Sei sulla pagina 1di 8

Materials and Design 53 (2014) 504511

Contents lists available at ScienceDirect

Materials and Design


journal homepage: www.elsevier.com/locate/matdes

Microstructures and mechanical property of laser butt welding


of titanium alloy to stainless steel
Shuhai Chen, Mingxin Zhang, Jihua Huang , Chengji Cui, Hua Zhang, Xingke Zhao
School of Materials Science and Engineering, University of Science and Technology Beijing, Beijing 100083, PR China

a r t i c l e

i n f o

Article history:
Received 8 March 2013
Accepted 12 July 2013
Available online 24 July 2013
Keywords:
Laser welding
Titanium alloy
Stainless steel
Intermetallic compounds

a b s t r a c t
Laser butt welding of titanium alloy to stainless steel was performed. The effect of laser-beam offsetting
on microstructural characteristics and fracture behavior of the joint was investigated. It was found that
when the laser beam is offset toward the stainless steel side, it results in a more durable joint. The intermetallic compounds have a uniform thickness along the interface and can be divided into two layers. One
consists of FeTi + a-Ti, and other consists of FeTi + Fe2Ti + Ti5Fe17Cr5. When laser beam is offset by 0 mm
and 0.3 mm toward the titanium alloy side, the joints fracture spontaneously after welding. Durable joining is achieved only when the laser beam is offset by 0.6 mm toward the titanium alloy. From the top to
the bottom of the joint, the thickness of intermetallic compounds continuously decreases and the following interfacial structures are found: FeAl + a-Ti/Fe2Ti + Ti5Fe17Cr5, FeAl + a-Ti/FeTi + Fe2Ti + Ti5Fe17Cr5
and FeAl + a-Ti, in that order. The tensile strength of the joint is higher when the laser beam is offset
toward the stainless steel than toward the titanium alloy, the highest observed value being 150 MPa.
The fracture of the joint occurs along the interface between two adjacent intermetallic layers.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
The joining of titanium alloy to conventional structural steel has
many industrial applications [1,2]. However, it is difcult to weld
titanium alloys to steel due to great differences in thermal, physical,
and chemical properties. According to the FeTi phase diagram [3],
the solubility of Fe in Ti is very low (0.1 at.%, at room temperature),
beyond which, intermetallic phases FeTi and then Fe2Ti (600 HV and
1000 HV respectively) begin to form [4]. These intermetallic phases
are highly brittle, causing conventional fusion-welded joints to
crack spontaneously, due to thermal-stress mismatch between the
two parent materials. Therefore, suppressing the formation of brittle intermetallic compounds (IMCs) is the key to realizing reliable
joining. Diffusion bonding [5,6], brazing [7,8], and explosive welding [9,10] of steel and titanium has been investigated for this
purpose. These methods are effective to some extent in controlling
the formation of brittle intermetallic compounds, but their applications are restricted by joint congurations.
Friction stir welding is a novel solid-state welding method
which has been considered for joining titanium alloys to steel. Fazel-Najafabadi et al. [11,12] achieved defect-free dissimilar lap
joints of CP-Ti to 304 stainless steel by adjusting friction-stir welding parameters. The maximum failure load of the joint reached 73%
of that of CP-Ti. Liao et al. [13] investigated the microstructures at
Corresponding author. Tel.: +86 010 62334859.
E-mail address: jhhuang@ustb.edu.cn (J. Huang).
0261-3069/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.matdes.2013.07.044

the interface of friction-stir lap joints of pure titanium and steel.


