huang, K. Ma, X. Zhao, and A. Vivek, Microstructures and Mechanical Properties of Laser Penetration Welding Joint
With/Without Ni-Foil in an Overlap Steel-on-Aluminum Configuration, Metall. Mater. Trans. A, vol. 45, no. 7, pp. 30643073, Mar. 2014.
J. H. Kong, M. Okumiya, Y. Tsunekawa, K. Y. Yun, S. G. Kim, and M. Yoshida, A Novel Bonding Method of Pure Aluminum and
SUS304 Stainless Steel Using Barrel Nitriding, Metall. Mater. Trans. A, vol. 45, no. 10, pp. 44434453, Jun. 2014.
G. am and S. Mistikoglu, Recent Developments in Friction Stir Welding of Al-alloys, J. Mater. Eng. Perform., vol. 23, no. 6, pp. 1936
1953, Apr. 2014.
B. T. Gibson, D. H. Lammlein, T. J. Prater, W. R. Longhurst, C. D. Cox, M. C. Ballun, K. J. Dharmaraj, G. E. Cook, and a. M. Strauss,
Friction stir welding: Process, automation, and control, J. Manuf. Process., vol. 16, no. 1, pp. 5673, Jan. 2014.
Y. Zhao, L. Zhou, Q. Wang, K. Yan, and J. Zou, Defects and tensile properties of 6013 aluminum alloy T-joints by friction stir welding,
Mater. Des., vol. 57, pp. 146155, May 2014.
X. Hou, X. Yang, L. Cui, and G. Zhou, Influences of joint geometry on defects and mechanical properties of friction stir welded AA6061T4 T-joints, Mater. Des., vol. 53, pp. 106117, Jan. 2014.
Honda Co. Ltd., http://world.honda.com/news/2013/4130218New-Technology-Join-Steel-Aluminum/, 2013.
W. H. Jiang and R. Kovacevic, Feasibility study of friction stir welding of 6061-T6 aluminium alloy with AISI 1018 steel, Proc. Inst.
Mech. Eng. Part B J. Eng. Manuf., vol. 218, no. 10, pp. 13231331, Jan. 2004.
K. K. A. T. WATANABE, Friction Stir Welding of Aluminum Alloy to Steel, Weld. J., vol. October, pp. 277282, 2004.
H. Uzun, C. Dalle Donne, A. Argagnotto, T. Ghidini, and C. Gambaro, Friction stir welding of dissimilar Al 6013-T4 To X5CrNi18-10
stainless steel, Mater. Des., vol. 26, no. 1, pp. 4146, Feb. 2005.
T. Watanabe, H. Takayama, and A. Yanagisawa, Joining of aluminum alloy to steel by friction stir welding, J. Mater. Process. Technol.,
vol. 178, no. 13, pp. 342349, Sep. 2006.
W.-B. Lee, M. Schmuecker, U. A. Mercardo, G. Biallas, and S.-B. Jung, Interfacial reaction in steelaluminum joints made by friction stir
welding, Scr. Mater., vol. 55, no. 4, pp. 355358, Aug. 2006.
M. Girard, B. Huneau, C. Genevois, X. Sauvage, and G. Racineux, Friction stir diffusion bonding of dissimilar metals, Sci. Technol.
Weld. Join., vol. 15, no. 8, pp. 661665, Nov. 2010.
H. Springer, a. Kostka, J. F. dos Santos, and D. Raabe, Influence of intermetallic phases and Kirkendall-porosity on the mechanical
properties of joints between steel and aluminium alloys, Mater. Sci. Eng. A, vol. 528, no. 1314, pp. 46304642, May 2011.
R. S. Coelho, a. Kostka, J. F. dos Santos, and a. Kaysser-Pyzalla, Friction-stir dissimilar welding of aluminium alloy to high strength
steels: Mechanical properties and their relation to microstructure, Mater. Sci. Eng. A, vol. 556, pp. 175183, Oct. 2012.
M. Dehghani, S. a. a. A. Mousavi, and a. Amadeh, Effects of welding parameters and tool geometry on properties of 3003-H18 aluminum
alloy to mild steel friction stir weld, Trans. Nonferrous Met. Soc. China, vol. 23, no. 7, pp. 19571965, Jul. 2013.
M. Ghosh, R. K. Gupta, and M. M. Husain, Friction Stir Welding of Stainless Steel to Al Alloy: Effect of Thermal Condition on Weld
Nugget Microstructure, Metall. Mater. Trans. A, vol. 45, no. 2, pp. 854863, Nov. 2013.
