You are on page 1of 5

Materials Science & Engineering A 866 (2023) 144641

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Bonding mechanism and fracture behavior of inertia friction welded joint of


2219 aluminum alloy to 304 stainless steel
Zongyu Dang , Guoliang Qin *, Juan Wang
School of Materials Science and Engineering, Shandong University, Jinan, 250061, PR China

A R T I C L E I N F O A B S T R A C T

Keywords: The 2219 aluminum (Al) alloy bars were welded to 304 stainless steel bars by inertia friction welding. The
Bonding mechanism intermetallic compounds (IMCs) at the friction interface were characterized and analyzed and the bonding
Bonding strength mechanism of the welded joint was studied in detail. Characterization of the microstructure at the friction
Intermetallic compounds
interface revealed that there were two types of IMCs layers, the nanoscale Fe4Al13 layer and the microscale Cu-
Cu-rich layer
Al alloy/steel
rich layer. In particular, the Cu-rich layer is bonded to the base material by Al2Cu and two Fe–Al–Cu ternary
Friction welding IMCs. Meanwhile, it was verified both theoretically and experimentally that the Fe4Al13 layer had less misfit with
the base material with higher bonding strength at the interface. The formation of the Cu-rich layer increases the
lattice mismatch, particularly at the interface of the ternary IMCs layer which becomes a weak area where the
crack initiates and propagates easily.

1. Introduction the one hand, Kimura et al. [9–12] reported that Al alloy/steel RFW
joints have optimal mechanical properties under the condition that no
The hybrid structure of aluminum (Al) alloy and stainless steel dis­ IMCs layer is generated at the friction interface, yet no explanation was
similar metals can realize lightweight products while meeting the me­ provided for how the welded joints are bonded when the IMCs layer is
chanical properties of the structure [1] and satisfying the urgent needs of not generated. On the other hand, Ma et al. [13] and Fukumoto et al.
modern industry of saving energy, light structures, and high strength [2, [14] revealed that the amorphous layer is a transition layer that trans­
3]. Therefore, the high-quality joining of Al alloy to stainless steel dis­ forms into IMCs layers with crystalline structure at a certain temperature
similar materials can promote the application of hybrid structures in in Al alloy/steel friction welding joints. Meanwhile, the bonding
automotive, aerospace and other fields requiring lightweight materials strength of the welding interface reaches a maximum when forming the
[4–6]. However, the physical and chemical properties of Fe and Al are amorphous layer, but the strength decreases when the thickness of the
very different from each other. In addition, the solid solubility of both IMCs layer exceeds 100 nm [13]. It is generally regarded that the gen­
are very low at room temperature, making it difficult to achieve a eration of IMCs marks when the welding interface achieves metallur­
joining by solid solution. In contrast, at high temperatures, brittle and gical bonding. The type, thickness, and distribution of Fe–Al IMCs have a
hard intermetallic compounds (IMCs) are easily formed, preventing the decisive influence on the mechanical properties of Al alloy/steel welded
use of conventional fusion welding and mechanical joining from joints. The research conducted on Al alloy/steel friction welding shows
achieving an effective bonding that meets industrial requirements [7,8]. that the IMCs at the friction interface mainly consist of FeAl, Fe2Al5 and
As a solid phase welding method, rotary friction welding (RFW) has Fe4Al13 [5,15–18]. Different types of IMCs match different bonding
the inherent advantages of low heat input and high efficiency to ensure strengths with the base material (BM). The thickness of the IMCs layer
the mechanical properties of Al alloy/steel joints, especially for the also has a critical value (0.5–3 μm), below which the joint matches the
joining of cylindrical structures. Currently, although there are many higher tensile strength [19–21]. Therefore, the phenomenon of inho­
papers investigating the bonding mechanism of Al alloy/steel friction mogeneous thermo-mechanical coupling during RFW makes the thick­
welding, certain controversial issues are still present. Three major ness of the IMCs layer at different locations of the friction interface
bonding mechanisms are currently available: (1) no IMCs layer gener­ different, thus affecting the integral mechanical properties of the welded
ated, (2) joining by the amorphous layer, (3) joining by IMCs layer. On joints [22,23]. Nevertheless, current research indicates that for 1xxx,