Swirling macro- and micro-intermixing zones of titanium and steel
were found along the interface, where tiny FeTi intermetallic particles were mixed with b titanium [13]. The investigations on
friction stir welding of titanium and steel focused mostly on the
lap conguration, but it was insufcient to study the butt
conguration.
Fusion welding, which is widely applied in industry, is confronted with a great challenge in joining titanium alloys to steel.
In traditional arc welding, it is very difcult to control the molten
pool of titaniumsteel mixture, so a mass of intermetallic compounds are formed during welding. Recently, the development of
high energy beam welding such as laser beam welding and electron beam (EB) wending have made fusion welding of titanium
alloy to steel possible.
To reduce the residual tensile stress of the joint and improve
metallurgical reaction of the molten pool, copper was selected to
be the ller metal for electron beam welding of titanium alloy to
steel [14,15]. Electron beam welding needs to be performed in a
vacuum environment, which restricts its application. Laser welding, a new joining technology, has great advantages, such as high
efciency, excellent controllability, and ability to be performed in
a non-vacuum environment. Hiraga et al. [16] applied a pulse laser
lap welding technique to join thin sheets of pure titanium and
stainless steel 304. Zhao et al. [17], suggested the laser lap welding
technique for joining titanium alloy to steel. They also obtained
further understanding of the welding process by FEM model

505

S. Chen et al. / Materials and Design 53 (2014) 504511

calculation. The laser butt welding of titanium alloy to steel need


Cu interlayer to improve metallurgical reaction and decrease residual stress. Tomashchuk et al. [18] reported that welding of Ti6Al4 V
to stainless steel AISI 316L through pure copper interlayer carried
out by pulsed Nd:YAG laser. The tensile strength of the welds is up
to 359 MPa and is limited by brittleness of CuTi2 + FeTi + a-Ti layer
situated next to the solid Ti6Al4 V. Groza et al. [19] welded
X5CrNi1810 to Ti6Al4V through a copper foil, with the thickness of 600 lm, as intermediary layer by continuous NdYAG laser.
The average value of tensile strength reached 328 Mpa. Thus, Using
Cu foil as an interlayer leads to acceptable strength characteristics
during laser welding of titanium alloy to steel.
Laser butt welding of titanium alloy to steel without ller metal
is not pay more attention. In fact, direct laser butt welding of titanium alloy to steel is convenience in technology. Satoh et al. [20]
attempt to weld titanium alloy to stainless steel without ller metal by laser welding, but tensile strengths of the joints were observed to be lower than the base materials with failure occurring
through brittle fracture. And, investigation on the microstructures
is insufcient. To improve mechanical property of the joint, it is
necessary to understand deeply the weldability of titanium alloy
to steel without ller. Therefore, the weld characteristics and
microstructure should be investigated systematically.
The present work focuses on laser butt welding of titanium alloy to stainless steel without a ller metal. To control metallurgical
reactions in the molten pool, the welding experiments were performed with various laser beam offsets, either toward the titanium
alloy or toward the stainless steel. The effect of laser beam offset
on the microstructural characteristics and mechanical property of
the joint was investigated.

Fig. 1. Schematic diagram of laser welding of titanium alloy to stainless steel. (a)
The process and (b) tensile specimen (thickness: 1 mm).

tometer (XRD). To increase diffracted intensity of the phases and


analyze broken position of the joints, the XRD tests were performed at the fracture surface of the tensile specimen after tensile
test. The tensile strength of the joints at room temperature was
evaluated by a mechanical testing frame (INSTRON MODEL 1186)
at a cross head speed of 1 mm/min. The schematic diagram of
the tensile specimen is shown in Fig 1b). The tensile test standards
are in accordance with ISO 6892: 1998 [23].

2. Experiments
3. Results
The plates of Ti6Al4V titanium alloy (100  50  1 mm3) and
201 stainless steel (100  50  1 mm3) were selected as laser
welding materials and their chemical compositions were shown
in Table 1. The plates manufacturing of titanium alloy and stainless
steel conrm to GB/T3621-94 [21] and GB/T3280-92 [22], respectively. The interface sides of the specimens were carefully polished
and then clamped in a butt-weld geometry. A high-power (4 kW)
CO2 laser with welding-head focal distance of 200 mm was used.
The laser beam was made to focus on the upper surfaces of the
specimens. All of the specimens were welded at a power of 2000 W
and a welding speed of 2 m/min to ensure full penetration of the
specimens. Inuence of the laser beam offsets on the mechanical
property and microstructures of the joints were investigated because the formation of the intermetallic compounds is sensitive
to the offsets. The offsets toward stainless steel were dened as positive and those toward titanium alloy as negative, as shown in
Fig. 1a). In this study, welding with offsets of
0.6 mm,
0.3 mm, 0 mm, 0.3 mm and 0.6 mm were performed.
The microstructures of the joint were observed by scanning
electron microscope (SEM) equipped with an energy-dispersive
X-ray spectrometer (EDS) which has 2% error for middle atom
number, after standard grinding and polishing procedures. Crystal
phases of the joints were identied by microbeam X-ray diffrac-