S. Kundu, D. Roy, R. Bhola, D. Bhattacharjee, B. Mishra, and S. Chatterjee, Microstructure and tensile strength of friction stir welded
joints between interstitial free steel and commercially pure aluminium, Mater. Des., vol. 50, pp. 370375, Sep. 2013.
S. G. Sajan, M. Meshram, P. Srinivas, and S. R. Dey, Advanced Materials Friction Stir Welding of Aluminum 6082 with Mild Steel and its
Joint Analyses, Adv. Mater. Manuf. Charact., vol. 3, no. 1, pp. 189194, 2013.
X. Liu, S. Lan, and J. Ni, Analysis of process parameters effects on friction stir welding of dissimilar aluminum alloy to advanced high
strength steel, Mater. Des., vol. 59, pp. 5062, Jul. 2014.
A. Elrefaey, M. Gouda, M. Takahashi, and K. Ikeuchi, Characterization of Aluminum/Steel Lap Joint by Friction Stir Welding, J. Mater.
Eng. Perform., vol. 14, no. 1, pp. 1017, Feb. 2005.
K. Kimapong and T. Watanabe, Lap Joint of A5083 Aluminum Alloy and SS400 Steel by Friction Stir Welding, Mater. Trans., vol. 46,
no. 4, pp. 835841, 2005.
Y. C. Chen, T. Komazaki, Y. G. Kim, T. Tsumura, and K. Nakata, Interface microstructure study of friction stir lap joint of AC4C cast
aluminum alloy and zinc-coated steel, Mater. Chem. Phys., vol. 111, no. 23, pp. 375380, Oct. 2008.
Y. C. Chen and K. Nakata, Effect of the Surface State of Steel on the Microstructure and Mechanical Properties of Dissimilar Metal Lap
Joints of Aluminum and Steel by Friction Stir Welding, Metall. Mater. Trans. A, vol. 39, no. 8, pp. 19851992, May 2008.
R. S. Coelho, a. Kostka, S. Sheikhi, J. dosSantos, and a. R. Pyzalla, Microstructure and Mechanical Properties of an AA6181-T4
Aluminium Alloy to HC340LA High Strength Steel Friction Stir Overlap Weld, Adv. Eng. Mater., vol. 10, no. 10, pp. 961972, Oct. 2008.

28

[39] M. Movahedi, a. H. Kokabi, S. M. S. Reihani, and H. Najafi, Mechanical and Microstructural Characterization of Al-5083/St-12 lap joints
made by friction stir welding, Procedia Eng., vol. 10, pp. 32973303, 2011.
[40] Z. W. Chen and S. Yazdanian, Friction Stir Lap Welding: material flow , joint structure and strength, J. Achiev. Mater. Manuf. Eng., vol.
55, no. 2, pp. 629637, 2012.
[41] T. Ogura, Y. Saito, T. Nishida, H. Nishida, and T. Yoshida, Partitioning evaluation of mechanical properties and the interfacial
microstructure in a friction stir welded aluminum alloy / stainless steel lap joint, Scr. Mater., vol. 66, no. 8, pp. 531534, 2012.
[42] Z. W. Chen, S. Yazdanian, and G. Littlefair, Effects of tool positioning on joint interface microstructure and fracture strength of friction
stir lap Al-to-steel welds, J. Mater. Sci., vol. 48, no. 6, pp. 26242634, Dec. 2013.
[43] H. Das, S. Basak, and G. Das, Influence of energy induced from processing parameters on the mechanical properties of friction stir welded
lap joint of aluminum to coated steel sheet, Int. J. Adv. Manuf. Technol., vol. 64, pp. 16531661, 2013.
[44] M. Movahedi, A. H. Kokabi, S. M. S. Reihani, W. J. Cheng, and C. J. Wang, Effect of annealing treatment on joint strength of aluminum /
steel friction stir lap weld, Mater. Des., vol. 44, pp. 487492, 2013.
[45] Y. Wei, J. Li, J. Xiong, and F. Zhang, Effect of Tool Pin Insertion Depth on Friction Stir Lap Welding of Aluminum to Stainless Steel, J.
Mater. Eng. Perform., vol. 22, no. 10, pp. 30053013, 2013.
[46] H. Das, R. N. Ghosh, and T. K. Pal, Study on the Formation and Characterization of the Intermetallics in Friction Stir Welding of
Aluminum Alloy to Coated Steel Sheet Lap Joint, Metall. Mater. Trans. A, vol. 45, no. 11, pp. 50985106, Jul. 2014.
[47] H. Das, S. S. Jana, T. K. Pal, and a. De, Numerical and experimental investigation on friction stir lap welding of aluminium to steel, Sci.
Technol. Weld. Join., vol. 19, no. 1, pp. 6975, Jan. 2014.