* Corresponding author. Qianfoshan Campus of Shandong University, 17923 Jingshi Road, Jinan, 250061, PR China.
E-mail address: glqin@sdu.edu.cn (G. Qin).

https://doi.org/10.1016/j.msea.2023.144641
Received 1 November 2022; Received in revised form 11 January 2023; Accepted 12 January 2023
Available online 21 January 2023
0921-5093/© 2023 Elsevier B.V. All rights reserved.
Z. Dang et al. Materials Science & Engineering A 866 (2023) 144641

5xxx and 6xxx Al alloy to steel friction welding joints, the joint strength off when a predetermined speed was reached. The energy stored by the
can match the BM of the Al alloy, even if the tensile fracture occurs at the inertia of the flywheel under the action of friction pressure bonds the
BM. The above results show that for the three series of Al alloy/steel dissimilar metals together, as shown in Fig. 1 (a). After welding, speci­
friction welding joints with different bonding mechanisms, their weld­ mens cut by electrical discharge machining were prepared for subse­
ability is better, and they can achieve high quality bonding with steel. quent characterization in the direction of the perpendicular friction
However, for the 2xxx Al alloy used extensively in the aerospace interface. Metallographic specimens followed the standard metallo­
field, the addition of a large number of Cu elements causes the micro­ graphic process with grinding and subsequent polishing.
structure and mechanical properties of Al alloy/steel friction welded
joints to be different from those of other series Al alloys. The IMCs layer 2.3. Microstructural characterization and tensile test
is dominated by Al–Cu IMCs and contains almost no Fe elements, which
is very distinct from the generation of the Fe–Al IMCs layer in the joints The microstructure and chemical composition of the friction inter­
of other series Al alloy [24]. In contrast, the weldability of 2xxx Al alloy face were investigated and analyzed by scanning electron microscope
to steel is generally poor, with the strength of the welded joints much (SEM, JSM-7800 F) and field emission electron microprobe analysis
lower than that of the BM of Al alloy [25–27]. The addition of Cu (EPMA, JXA-8530 F PLUS). To further investigate the bonding mecha­
element alters the solid state metallurgical reaction behavior of Al nism of the friction interface, the focused ion beam (FIB, FEI Strata 400
alloy/steel friction welding joints, which severely damages their me­ S) technique was used to sample the selected interface zone. The
chanical properties [28]. However, there is no definitive conclusion microstructure at the friction interface was studied by transmission
about the bonding mechanism and bonding strength of the IMCs layer to electron microscopy (TEM, FEI TECNAI G2 F20). Scanning transmission
the BM for 2xxx Al alloy/steel joints. electron microscopy (STEM) mode using selected area diffraction (SAD)
Based on the above understanding and analysis, this work has been patterns, high angle annular dark field (HAADF) imaging and high
conducted to characterize and study 2219 Al alloy (AA2219) to 304 resolution TEM (HRTEM) imaging were applied in combination with
stainless steel (304SS) inertia friction welding (IFW) joints. The bonding energy dispersive X-ray spectroscopy (EDS, genesis 4000) for the
mechanism of the friction interface is first clarified. Then, the bonding elemental analysis.
methods and strength of the Al–Cu IMCs layer with the BM are eluci­ To measure the tensile strength, the flash of the welded joints was
dated both experimentally and theoretically. The results of the study are removed by the lathe, as shown in Fig. 1 (c). Then a universal tensile
useful in guiding the regulation of the welding technology and metal­ machine (E45, MTS) was used to test the tensile samples under a tensile
lurgical reaction aspects of 2xxx Al alloys/steel friction welding joints rate of 1 mm/s in a displacement-controlled mode. Multi-stage tensile
