Microstructure and mechanical properties of 2060 Al–Li alloy welded by alternating current cold metal transfer with high-frequency pulse current
-
Liwei Wang
and Dianlong Wang
Abstract
Al–Li alloy has been widely used in the aerospace field owing to its high strength and low density. In this study, alternating current cold metal transfer (AC CMT) along with a high-frequency pulse current technique was used to weld a 2060 Al–Li alloy using an ER5356 wire. The effect of pulse frequency on the arc shape, microstructure, and mechanical properties of the welded joints was examined, and mechanical performance testing was conducted. The results revealed that the arc diameter, arc length, and arc volume showed a trend of increasing first and then decreasing with an increase in the pulse frequency and reached their peak values when the pulse frequency was 50 kHz. Coupling the welding process with a high-frequency pulse resulted in grain refinement, which was attributed to the stirring action of the arc force. Both the porosity levels and grain size decreased with increasing frequency. When the pulse frequency was 70 kHz, the porosity level was the lowest, and the grain size was refined to 24.1 μm. The tensile strength of the welded joints also increased with the pulse frequency, and a maximum tensile strength of 249 MPa was observed at 70 kHz.
1 Introduction
Owing to the advantages of low density, excellent resistance to stress corrosion cracking, and enhanced mechanical properties such as high strength and good fracture toughness, Al–Li-based alloys are widely used in aircraft and aerospace structures. Compared with the previous second generation of Al–Li alloys, the third generation of Al–Li alloys has displayed superior mechanical behavior in terms of strength and hardness and improved weldability with lower anisotropy. This could be attributed to the chemical composition of third-generation Al–Li alloys, which have a lower content of Li and an increased content of Cu, Zr, and Mn [1,2]. Nevertheless, there are still some issues with Al–Li alloys during welding, such as weld metal porosity and cracks [3]. Studies have shown that the use of appropriate filler materials and preparation of high-quality weld surface are good methods for reducing the weld defects of Al–Li alloys [4]. Therefore, the strength of Al–Li weld joints can be greatly improved by choosing appropriate filler materials and effective welding parameters. Mousavi and Sabzi [5] studied the effect of pouring temperature and the surface angle of vortex casting on the microstructural evolution and mechanical properties of 7050 Al–3wt% SiC composite. Mechanical property measurements showed that the yield strength, hardness, and fracture toughness of the composite were increased by increasing the pouring temperature and reducing the surface angle during vortex casting.
Effective methods for the welding of Al–Li alloys include laser beam welding (LBW) using high-energy beams and friction stir welding (FSW) using solid-phase connections [6,7,8]. LBW offers several advantages such as a lower heat application than conventional arc welding and a reduction in porosity and cracks during the welding process. Different Al–Li alloys (2060 and 2198) were laser-welded without the addition of filler materials [9,10]. The effects of welding parameters on the formation of welded joints, microstructure evolution, solute segregation, porosity, and their relationship with the mechanical properties of the joints were studied. This research has found that reducing welding heat input can effectively prevent grain coarsening, while reducing porosity and thermal cracking tendency, thereby improving joint performance. Liu et al. [10] further studied the influence of the microscopic morphology of different regions on the mechanical properties of welded joints and found that the microhardness distribution corresponds to the grain morphology distribution. The average microhardness at the center of the weld was the smallest because the grain size at the center of the weld was the largest. In addition, Cu and Mg segregation occurred during the welding process, which reduced the strength of the weld. However, Al–Li alloys have high reflectivity for laser beams, and alloying elements with lower melting points evaporate. The radiation of FSW is extremely small, the welded joints are flawless, and there is no need for chamfering when welding thick plates. Solutions have also been found for the three main problems associated with FSW, which are back support, weld thinning, and small hole defects [11]. Furthermore, refill friction stir spot welding (RFSSW) is a new technique for joining metal structures [12], which is used to connect thin aluminum elements; the welded joint exhibits good performance when the welding parameters are appropriate. The current research focus is to explore other efficient welding technologies to meet the requirements of low-cost production of Al–Li alloy products.