[48] S. Bozzi, A. L. Helbert-etter, T. Baudin, B. Criqui, and J. G. Kerbiguet, Intermetallic compounds in Al 6016 / IF-steel friction stir spot
welds, Mater. Sci. Eng. A, vol. 527, no. 1617, pp. 45054509, 2010.
[49] K. Feng, M. Watanabe, and S. Kumai, Microstructure and Joint Strength of Friction Stir Spot Welded 6022 Aluminum Alloy Sheets and
Plated Steel Sheets, Mater. Trans., vol. 52, no. 7, pp. 14181425, 2011.
[50] Y. Uematsu, T. Kakiuchi, Y. Tozaki, and H. Kojin, Comparative study of fatigue behaviour in dissimilar Al alloy/steel and Mg alloy/steel
friction stir spot welds fabricated by scroll grooved tool without probe, Sci. Technol. Weld. Join., vol. 17, no. 5, pp. 348356, Jul. 2012.
[51] Y. F. Sun, H. Fujii, N. Takaki, and Y. Okitsu, Microstructure and mechanical properties of dissimilar Al alloy/steel joints prepared by a
flat spot friction stir welding technique, Mater. Des., vol. 47, pp. 350357, May 2013.
[52] O. Lorrain, V. Favier, H. Zahrouni, and D. Lawrjaniec, Understanding the material flow path of friction stir welding process using
unthreaded tools, J. Mater. Process. Technol., vol. 210, no. 4, pp. 603609, Mar. 2010.
[53] T. Yasui, N. Ando, S. Morinaka, H. Mizushima, and M. Fukumoto, Effect of weld line shape on material flow during friction stir welding
of aluminum and steel, IOP Conf. Ser. Mater. Sci. Eng., vol. 61, p. 012009, Aug. 2014.
[54] Y. Morisada, T. Imaizumi, H. Fujii, M. Matsushita, and R. Ikeda, Three-Dimensional Visualization of Material Flow During Friction Stir
Welding of Steel and Aluminum, J. Mater. Eng. Perform., vol. 23, no. 11, pp. 41434147, Aug. 2014.
[55] T. Khaled and D. Ph, An Outsider Looks at Friction Stir Welding, vol. 90712, no. July, 2005.
[56] H. Zhang, S. B. Lin, L. Wu, J. C. Feng, and S. L. Ma, Defects formation procedure and mathematic model for defect free friction stir
welding of magnesium alloy, Mater. Des., vol. 27, no. 9, pp. 805809, Jan. 2006.
[57] Y. Su, X. Hua, and Y. Wu, Effect of input current modes on intermetallic layer and mechanical property of aluminumsteel lap joint
obtained by gas metal arc welding, Mater. Sci. Eng. A, vol. 578, pp. 340345, Aug. 2013.
[58] G. Qin and Y. Su, Microstructures and properties of welded joint of aluminum alloy to galvanized steel by Nd:YAG laser + MIG arc
hybrid brazing-fusion welding, Trans. Nonferrous Met. Soc. China, vol. 24, pp. 989995, 2014.
[59] S. Meco, G. Pardal, S. Ganguly, S. Williams, and N. McPherson, Application of laser in seam welding of dissimilar steel to aluminium
joints for thick structural components, Opt. Lasers Eng., vol. 67, pp. 2230, 2015.
[60] J. Ma, M. Harooni, B. Carlson, and R. Kovacevic, Dissimilar joining of galvanized high-strength steel to aluminum alloy in a zero-gap lap
joint configuration by two-pass laser welding, Mater. Des., vol. 58, pp. 390401, 2014.
[61] J. L. Song, S. B. Lin, C. L. Yang, and C. L. Fan, Effects of Si additions on intermetallic compound layer of aluminum-steel TIG weldingbrazing joint, J. Alloys Compd., vol. 488, no. 1, pp. 217222, 2009.
[62] H. Zhang, J. Liu, and J. Feng, Effect of auxiliary TIG arc on formation and microstructures of aluminum alloy/stainless steel joints made
by MIG welding-brazing process, Trans. Nonferrous Met. Soc. China, vol. 24, no. 9, pp. 28312838, 2014.
[63] W. Zhang, D. Sun, L. Han, and Y. Li, Optimised design of electrode morphology for novel dissimilar resistance spot welding of
aluminium alloy and galvanised high strength steel, Mater. Des., vol. 85, pp. 461470, 2015
[64] J. Schneider and R. Radzilowski, Welding of Very Dissimilar Materials (Fe-Al), Jom, vol. 66, no. 10, pp. 21232129, Sep. 2014.