from a microscopic perspective. tests under the 160–1300 welding parameters were conducted to
investigate the fracture behavior of the dissimilar joints, which meant
2. Materials and experiments that the tensile process was manually stopped when the specimen was
stretched to the specified strength before the joint was fractured. Then
2.1. Materials tensile specimens at different stages were fabricated according to the
procedure in section 2.2 to make standard metallographic specimens
AA2219-T8 and 304SS bars with diameters of 15 mm were used as which were used to study the bonding strength and fracture behavior of
the BM for the IFW tests. The chemical compositions of the materials the friction interface.
were measured by optical emission spectrometer (M12, Spectro-Lab), as
shown in Table 1. It is noticeable that AA2219 is a typical aging 3. Results
strengthening Al–Cu alloy. The solid solubility of Cu in Al reaches 5.65
wt% at the eutectic temperature, whereas it is only 0.5 wt% when 3.1. Microstructural appearance of friction interface
reduced to room temperature [29]. The Al–Cu phases at room temper­
ature mainly include θ”, θ’ and θ, where θ is a stable equilibrium phase Fig. 2 showed the SEM and EPMA results for the distribution and
[30]. Thus, the Al–Cu IMCs at the friction interface are related to the chemical composition of IMCs layers at the friction interface. The fric­
strengthening phase of the Al alloy [24]. tion interface clearly existed with the IMCs layer and without the IMCs
layer in two different bonding mechanisms, as shown in Fig. 2 (a). The
2.2. Welding parameters and sample preparation IMCs layer formed at the interface was also noticed to be discontinuous
and inhomogeneous, with a thickness of about 2 μm or less. The EPMA
The experiment was conducted on an HSMZ-4 inertia friction weld­ results of the elemental mapping scanning of the two typical welded
ing machine with a maximum loading force of 40 kN. The main IFW interfaces are shown in Fig. 2 (b). The location of the layer without IMCs
parameters were the initial rotation speed, friction pressure and rota­ was dominated by the interdiffusion of Fe–Al elements, with a small
tional inertia of the flywheel. The optimal parameters of 1300 rpm, 160 number of Cu elements aggregated at the interface, whereas the IMCs
MPa, and 0.49 kg m2 (abbreviated to 160–1300), respectively, were layer contained significant amounts of Al and Cu elements, which far
obtained after numerous welding experiments. Under these conditions, exceeded the amount of Fe element. This phenomenon was also found at
the tensile strength of the welded joint reached 241.5 MPa, which is the AA2219-O/304SS friction welded joint and this layer was identified
51.5% of that of the Al alloy BM [31]. The standard deviation is 10.46 as the Al2Cu layer [24]. It was also noticed that the Cu content was
and the average axial shortening of the welding joints at the optimal higher in the IMCs layer near the Al alloy side, which may indicate that
parameters is 8.69 mm. the IMCs layer was not composed of a single IMC. Therefore, we defined
Before welding, a lath was used to remove the oxide layer from the this layer as the Cu-rich layer.
BM end face before wiping with alcohol to clean the grease. During the The formation and distribution of the Cu-rich layer were not
welding process, the motor that drove the flywheel rotation was turned consistent with previously reported results for IMCs layers et al. alloy/

Table 1
Chemical compositions of base materials (wt. %).
Base materials Al Si Cu Mn C Ni Cr Zr Fe

AA2219 Bal 0.0542 6.06 0.292 – – – 0.147 0.147


304SS – 0.46 – 0.97 0.16 8.82 17.5 – Bal

2
Z. Dang et al. Materials Science & Engineering A 866 (2023) 144641

Fig. 1. (a) Al alloy/steel IFW process; (b) typical zones in metallurgic sample for microstructure analysis; (c) tensile sample of the welded joint.