3.1. Weld appearance


Fig. 2 shows the effect of the laser beam offsets on the surface
appearance of weld. When the offset is 0.6 mm (toward titanium
alloy), a durable joining was realized, as shown in Fig. 2a. If the offsets are 0 mm and 0.3 mm (toward titanium alloy), the specimens crackle noticeably during welding and then fractured
spontaneously after welding. Fig. 2b shows surface appearance of
the weld with a 0 mm offset. A number of cracks appear in the center of the weld, resulting in spontaneous fracture after welding. It
appears that the weldability is bad when laser beam is offset toward titanium alloy.
When the laser beam is offset toward stainless steel, all of the
specimens are joined durably, and the quality of the joint is improved signicantly. Fig. 2c shows the surface appearance of a
weld with a 0.6 mm offset. Compared to the 0.6 mm-offset weld,
the surface appearance of weld is relatively smooth. The weld zone
was narrower because stainless steel has higher thermal conductivity (12.1 W/m K, 20 C) than titanium alloy (5.44 W/m K, 20 C)
[24]. Therefore, laser butt welding of titanium alloy to stainless
steel has better weldability when the laser beam is offset toward
stainless steel than toward titanium alloy.

Table 1
Chemical composition of materials used (wt%).
Name
Titanium
Name
Steel

Fe
60.3
C
60.15

C
60.1
Si
60.75

N
60.05
Mn
5.57.5

H
60.015
Cr
16.018.0

O
60.2
N
60.25

Al
5.56.8
Ni
3.55.5

V
3.54.5
P
60.03

Ti
rest
S
60.06

Fe
rest

506

S. Chen et al. / Materials and Design 53 (2014) 504511

Fig. 2. Inuence of the laser beam offsets on the surface appearance of welds. (a)
Offset of 0.6 mm, (b) offset of 0 mm and (c) offset of 0.6 mm.

3.2. Microstructures of the joint


3.2.1. Offset toward stainless steel
Fig. 3 shows the microstructures of the joint made by a laser
beam with a 0.6 mm offset toward the stainless steel. A layer of
intermetallic compounds with uniform thickness formed along
the interface between titanium alloy and stainless steel, as shown
in Fig. 5a. The total thickness of the intermetallic compounds was
about 30 lm. The structures can be divided into two reaction layers, as shown in Fig. 3b. Let the reaction layer adjacent to titanium
alloy be called layer I and the one adjacent to stainless steel,
layer II. Layer I consisted of two phases in a matrix-and-reticular
structure. Layer II seemed to have complex structures composed of
four phases, as shown in Fig. 3c.
To identify phase compositions of the reaction layers, EDS was
performed on zones A through F in Fig. 3c. The results are shown
in Table 2. According to the EDS results, these phases have complex
element compositions. The substitution of the atoms in the lattice
should be considered in identifying phase composition. Since the
binary systems of TiV, TiCr, FeCr and FeNi are mutually soluble in solid state at high temperature, V and Cr can replace Ti atoms
and Cr and Ni can replace Fe atoms in the intermetallic lattice. It
appears that the Fe atoms in the reticular phase (A zone) can be replaced by Cr and Ni atoms while Ti atoms can be replaced by V
atoms. In contrast, the Fe atoms in the matrix phase (B zone) can

Fig. 3. Microstructures of the joint at the offset of laser beam toward to stainless
steel. (a) Macro cross-section, (b) interfacial microstructures and (c) magnication
in the Fig. 3b).