Recent studies on arc welding have demonstrated the positive effect of impulse current and frequency in grain refinement during the solidification of the weld region in aluminum alloys. For instance, the grain size was reduced with a higher pulse current frequency in an arc welding process [13,14]. During alternating current (AC) arc welding, the coupling of a high-frequency and moderate pulse current frequency has significant effects on the solidification process of the molten metal [8]. This has often resulted in improved mechanical properties of weld joints. Researchers have demonstrated that the issue of porosity can be effectively addressed by implementing AC arc welding. A reasonable adjustment of the time ratio of the EN phase and the EP phase in AC arc welding has resulted in a reduction in porosity and grain size [15]. Moreover, optimizing the AC arc welding parameters has been proven to have a positive effect on the metallurgical characteristics of the weld metal [16].
In the recent past, there have been several advancements in gas metal arc welding processes (GMAW). Cold metal transfer (CMT) is a type of GMAW (Figure 1) with substantial benefits, such as maintaining a “controlled heat” during the welding process. In this automated welding process, there is minimal heat due to a short circuit that controls (i.e., retracts and advances) the welding filler material [17]. Therefore, the arc introduces heat for a brief period, resulting in a spatter-free material transfer and a smooth weld. This unique characteristic (i.e., controlled heat input) makes this an ideal technique for welding Al–Li alloys. In addition, the CMT welding process has several advantages, such as a narrow welding thermal zone, and uniform welding seam. Moreover, it can effectively clean the oxide film while welding aluminum alloys [17].
![Figure 1
CMT droplet transition [17].](/document/doi/10.1515/htmp-2021-0026/asset/graphic/j_htmp-2021-0026_fig_001.jpg)
CMT droplet transition [17].
Therefore, this study chose a new process that combines high-frequency pulse current with an AC CMT welding process to weld aluminum–lithium alloys. This process has been used to weld 2198 Al–Li alloy and has achieved good results. When the appropriate process parameters were selected, grain refinement of the welded joint was obvious [16]. Research on the excellent combination of AC CMT and high-frequency pulse current for welding 2060 Al–Li alloy can provide a reference for the preparation of aerospace structural parts.
2 Experiment
The welding set-up houses two parallel independent power systems (i.e., CMT advanced power supply and high-frequency power supply) that are connected to a welding torch. Therefore, there is a single arc during the welding process that combines the AC CMT with a high-frequency pulse current (Figure 2) [16]. There are diodes in the circuit of the high-frequency power supply in parallel with CMT power supply. The diode has the characteristics of unidirectional current, so two parallel power supplies will not interfere with each other. A photograph of the welding system is presented in Figure 3. This welding system is composed of a KUKA robot and its control cabinet, Austria Fronius CMT Advanced 4000R AC CMT welding machine, and a high-frequency power supply. The peak current of the high-frequency pulse was 30 A, the duty ratio was 50%, and the frequency range selected in the current study was 20–80 kHz. In accordance with the CMT welding database, the optimal welding parameters (i.e., without coupling high-frequency current welding) of the welding wire were chosen (Table 1).
![Figure 2
The experimental set-up of AC CMT advanced supply coupled with high-frequency power supply: (a) AC CMT advanced supply and high-frequency generator and (b) welding torch [16].](/document/doi/10.1515/htmp-2021-0026/asset/graphic/j_htmp-2021-0026_fig_002.jpg)
The experimental set-up of AC CMT advanced supply coupled with high-frequency power supply: (a) AC CMT advanced supply and high-frequency generator and (b) welding torch [16].

Photograph of the welding system.
Process parameters for AC CMT welding
Wire brand | Arc voltage U/V | Welding current I/A | Welding rate v/(cm min−1) | Gas flow rate q/(L min−1) | EP:EN |
---|---|---|---|---|---|
ER5356 | 9.8 | 90 | 70 | 20 | 10:10 |
The base metal was an Al–Li alloy (2060-T8) plate (135 × 95 × 2 mm). Filler wire of ER5356 with a 1.2 mm diameter was used. The Mg content in the ER5356 filler wire can play a vital role in dispersion strengthening due to the precipitation of the β(Al3Mg2) phase in the Al alloy [18]. Therefore, ER5356 filler wires were chosen to weld the 2060 plate. In addition, a comparison of the performance of ER5356- and ER4043-filled 2060 Al–Li alloy welding joints shows that ER5356 welding wire has better performance than ER4043 in reducing pore distribution, and it is not conducive to the generation of pores in terms of wire composition [19]. The chemical compositions of the 2060-T8 Al–Li alloy and filler material ER5356 are shown in Table 2. Photographs of the welded plates are shown in Figure 4.