Fig. 2. Distribution and chemical composition of IMCs layer at the friction interface: (a) typical distribution of IMCs layer at the friction interface; (b) and (c)
chemical composition of the friction interface at different zones.

steel friction welded joints. However, how the metallurgical bonding 3.1.1. Cu-rich IMCs layer
between the Cu-rich layer and the BM was achieved needs to be further Fig. 3 illustrated the specific information of the Cu-rich layer struc­
characterized and studied. For locations where the IMCs layer was not ture, as derived from the TEM analysis. Fig. 3 (a)–(c) showed the dis­
formed, only the interdiffusion of Fe–Al elements was noticed, and tribution of the main elements of Fe, Cu and Al in the IMCs layer. The
whether the IMCs layer was formed to achieve metallurgical bonding IMCs layer was enriched by the Al and Cu elements, with few Fe ele­
was unclear. ments, which is in accordance with the results of the EPMA mapping
analysis shown in Fig. 2. Fig. 3 (d) showed an STEM image of the IMCs
layer and the SAD patterns for different position in the IMCs layer were

3
Z. Dang et al. Materials Science & Engineering A 866 (2023) 144641

Fig. 3. Microstructure of Cu-rich IMCs layer: (a)–(c) distribution of Fe, Cu and Al elements in IMCs layer; (d) HADDF image of IMCs layer and the corresponding SAD
patterns of different grains; (e) magnified view of the white rectangle in (d); (f) HRTEM image of the white rectangle in (e); (f1) fast Fourier transformation (FFT) of
the HRTEM image in (f); (i) result of EDS line scanning in (e).

analyzed. These grains could be identified as two different IMCs layers according to the measurement of the interplanars distance and their
according to their SAD patterns: a Al2Cu layer near the Al alloy and an angles and in HRTEM image, as shown in Fig. 3 (f).
intermediate transition layer of Al63⋅5Cu24Fe12.5, as shown in Fig. 3 (d). Therefore, a thicker Al2Cu IMCs layer was formed near the Al alloy
There was also a lamellar IMCs layer near the steel, which was difficult mainly at the position of the Cu-rich IMCs layer, with the former
to determine by SAD due to its thinness. According to the results of the achieving an effective bond with the steel through two different
EDS mappings shown in Fig. 2 (a), it was inferred that this lamellar layer Fe–Al–Cu ternary IMCs layers. Currently, no paper has reported the
may be a ternary IMCs layer of Fe–Al–Cu. Fig. 3 (e)–(i) showed a further formation of the Fe–Al–Cu ternary IMCs layer for the structure of the Cu-
characterization of the lamellar IMCs layer near the steel by HRTEM. rich IMCs layer at the friction interface of 2xxx series Al alloy/steel
The selected area in Fig. 3 (e) was enlarged to obtain Fig. 3 (d), on which friction welded joints.
EDS line scanning analysis were conducted. The elemental distribution
results demonstrated that this lamellar layer had significantly less Cu 3.1.2. Fe–Al IMCs layer
and more Fe elements, and no noticeable change in Al elements, as In section 3.1.1, the structure of the Cu-rich IMCs layer was inves­
shown in Fig. 3 (i). There was also no significant diffusion of Cr and Ni tigated, but IMCs layers were not observed in SEM images of most po­
elements in the steel. Therefore, we concluded that a Fe–Al–Cu ternary sitions of the friction interface. Therefore, how the friction interface
IMCs layer was formed, and its Fe element content was higher than that realized an effective bonding when IMCs layers were not formed re­
of Al63⋅5Cu24Fe12.5, whereas the Cu element was lower than that of quires further study. Exemplary HRTEM analysis results from the posi­
Al63⋅5Cu24Fe12.5. This lamellar IMCs layer was identified as Al65Cu20Fe15 tions of the unformed IMCs layer under the 160–1300 welding