be replaced by Ni atoms and Ti atoms can be replaced by Cr and


V atoms. The positions of components of reticular and matrix
phases on the TiFeAl ternary phase diagram [25] are shown by
A and B in Fig. 4. The components are close to the eutectic liquidus
of FeAl and b-Ti. The reticular and matrix phases seem to be composed of FeAl and b-Ti. However, the b-Ti may transform into a-Ti
at a low temperature. The crystal structure of the Ti needs to be
determined by XRD.
The compositions of the dark (C zone) and gray (D zone) phases
are shown by C and D in Fig. 4. It can be deduced that the dark and
gray phases are FeTi and Fe2Ti, respectively. In contrast, the light
grey phase (E zone) and white phase (F zone) have a low content

S. Chen et al. / Materials and Design 53 (2014) 504511

507

deduced that the light grey phase (E zone) and the white phase
(F zone) are Fe2Ti and s (Ti5Fe17Cr5), respectively.
It can be seen that gray phases (D zone) and light grey phase (E
zone) are Fe2Ti phases. The contrast difference of the phases is
caused by different average atom numbers between D and C zones.
Therefore, interfacial structures of the joint are FeAl + Ti/
FeTi + Fe2Ti + Ti5Fe17Cr5.
3.2.2. Offset toward titanium alloy
Fig. 6 shows the microstructures of the joint made by the laser
beam offset of 0.6 mm toward titanium alloy. Complex structures
appear at the interface. First, partial melting of stainless steel

Fig. 4. Liquidus surface of the FeTiAl ternary phase diagram [25].

Table 2
EDS results of the joints (at.%).
Position

Ti

Fe

Cr

Ni

Al

A
B
C
D
E
F
G
H
I
J

57.51
60.77
49.58
40.92
25.99
12.24
27.35
15.71
4.43

24.31
20.90
32.18
36.38
51.31
58.87
53.44
57.34
66.06
70.02

5.31
5.58
7.06
9.97
11.84
18.55
13.36
19.17
22.55
21.06

1.78
2.10
3.22
5.18
7.17
5.73
3.47
4.85
4.95
7.97

1.52
1.77
1.14
1.59
0.49
1.01
0.64
0.53
0.69
0.24

9.57
8.87
6.83
5.96
3.13
3.60
1.73
2.40
1.32
0.70

Fig. 5. Liquidus surface of the FeCrTi ternary phase diagram [26].

of Al, as shown in Table 2. Zones E and F can then be considered a


FeCrTi ternary system and phase compositions can thus be conrmed. The locations of these phases on the FeCrTi ternary
phase diagram [26] are as shown by E and F in Fig. 5. It can be

Fig. 6. Microstructures of the joint at the offset of laser beam toward to titanium
alloy. (a) Macro cross-section, (b) microstructures at the top of the joint and (c)
microstructures at the bottom of the joint.