[65] D.-H. Choi, C.-Y. Lee, B.-W. Ahn, J.-H. Choi, Y.-M. Yeon, K. Song, S.-G. Hong, W.-B. Lee, K.-B. Kang, and S.-B. Jung, Hybrid Friction
Stir Welding of High-carbon Steel, J. Mater. Sci. Technol., vol. 27, no. 2, pp. 127130, Feb. 2011.
[66] Y. F. Sun, Y. Konishi, M. Kamai, and H. Fujii, Microstructure and mechanical properties of S45C steel prepared by laser-assisted friction
stir welding, Mater. Des., vol. 47, pp. 842849, May 2013.
[67] J. Luo, W. Chen, and G. Fu, Hybrid-heat effects on electrical-current aided friction stir welding of steel, and Al and Mg alloys, J. Mater.
Process. Technol., vol. 214, no. 12, pp. 30023012, Dec. 2014.
[68] M. Merklein and a. Giera, Laser assisted Friction Stir Welding of drawable steel-aluminium tailored hybrids, Int. J. Mater. Form., vol. 1,
no. S1, pp. 12991302, Apr. 2008.
[69] H. Bang, H. Bang, G. Jeon, I. Oh, and C. Ro, Gas tungsten arc welding assisted hybrid friction stir welding of dissimilar materials Al6061T6 aluminum alloy and STS304 stainless steel, Mater. Des., vol. 37, pp. 4855, May 2012.
[70] T. Azimzadegan and S. Serajzadeh, An Investigation into Microstructures and Mechanical Properties of AA7075-T6 during Friction Stir
Welding at Relatively High Rotational Speeds, J. Mater. Eng. Perform., vol. 19, no. 9, pp. 12561263, Mar. 2010.
[71] Z. Y. Ma, Friction Stir Processing Technology: A Review, Metall. Mater. Trans. A, vol. 39, no. 3, pp. 642658, Feb. 2008.
[72] . Frigaard, . Grong, and O. T. Midling, A process model for friction stir welding of age hardening aluminum alloys, Metall. Mater.
Trans. A, vol. 32, no. 5, pp. 11891200, May 2001.
[73] B. Y. M. Roulin, J. W. Luster, G. Karadeniz, and A. Mortensen, Strength and Structure of Furnace-Brazed Joints between Aluminum and
Stainless Steel, Weld. Res., no. May, pp. 151155, 1999.

29

[74] M. Potesser, T. Schoeberl, H. Antrekowitsch, and J. Bruckner, The Characterization of the Intermetallic Fe-Al Layer of Steel-Aluminum
Weldings, Miner. Met. Mater. Soc., vol. EPD Congre, pp. 167176, 2006.
[75] H. Ozaki and M. Kutsuna, Dissimilar Metal Joining of Zinc Coated Steel and Aluminum Alloy by Laser Roll Welding, in Welding
Processes, Ch 2, 2012, p. 35.
[76] Y. Su, X. Hua, Y. Wu, Y. Zhang, and Y. Guo, Characterization of intermetallic compound layer thickness at aluminumsteel interface
during overlaying, Mater. Des., vol. 78, pp. 14, Aug. 2015.
[77] S. Kobayashi and T. Yakou, Control of intermetallic compound layers at interface between steel and aluminum by diffusion-treatment,
Mater. Sci. Eng. A, vol. 338, no. 12, pp. 4453, Dec. 2002.
[78] C.-Y. Lee, D.-H. Choi, Y.-M. Yeon, and S.-B. Jung, Dissimilar friction stir spot welding of low carbon steel and Al--Mg alloy by
formation of IMCs, Sci. Technol. Weld. Join., vol. 14, no. 3, pp. 216220, 2009.
[79] D. Trimble, J. Monaghan, and G. E. ODonnell, Force generation during friction stir welding of AA2024-T3, CIRP Ann. - Manuf.
Technol., vol. 61, no. 1, pp. 912, 2012.
[80] T. DebRoy and H. K. D. H. Bhadeshia, Friction stir welding of dissimilar alloys a perspective, Sci. Technol. Weld. Join., vol. 15, no. 4,
pp. 266270, May 2010.
[81] F. X. Zhao, X. C. Xu, H. Q. Liu, and Y. L. Wang, Effect of annealing treatment on the microstructure and mechanical properties of
ultrafine-grained aluminum, Mater. Des., vol. 53, pp. 262268, Jan. 2014.
[82] C. Mertin, A. Naumov, L. Mosecker, M. Bambach, and G. Hirt, Influence of the Process Temperature on the Properties of Friction Stir
Welded Blanks Made of Mild Steel and Aluminum, Key Eng. Mater., vol. 611612, pp. 14291436, May 2014.

30

You might also like