4
Z. Dang et al. Materials Science & Engineering A 866 (2023) 144641

parameters were shown in Fig. 4. Fig. 4 (a) and (b) represent the center
zone and the 1/2 R zone of the welded joint, respectively. The TEM
specimen in the center zone still had no obvious IMCs layer present in
HADDF image, as shown in Fig. 4 (a). Further magnification by HRTEM
revealed that the lattice fringe at the interface was quite different from
that of the BM. The FFT image shown in Fig. 4 (a1) provided better
contrast to confirm the existence of the intermediate phase. A nanoscale
Fe4Al13 layer was generated at the interface with a thickness of only
10–15 nm, as determined by measuring the planers distance and their
angles. However, for the 1/2 R zone, as shown in Fig. 4 (b), there was an
IMCs layer formed at the interface. Combined with HRTEM and FFT
images shown in Fig. 4 (b1) and (b2), it was confirmed that the nano­
scale Fe4Al13 layer was formed near the steel side, whereas the Al2Cu
layer was on the Al alloy side. The precipitation of Al2Cu layers at the 1/
2 R zone was associated with the inhomogeneity of frictional heat
generation during IFW. In the center zone, the welding temperature was
lowest at the welded joint due to the near zero linear velocity during
rotation; in comparison to the higher welding temperature in the 1/2 R
zone, the Al2Cu layer was formed by the precipitation of more fine
strengthening phases in the Al alloy at the welding interface.
Therefore, it was clear that nanoscale IMCs layers were formed at Fig. 5. Multi-stage tensile tests under the 160–1300 welding parameter.
locations where IMCs layers were not observed in the SEM field of view,
which were mainly composed of Fe4Al13 layer or Fe4Al13 and Al2Cu occurred at the partial location of the Cu-rich layers, where the cracks
layers. were mainly located at the bonding interface between the Cu-rich layer
and the steel, as shown in Fig. 6 (d). The Fe4Al13 layer still maintains a
3.2. Bonding strength and fracture behavior high strength as the tensile strength was increased to 180 MPa. Not only
were no obvious welding defects found at the Fe4Al13 layer, but also
Section 3.1 investigated the bonding method of the friction interface, found that some cracks extend through the steel side without occurring
where the inhomogeneous Cu-rich layers were the main difference be­ at the interface, as shown in Fig. 6 (e). Certainly, the extension of the
tween 2xxx Al alloy/steel joints and other series Al alloy/steel joints. cracks on the steel side did not mean that the bonding strength of the
The effects of the two different IMCs layers on the tensile strength of the interface had exceeded that of the steel. This was mainly due to the
welded joints had not been reported in previous publications. micro-cracks left when using the lathe to remove the oxide layer on the
To study the interfacial strength and fracture behavior, the tensile steel side before welding, which expanded when stretched under higher
process was manually stopped at strengths of 100, 180 and 200 MPa, as stresses. However, when the cracks extended to the interface, the
shown in Fig. 5. The fracturing at the friction interface at different stages Fe4Al13 layer did not fracture, which also indicated that the interface
was also investigated. Firstly, the microstructure at the interface of the maintained a high strength. Fig. 6 (g) demonstrates that the Fe4Al13
welded joint in the case of a non-tensioning process was presented in layer in parts of the interface still did not crack when the tensile strength
Fig. 6 (a) and (b). No obvious welding defects, such as cracks or the was increased to 200 MPa. However, most of the Cu-rich layers were
unbonded areas, were found at either the Fe4Al13 layer or the Cu-rich cracked, which were mainly located at the interface between the Cu-rich
layer. When the tensile strength reached 100 MPa, no crack was layer and the steel side. Cracks extension also occurred inside the Cu-
observed at the Fe4Al13 layer, as shown in Fig. 6 (c). However, cracks rich layer and between the Cu-rich layer and the Al alloy side, and

Fig. 4. Microstructure of the Fe–Al layer: (a) Fe–Al layer at the center zone; (b) Fe–Al layer at the 1/2 R zone.

You might also like