508

S. Chen et al. / Materials and Design 53 (2014) 504511

appears at the top of the joint, as shown in Fig. 6a. Second, the
thickness gradually decreases from the top to the end of the joint.
Third, different interfacial structures appear at the top, middle and
bottom of the joint. At the top of the joint in Fig. 6b, two reaction
layers are formed. A grey layer is close to the fusion zone, dened
as Layer III in this paper. It can be deduced that the phase compositions of Layer III is same with Layer I because they have the
same morphology. The other white layer is close to stainless steel,
dened as Layer IV. In this layer, a light grey dendritic structure
grows into Layer IV from the interface of Layer III and Layer IV. A
light grey fusion line appears between Layer IV and parent stainless steel. In the middle of the joint indicated by a rectangle in
Fig. 6b, the interfacial structures are the same as that made by
the laser beam offset toward stainless steel, as shown in Fig. 3b.
The structures consist of two layers of FeAl + Ti/FeTi + Fe2Ti + Ti5Fe17Cr5. At the bottom of the joint in Fig. 6c, only one reaction layer
is formed between titanium alloy and stainless steel. It is conrmed that this layer has same structures with Layer I (FeAl + Ti).
According to above observation on the microstructures of the
joint, the phase compositions of Layer IV need to be conrmed.
In Layer IV, since the Al content is relatively low, the phase compositions can be identied as a FeCrTi ternary system. The compositions of the grey dendritic phase growing into Layer IV (G zone)
and white phase of the Layer IV (H zone) in the FeCrTi ternary
phase diagram are shown by D and E in Fig. 4. It can be believed
that the grey dendritic phase is Fe2Ti and the white phase is ternary intermetallic compound s (Ti5Fe17Cr5). At fusion line (I zone),
the content of Ti slightly increases, compared to parent stainless
steel (J zone).
Therefore, the interfacial structures of the joint are from the top
to the bottom conrmed orderly as FeAl + Ti/Fe2Ti + Ti5Fe17Cr5 ?
FeAl + Ti/ FeTi + Fe2Ti + Ti5Fe17Cr5 ? FeAl + Ti.
3.3. mechanical property of the joint
To evaluate the mechanical property of the joint, tensile
strength tests were performed. Fig. 7 shows the inuence of laser-beam offsets on the tensile strength of the joints. Because the
specimens with the offsets of 0.3 mm and 0 mm fracture spontaneous after the welding, the tensile strength is zero. When the offset is 0.6 mm, the tensile strength of the joint is only 24.75 MPa.
In contrast, when the offsets are 0.3 mm and 0.6 mm, the tensile
strength of the joints increase signicantly. When laser beam is
offset to 0.6 mm toward stainless steel, the tensile strength of
the joint reaches 150 MPa. Therefore, the joints have higher tensile

Fig. 7. Inuence of the offset of laser beam on the tensile strength.

strength when the laser beam is offset toward the stainless steel
than the toward the titanium alloy.
3.4. Fracture behavior of the joint
Fig. 8 shows fracture surface of the joint made with laser-beam
offset of 0.6 mm toward titanium alloy. The fracture exhibits typical brittle characteristics. A relatively smooth fracture surface with
a river patten appears at the top of the joint. In contrast, the fracture at the middle and end of the joint exhibits relatively rough
morphology. In addition, many cracks along the thickness direction
of the plate appear at the fracture surface. When the offset is
0.6 mm toward stainless steel, the fracture morphology still exhibits typical brittle characteristics. However, the relatively smooth
fracture surface with river patten disappears and all fracture surface becomes to relatively rough, as shown in Fig. 9. Also, the
cracks along the thickness direction of the plate is more dense than
that with laser beam offset of 0.6 mm toward titanium.
To conrm phase compositions at the fracture surface, XRD was
performed on all the samples, as shown in Fig. 10. Because there
are complex element compositions of the joints, distortion of crystal lattice is inevitable. Thus, the location of the diffraction peaks of
the phase compositions may be offset from the location of the standard diffraction peak. It was found that a-Ti and Fe2Ti appear at all
of the fracture surfaces. The b-Ti do not appears at the fracture surfaces, which indicates that the b-Ti in the eutectic structures of b-Ti
and FeTi transforms into a-Ti during the cooling of the weld.
Therefore, it is conrmed that the Layer I consists of a-Ti and FeTi.
The presence of the a-Ti at all fracture surfaces may indicatethat the joint fractures along the layer which contains a-Ti and FeTi
(Layer I or Layer III). However, the presence of the Fe2Ti seems to
indicate that the joint fractures along Layer II or Layer IV. Therefore, it can be concluded that fracture only occurs along the interface between two adjacent layers.
It should be noted that some of the phases observed in SEM
were not found by the XRD. In general, the XRD results depend
on diffracted intensity of the phases. Thus, the XRD were performed at fracture surface to increase the diffracted intensity
[27]. However, some of the intermetallic compounds in this study
are uncontinuous and inhomogeneous in distribution, such as Ti5Fe17Cr5. And, the content is relatively low. Therefore, some of the
phases maybe are not found by the XRD, although the XRD were
performed at fracture surface.
4. Discussion
When titanium alloy is mixed with the stainless steel in the liquid state, a number of brittle intermetallic compounds are formed
inside fusion zone during welding. Meanwhile, a stress mismatch
arises in the seam due to great difference in thermal expansion
coefcients between titanium alloy (7.89  10 6/C) and stainless
steel (16.9  10 6/C) [24]. The specimens fracture spontaneously
when a sufcient amount of intermetallic compounds has formed.
To improve the weldability of titanium alloy to steel, the Cu interlayer was used during laser welding [18,19]. On the one hand, the
Cu interlayer improved metallurgical reaction of weld pool, which
leads to the formation of TiCu intermetallic compounds [19]. On
the other hand, the interlayer decrease residual stress of the joint
[14]. Therefore, the tensile strength of the joint with the Cu interlayer reached to 359 MPa [18].
Outside using the Cu interlayer, suppressing liquid-state mixing
between titanium alloy and steel is an important way to realize
strong joining. Satoh et al. [20] attempted to weld titanium alloy to
steel without interlayer by Nd: YAG laser welding, but liquid-states
mixing between titanium alloy and steel is not suppressed fully.