Composition of the base metal Al–Li 2060 and ER5356 filler wire (wt%)
Cu | Li | Zn | Mn | Mg | Si | Fe | Al | |
---|---|---|---|---|---|---|---|---|
2060 | 4.2 | 0.9 | 0.45 | 0.5 | 1.1 | 0.05 | 0.01 | Bal. |
ER5356 | 0.05 | — | 0.05 | 0.15 | 5.0 | 0.1–0.2 | 0.1–0.2 | Bal. |

Welded plates made with different frequencies: (a) 0 kHz, (b) 40 kHz, (c) 50 kHz, (d) 60 kHz, and (e) 70 kHz.
Unlike conventional welding methods, the AC CMT current and voltage have unique characteristics in both the EP and EN phases [20]. The characteristics of the EP and EN phases of the AC CMT during welding are shown in Figure 5. During the EP phase, the cations hit the surface of the workpiece and molten pool, which can quickly clean the oxide film on the surface of the workpiece. Moreover, the arc is more divergent, the workpiece generates more heat as a cathode, and the temperature of the molten pool is higher. In addition, the transition period is longer than that of the EN phase. During the EN phase, the welding wire as the cathode generates more heat. Therefore, the welding wire melts faster, the transition period is shorter than in the EP phase, and the arc shrinks more. Compared with the EN phase, the arc is divergent during the EP phase, the workpiece produces a large amount of heat, and the arc burns for a longer time, which is not conducive to reducing the heat input. Thus, the EP phase affects the welding quality of the aluminum–lithium alloy [21]. Consequently, the EP phase coupled with a high-frequency pulse current was used in this study to improve the arc characteristics.

The characteristics of EP phase and EN phase of AC CMT: (a) AC CMT current and voltage, (b) arc morphology of different phases of AC CMT, (c) AC CMT current waveform coupled with high-frequency pulse current, and (d) AC CMT voltage waveform coupled with high-frequency pulse current.
The surfaces were carefully prepared before welding. This involved cleaning the oxide films on the butt using a steel wire brush and scraper (thickness at least 0.15 mm) to effectively eliminate the pores. Acetone was used to remove oil and water stains. Flat welding with a set arc guide plate and an arc receiving plate were used, along with high-purity (99.99%) argon gas for shielding purposes. The microstructures of the weld joints were characterized using a Zeiss optical microscope. The samples were subjected to metallographic preparation techniques (i.e., mount, ground, and polished) and later etched using Keller’s reagent (95% H2O + 2.5% HNO3 + 1.5% HCl + 1.0% HF, %vol). Vickers microhardness tests were performed on the cross-sectional area of the weld joints at a load of 200 gf for a dwell time of 15 s. The hardness tester was a THV-1MD automatic turret digital microhardness tester. The samples for tensile testing were wire-cut using a wire-rod electronic discharger in accordance with ISO 4136:2001 tensile standards (Figure 6). The tensile tests were performed at a speed of 1 mm min−1 at room temperature using a SANS CMT5204 electronic universal test machine.

Schematic representation of the tensile sample (all dimensions are in mm).
3 Results and discussions
3.1 Morphological characteristics of arc
During welding, the morphology of the arc while using the filler wire ER5356 was analyzed (Figure 7). With the addition of a high-frequency pulse current, there was a significant drop in the arc length (Figure 8(b)). However, the arc burning period did not change. This could be ascribed to the AC CMT welder control system, which uses the arc voltage signal as the feedback signal. Based on the arc voltage in the control system, the wire feeding system responds by either withdrawing or feeding the wire [22]. For example, when the control system detects that the arc voltage is zero (or below a threshold value), the wire feeding system starts to withdraw the wire. In contrast, when the arc voltage increases above a critical value, the wire feeder begins to feed the wire. Therefore, when a high-frequency pulse current was added, the arc voltage was increased, and the filler wire was drawn back. This resulted in an increase in the dry elongation and a reduction in the arc length. As seen in Figure 8(b), adding the high-frequency pulse initially lowers the arc length at 20 and 30 kHz, then increases it to near the 0 kHz value for higher frequencies, with a maximum of 6.9 mm at 50 kHz. The above phenomenon could be attributed to the following reasons: (a) the arc voltage has been used as a feedback quantity to control wire movement and (b) the frequency of the coupling high-frequency pulse current is much greater than the change frequency of the welding current and voltage. Therefore, the high-frequency current applied to the AC CMT arc was calculated using the root-mean-square (RMS) value. The RMS value I e can be determined using the following equation:
where I e is the effective value of pulse current, I p is the peak pulse current, and D is the pulse current duty cycle.

Arc forms at different frequencies.

Arc characteristics versus frequencies: (a) arc diameter, (b) arc length, and (c) arc volume.
According to equation (1), the RMS value of the pulse current is only related to the peak value of the pulse current I p and duty cycle D, and has very little correlation with the high-frequency current frequency. If, I p and D remain unchanged, the RMS value of the pulse current remains unaffected. If the average impedance and the voltage imposed on the arc are assumed to be constant, then the arc length should remain unchanged with the addition of the high-frequency pulse current (Ohm’s law) [22]. However, the results show that the arc length changes according to the frequency of the high-frequency impulse current. This also indicates that the average arc impedance follows the same trend as that of the arc length, which first decreases and then increases. This arc impedance is determined by the degree of arc ionization (i.e., the number of carriers) [23]. This clearly shows that high-frequency pulse currents of different frequencies have multiple effects on the degree of ionization of the arc. With an increase in the pulse frequency, the degree of arc ionization and density of the conductive ions increased, whereas the arc impedance decreased. The maximum ionization degree was observed for a particular frequency range (50–70 kHz), followed by a sharp drop at higher frequencies [24]. This could be ascribed to the limited excitation response of the arc plasma to the impulse energy. The energy response and transmission of the arc as a heat source are suppressed by energy excitation with very high frequencies. This leads to the failure of the immediate transmission of conductive ions and an increase in the arc impedance. When the welding current changes from the base current to the peak current, the anode current density forms a voltage spike. The leading edge of the pulse voltage also has a spike, and the duration of this high voltage does not change with the frequency. At this time, the arc voltage is in an unsteady state, which can be maintained for a period of time. This shows that there is a transition process when the pulsed arc changes from one stable state to another. The time of this transition does not change with frequency; it is a characteristic of the arc itself. Because the arc can be controlled in an unstable state at high frequency, the stiffness of the arc, the energy density, and the power all increase, which is the cause of the high-frequency effect [24].
3.2 Porosity in weld joints
Hydrogen pores are one of the main problems in the welding of Al–Li alloys. Al–Li alloys are rich in active metallic elements such as Mg and Li, which promote the formation of oxidation layers after high-temperature processing. These oxide layers absorb moisture from the surrounding environment during welding. Consequently, the hydrogen atoms can enter the molten pool, leading to the formation of hydrogen pores because hydrogen solubility is higher in the liquid state than in the solid. The presence of pores can be detrimental to the strength and quality of weld joints. Therefore, it is necessary to investigate the relationship between weld material porosity and the corresponding welding parameters. Welding samples coupled with high-frequency pulse current were chosen for the porosity investigation as they had a better appearance with no weld defects, no obvious deformation of plates, and clear and continuous fish scales.
Figure 9 shows the occurrence of porosity in the Al–Li weld joints. In the absence of a coupled high-frequency pulse, the pores are not evenly distributed. For example, pores of different sizes are observed in the center, fusion line, and edge of the weld (Figure 9(a)). When coupled with a high-frequency current, the number of pores in the center of the weld is comparatively small, and most of the pores are concentrated along the fusion line and the edge of the weld. During the welding process, a high-frequency electrical pulse flows through the molten pool formed by a pulsation-induced electromagnetic field of a certain intensity. Under the action of a pulsed induction electromagnetic field, the molten pool is affected by a strongly pulsed electromagnetic force. For a given cross-section of the hemispherical molten pool, there is a circulating flow driven by an electromagnetic force that originates at the center of the surface and then flows downward. The pores in the molten pool are effectively concentrated at the edge of the liquid metal hemisphere by stirring and then move away. Hence, the pores were distributed around the edge of the weld and along its centerline during the solidification process. The above phenomenon showing the flow direction of the liquid metal under stress in the molten pool is shown in Figure 10 [25,26].