S. Chen et al. / Materials and Design 53 (2014) 504511

509

Fig. 8. Fracture surface of the joint with laser beam offset of 0.6 mm toward to
titanium. (a) Macro fracture surface, (b) magnication of P1 zone and (c)
magnication of P2 zone.

Fig. 9. Fracture surface of the joint with the offset of 0.6 mm toward to stainless
steel. (a) Macro fracture surface, (b) magnication of P3 zone and (c) magnication
P4 zone.

Perhaps samples were processed at laser beam offsets from the


interface toward the Ti plate, which is not good way proved in this paper. In our study, the liquid-states mixing is suppressed better by

controlling laser beam offset toward titanium alloy. Thus, laser beam
offsetting toward titanium alloy is an effective way to suppress liquid-state mixing during welding of titanium alloy to stainless steel.

510

S. Chen et al. / Materials and Design 53 (2014) 504511

Fig. 10. XRD patterns of fracture surface of the joint. (a)

According to the results of the experiment, laser butt welding of


titanium alloy to stainless steel has better weldability when the
laser beam is offset toward stainless steel than toward titanium
alloy. This is closely related to the distribution and compositions
of the intermetallic compounds, which depend on the laser beam
offsets.
The intermetallic compounds along the interface between
titanium alloy and stainless steel are more uniform in thickness
when the laser beam is offset toward stainless steel than when
it is offset toward titanium alloy. This difference in thickness is
a result of great difference in thermal conductivity between
stainless steel (12.1 W/m K, 20 C) and titanium alloy (5.44 W/
m K, 20 C) [24]. Stainless steel has a higher thermal conductivity
than titanium alloy, which means that the thermal conduction
from the high temperature zone of the molten pool to the low
temperature zone is easier when laser beam is offset toward
the stainless steel rather than titanium. Thus, a more uniform
temperature distribution is achieved when the laser beam is offset toward stainless steel, which is the side with higher thermal
conductivity. The low temperature gradient along the interface
induces uniform thickness of intermetallic compounds. The
uneven thickness distribution of the intermetallic compounds
has negative inuence on the mechanical property of the joint.
First, it causes an uneven distribution of residual stress in the
seam, compared to the joint with uniform distribution of the
intermetallic compounds. Furthermore, stress concentration is
induced inside the seam. The residual stress compromises
mechanical property of the joint. Second, the residual stress
causes cracks to initiate at the top of the joint, where the thick
intermetallic compounds layer is. In particular, this causes the
specimens with the 0.3 mm and 0 mm offsets fracture spontaneously after welding. Third, the uneven thickness distribution
of the intermetallic compounds makes it more difcult to
control mechanical property of the joint: when the amount intermetallic compounds at the top of the joint are limited to a
reasonable level, the bottom of the joint would not have been
reliably welded yet; when the bottom of the joint is welded with
a reasonable amount of intermetallics, too much intermetallics
would have formed on the top of the joint. Therefore, the joint

0.6 mm, (b)

0.3 mm, (c) 0.3 mm and (d) 0.6 mm.