Comparative study on the porosity level and microstructure at weld joints made with different frequencies: (a and f) 0 kHz, (b and g) 40 kHz, (c and h) 50 kHz, (d and i) 60 kHz, and (e and j) 70 kHz.

The flow direction of molten metal under stress.
The fewest pores were seen in the weld joints at a coupling frequency of 70 kHz. The above results clearly show that the porosity of the welded Al–Li alloy was reduced after the introduction of high-frequency pulse currents.
3.3 Microstructure
The microstructures of the different weld areas and the corresponding effect of high-frequency on grain morphologies are shown in Figure 9(f–j). Han et al. [27] studied LBW in an Al–Li alloy and showed that the microstructure near the fusion line can be divided into several zones, that is, the fusion zone (FZ), equiaxed grain zone (EQZ), partially melted zone (PMZ), and heat-affected zone (HAZ). The grain structure at the weld center without high-frequency pulse current was coarse and equiaxed (Figure 9(f)). The weld microstructure displayed significant grain refinement by using high-frequency pulse currents during the welding process (Figure 9(g–j)). Moreover, there were very few coarse grains, and there was a marked increase in grain refinement for the coupling frequency of 70 kHz (Figure 9(j)). In summary, the weld microstructures displayed grain refinement with an increase in the coupling frequency.
Based on the above observations, an in-depth analysis of the rationale behind the grain refinement was carried out. The original CMT welding current, along with the coupling of the high-frequency pulse current, changes the induced pulsating magnetic field. According to Faraday’s electromagnetic induction law, the molten metal at the weld joint produces an induced current, J e. Owing to the electromagnetic induction, J e generates a Lorentz force F under the action of a magnetic field, B. The Lorentz force can be expressed by equations (2)–(4) [28].
According to Maxwell’s equation,
Therefore, it can be inferred that
where
In the molten pool, the current flow direction and current density were not completely consistent and uniform. The current density at the weld center was high, whereas the density on the sides was low. Therefore, a pressure difference was formed from the center to the sides owing to the electromagnetic force created by the pulsating magnetic field. In addition, this pressure difference caused forced convection on the molten metal inside the weld. This can flush down some undissolved intermetallic particles (for example, Al3Zr and Al3Ti) with high melting points in the FZ and bring them into the pool [29]. Therefore, the number of non-uniform nucleation spots in the weld pool was significantly increased, which could further help in the grain refinement process. Moreover, these forces can produce a shearing effect on the growing dendrites, causing the dendrites to be fractured and dispersed in the weld pool. When the weld metal begins to solidify, the broken dendrites in the weld pool increase the number of crystal nuclei, resulting in the transformation of columnar crystals into equiaxed crystals [30,31].
The current intensity and induced magnetic induction intensity vary with the position in the melt [32]. The contraction force gradient and the velocity difference between the flow groups are formed in the melt, resulting in shear stresses that are given by the following equations [33]:
and
where
Combined with the formula given earlier, it can be seen that
The microscopic analysis clearly shows that the high-frequency pulse current significantly refined the microstructure. The 2060 + ER5356 weld combination yielded grain sizes of 36.6, 28.7, 26.8, 25.9, and 24.1 μm for the corresponding coupling current frequencies of 0, 40, 50, 60, and 70 kHz, respectively. In addition, there may be some variation in the frequency range at which the high-frequency begins to have a significant effect because of the filler wire properties (such as electrical conductivity, viscosity of the molten flow and surface tension).
3.4 Effect of welding parameters on the mechanical properties of weld joints
3.4.1 Microhardness
Microhardness patterns are measured to help understand the mechanical properties of welded joints, which depend on the microstructure [34]. The microhardness measurements are shown in Figure 11. The base metal (BM) was found to have an average microhardness of 146 ± 5 HV, which then drops in the HAZ and is further reduced in the weld zone (WZ). The microhardness pattern across the weld joints is consistent for all the coupling frequencies. The average microhardness is almost the same (70 ± 5 HV) across the center of the weld (Figure 11) for each condition. When the coupling frequency was 70 kHz, the average microhardness of the fusion line was the highest 118 ± 3 HV (i.e., ∼74% of the base material). Based on these results, the highest hardness corresponded to the frequency level with the finest grain microstructure. A fine-grained microstructure resulted in improved mechanical properties (e.g., yield strength and hardness). This is in conjunction with the Hall–Petch equation:

Hardness distribution on the weld joints with ER5356.