with uniform thickness of intermetallic compounds, made by a


laser beam offset toward the stainless steel, has high tensile
strength.
When the laser beam is offset toward titanium alloy, the relatively smooth fracture surface with river patten at the top of the
joint indicates that the crack propagation only required low energy, as shown in Fig. 8b. The low propagation energy of the crack
leads to low tensile strength of the joint. According to the microstructures in Fig. 6, slight melting appears at the top of the joint,
which indicates the mixing between titanium alloy and stainless
steel. This proves that the mixing between titanium alloy and
stainless steel adversely affects the mechanical property of the
joint. The fracture morphologies when laser beam is offset toward
titanium alloy are similar with that observed by Satoh et al. [20],
because similar processing parameters were used. Meanwhile, the
phase compositions of this zone (FeAl + a-Ti/Fe2Ti + Ti5Fe17Cr5)
greatly differ with other areas (FeAl + a-Ti/FeTi + Fe2Ti + Ti5Fe17
Cr5 or FeAl + a-Ti) of the joint. Therefore, phase composition
may be another important factor which contributes to low tensile
strength of joint made when the laser beam is offset toward the
titanium alloy.

5. Conclusions
From the investigation on microstructures and mechanical
property of laser butt welding of titanium alloy to stainless steel,
conclusions were summarized as following:
(1) When the laser beam is offset toward the stainless steel, durable joining is realized. The intermetallic compounds have
uniform thickness along the interface and can be divided into
two layers: FeTi + a-Ti and FeTi + Fe2Ti + Ti5Fe17Cr5.
(2) When the laser beam is offset by 0 mm and 0.3 mm toward
the titanium alloy, the joints fracture spontaneously subsequent to welding. Durable joining is obtained only when
the laser beam is offset by 0.6 mm toward the titanium alloy.
From the top to the bottom of the joint, the thickness of
intermetallic compounds continuously decreases and the

S. Chen et al. / Materials and Design 53 (2014) 504511

following interfacial structures are found: FeAl + a-Ti/Fe2


Ti + Ti5Fe17Cr5,
FeAl + a-Ti/FeTi + Fe2Ti + Ti5Fe17Cr5
and
FeAl + a-Ti, in that order.
(3) The tensile strength of the joint is higher when the laser
beam is offset toward stainless steel than toward titanium
alloy, the highest value being 150 MPa. Fracture occurs along
the interface between two adjacent layers.

Acknowledgments
The authors appreciate the nancial support from the National
Natural Science Foundation of China (No. 51004009) and the Fundamental Research Funds for the Central Universities (FRF-TP-12044A). The authors would like to express their gratitude to Dr. Bert
Liu, Department of Materials Science and Engineering, The Ohio
State University, for his help in English revision of this article.
References
[1] Mudali UK, Rao BMA, Shanmugam K, Natarajan R, Raj B. Corrosion and
microstructural aspects of dissimilar joints of titanium and type 304L stainless
steel. J Nucl Mater 2003;321:408.
[2] Ghosh M, Chatterjee S, Mishra B. The effect of intermetallics on the strength
properties of diffusion bonds formed between Ti5.5Al2.4V and 304 stainless
steel. Mater Sci Eng A 2003;363:26874.
[3] Murray JL. Phase diagrams of binary titanium alloys. ASM, International; 1987.
p. 99111.
[4] Akbari Mousavi SAA. Sartangi PF. Effect of post-weld heat treatment on the
interface microstructure of explosively welded titaniumstainless steel
composite. Mater Sci Eng A 2008;494:32936.
[5] Kundu S, Roy D, Chatterjee S, Olson D, Mishra B. Inuence of interface
microstructure on the mechanical properties of titanium/17-4 PH stainless
steel solid state diffusion bonded joints. Mater Des 2012;37:5608.
[6] Kundu S, Sam S, Chatterjee S. Interface microstructure and strength properties
of Ti6Al4V and microduplex stainless steel diffusion bonded joints. Mater
Des 2011;32:29973003.
[7] Yue X, He P, Feng JC, Zhang JH, Zhu FQ. Microstructure and interfacial reactions
of vacuum brazing titanium alloy to stainless steel using an AgCuTi ller metal.
Mater Charact 2008;59:17217.
[8] Lee MK, Lee JG, Lee JK, Hong SM, Lee SH, Park JJ, et al. Formation of interfacial
brittle phases sigma phase and IMC in hybrid titanium-to-stainless steel joint.
Nonferrous Met Soc China 2011;21:s7s11.