In general, microhardness is affected by the grain size and the presence of other metallurgical phases. The hardness at the weld center is slightly higher than that of the fusion line (Figure 11). This can be ascribed to the finer grains at the weld center, which were formed by the coupling of high frequencies in the welding process. Moreover, a decrease in grain size leads to an increase in the number of grain boundaries, which are high-energy sites that act as strong barriers against crack propagation [37].
Moreover, the weld zone is the region with the highest temperature during welding. Therefore, intermetallic particles (2060 + ER5356: β(Mg2Al3), θ(CuAl2), S(MgCuAl2)) would have gathered, grown, and dissolved in the weld zone under the action of the high temperature. Despite significant grain refinement in the weld zone, the hardness value was still lower than that of the BM. The hardness in HAZ was reduced owing to over-aging, dissolution, and growth of the enhanced phase under the action of a high temperature. The heat transferred from the weld area to HAZ is relatively small because of the lower heat input of CMT. Therefore, the second-phase particles in HAZ aggregated and dissolved, but the dissolution was not as significant as in the weld zone. In addition, the BM is in a T8 state (i.e., cold reduction after solution treatment and artificial aging). Therefore, HAZ has a higher hardness than the weld zone but smaller than that of the BM.
3.4.2 Tensile properties of the weld joints
The engineering stress–strain curves and tensile strength of welded joints at different coupling frequencies are presented in Figure 12. The tensile strength of the base material was 490 MPa, and the efficiency of the welded joint was calculated using the following equation [38]:

Stress–strain curves and tensile strength of weld joints made with different frequencies: (a) 0 kHz, (b) 40 kHz, (c) 50 kHz, (d) 60 kHz, (e) 70 kHz, and (f) tensile strength.
Without coupling with the high-frequency pulse current, the tensile strength of the weld joint was 230 MPa, and the minimum joint efficiency was calculated to be approximately 46.9%. However, the tensile strength of the weld joints was improved after coupling with the high-frequency pulse current. A frequency of 70 kHz resulted in a superior tensile strength of 249 MPa, with a maximum joint efficiency of about 50.8%. This clearly shows an increase of 19 MPa in the tensile strength of the 70 kHz weld joint above that made without high-frequency pulse current.
Tensile strength is related to the mobility of dislocations in the welded joint. As shown in Figure 9, the grain size of the welded joint was significantly reduced under the coupled high-frequency pulse. The associated grain boundaries have high energy owing to their structural disorder. Therefore, mobile dislocations must possess high amounts of energy to cross the grain boundaries. This is why mobile dislocations accumulate behind the grain boundaries when they reach them and stop moving. This results in an increase in the dislocation density behind the grain boundaries and finally increases the tensile strength [37].
The weld joint fracture surface characteristics were also studied (Figure 13). From the fracture analysis, it was clear that the fusion line was the weakest link, as all the fracture locations were near the fusion line. From the macroscopic fracture morphology shown in Figure 13(b), there is a slight necking phenomenon (plastic deformation) at the fracture, and the fracture surface was relatively flat and perpendicular to the direction of the tensile load. From the SEM fractographs (Figure 14), the fracture surface shows obvious dimple features and a few tear ridges (Figure 14(a–e)), which indicates a mainly ductile fracture mode. A small amount of tear ridges indicates that the fracture has a tendency for quasi-cleavage fracture [39–41]. Most metals and alloys (especially those with FCC lattice, such as Al–Li alloy) exhibit ductile fracture at ambient temperature. The ductile fracture characteristics under tensile stress include the development of very fine cavities within the necking region, which subsequently connect to form a fine crack that grows to final fracture [42,43]. It can be confirmed that the 70 kHz sample would be more ductile (with more dimples) because of its smaller grain size [44,45]. The coupling high-frequency pulse current had very little effect of on the fracture modes of the welding materials.

Samples after fracture: (a) tensile samples and (b) magnified view of fractures.

SEM fractographs of weld joints made with different frequencies: (a and a1) 0 kHz, (b and b1) 40 kHz, (c and c1) 50 kHz, (d and d1) 60 kHz, and (e and e1) 70 kHz.