511

[9] Akbari Mousavi SAA. Farhadi Sartangi P. Experimental investigation of


explosive welding of cp-titanium/AISI 304 stainless steel. Mater Des
2009;30:45968.
[10] Song J, Kostka A, Veehmayer M, Raabe D. Hierarchical microstructure of
explosive joints: example of titanium to steel cladding. Mater Sci Eng A
2011;528:26417.
[11] Fazel-Najafabadi M, Kashani-Bozorg SF, Zarei-Hanzaki A. Dissimilar lap joining
of 304 stainless steel to CPTi employing friction stir welding. Mater Des
2011;32:182432.
[12] Fazel-Najafabadi M, Kashani-Bozorg SF, Zarei-Hanzaki A. Joining of CPTi to
304 stainless steel using friction stir welding technique. Mater Des
2010;31:48007.
[13] Liao JS, Yamamoto N, Liu H, Nakata K. Microstructure at friction stir lap joint
interface of pure titanium and steel. Mater Lett 2010;64:231720.
[14] Zhang BG, Wang T, Duan XH, Chen GQ, Feng JC. Temperature and stress elds
in electron beam welded Ti-15-3 alloy to 304 stainless steel joint with copper
interlayer sheet. Trans Nonferrous Met Soc China 2012;22:398403.
[15] Wang T, Zhang BG, Chen GQ, Feng JC, Tang Q. Electron beam welding of Ti-15-3
titanium alloy to 304 stainless steel with copper interlayer sheet. Trans.
Nonferrous Met Soc China 2010;20:182934.
[16] Hiraga H, Fukatsu K, Ogawa K, Nakayama M, Mutoh Y. Nd: YAG laser welding
of pure titanium to stainless steel. Quart J JWS 2001;19:71726.
[17] Zhao SS, Yu G, He XL, Zhang YJ, Ning WJ. Numerical simulation and
experimental investigation of laser overlap welding of Ti6Al4V and 42CrMo.
J Mater Process Technol 2011;211:5307.
[18] Tomashchuk I, Sallamand P, Andrzejewski H, Grevey D. The formation of
intermetallics in dissimilar Ti6Al4V/copper/AISI 316L electron beam and
Nd:YAG laser joints. Intermetallics 2011;19:146673.
[19] Groza C, Mitelea I, Utu ID, Craciunescu CM. Melted zone morphology by laser
welding of Ti6Al4V With X5CRNI18-10. In: 21nd International Conference
on Metallurgy and Materials. Brno, Czech Republic; 2012, p. 113.
[20] Satoh G, Yao YL, Qiu C. Strength and microstructure of laser fusion-welded Ti
SS dissimilar material pair. Int J Adv Manuf Tech 2013;66:46979.
[21] GB/T3621-94. Titanium and titanium alloy sheet and plate. The State Standard
of the Peoples Republic of China, 1994.
[22] GB/T3280-92. Cold rolled stainless steel sheets and plates. The State Standard
of the Peoples Republic of China, 1992.
[23] ISO 6892: 1998. Metallic materials Tensile testing at ambient temperature.
International Organization for Standardization; 1998.
[24] ASM: ASM metals handbook, In: 10th ed., ASM International, Metals Park
(OH); 1990.
[25] Palm M, Lacaze J. Assessment of the AlFeTi system. Intermetallics
2006;14:1291303.
[26] Ivanchenko V, Pryadko T. Chromium Iron Titanium MSIT. In: Effenberg G,
Ilyenko S, editors. Berlin: Springer-Verlag; 2008. p. 117.
[27] Kundu S, Anand G, Chatterjee S. Diffusion bonding of 174 precipitation
hardening stainless steel to Ti alloy with and without Ni alloy interlayer:
Interface microstructure and mechanical properties. Metall Mater Trans A
2013;44:2196211.

Potrebbero piacerti anche