4 Conclusion
In general, the arc shape was largely based on the pulse frequencies. In addition, the arc diameter, arc length, and arc volume show a trend of rising first and then falling with the increase of the pulse frequency and reached their peak values when the pulse frequency was 50 kHz. The peak arc diameter, arc length, and arc volume were 14.5 mm, 6.9 mm, and 1,134 mm3, respectively.
In the process of welding 2060 Al–Li alloy with AC CMT coupled high-frequency pulse current, the high-frequency pulse current had a significant effect on the porosity, grain size, and tensile strength of the weld joints.
The porosity levels and grain size at the weld joints varied depending on the pulse frequencies employed for welding. The porosity levels and grain size both decreased with increasing frequency. When the pulse frequency was 70 kHz, the porosity level was the lowest, and the grain size was refined to 24.1 μm.
When using filler wire ER5356 along with the high-frequency effect, there was a significant increase in the tensile strength of the weld joints. For example, a maximum tensile strength of 249 MPa was observed for the weld combination of ER5356 and 70 kHz.
-
Funding information: This work was supported by the National Natural Science Foundation of China (Grant No. 51875168), Key Research and Development Program of Hebei Province (Grant No. 19212108D/19250202D/20312205D), and Natural Science Foundation of Hebei Province (Grant No. E2019208089/E2020208089/E2020208083).
-
Authors contribution: Liwei Wang: Supervision; Huan Hu: Writing-Original draft preparation; Huan Yan: Data curation, Validation; Ying Liu: Formal analysis; Ziqin Wu: Data curation, Validation; Balaji Narayanaswamy: Writing-Review & Editing; Zhimin Liang: Conceptualization; Dianlong Wang: Conceptualization.
-
Conflict of interest: Authors state no conflict of interest.
References
[1] Williams, J. C. and E. A. Starke. Progress in structural materials for aerospace systems 1. Acta Materialia, Vol. 51, No. 19, 2003, pp. 5775–5599.10.1016/j.actamat.2003.08.023Search in Google Scholar
[2] Rioja, R. J. and J. Liu. The evolution of Al–Li base products for aerospace and space applications. Metallurgical and Materials Transactions. A, Vol. 43, No. 9, 2012, pp. 3325–3337.10.1007/s11661-012-1155-zSearch in Google Scholar
[3] Zhao, T. B., S. S. Yutaka, R. S. Xiao, T. Huang, and J. Q. Zhang. Laser pressure welding of Al–Li alloy 2198: effect of welding parameters on fusion zone characteristics associated with mechanical properties. High Temperature Materials and Processes, Vol. 39, 2020, pp. 146–156.10.1515/htmp-2020-0047Search in Google Scholar
[4] Zhang, X. Y., X. S. Wang, X. Ye, and L. J. Guo. Study on influencing factors of porosity of 6061 aluminum alloy by CMT welding. Hot Working Technology, Vol. 48, No. 19, 2019, pp. 43–47.Search in Google Scholar
[5] Mousavi, S. H. and M. Sabzi. The effect of pouring temperature and surface angle of vortex casting on microstructural changes and mechanical properties of 7050Al-3 wt% SiC composite. Materials Science and Engineering A, Vol. 737, 2018, pp. 230–235.10.1016/j.msea.2018.09.057Search in Google Scholar
[6] Wei, S. T., C. Y. Hao, and J. C. Chen. Study of friction stir welding of 1420 aluminum–lithium alloy. Materials Science and Engineering. A, Vol. 452, 2007, pp. 170–177.10.1016/j.msea.2006.10.081Search in Google Scholar
[7] Cai, B., Z. Q. Zheng, D. Q. He, S. C. Li, and H. P. Li. Friction stir weld of 2060 Al–Cu–Li alloy: microstructure and mechanical properties. Journal of Alloys and Compounds, Vol. 649, 2015, pp. 19–27.10.1016/j.jallcom.2015.02.124Search in Google Scholar
[8] Chang, C. C., C. P. Chou, S. N. Hsu, G. Y. Hsiung, and J. R. Chen. Effect of laser welding on properties of dissimilar joint of Al–Mg–Si and Al–Mn aluminum alloys. Journal of Materials Science and Technology, Vol. 26, No. 3, 2010, pp. 276–282.10.1016/S1005-0302(10)60046-1Search in Google Scholar
[9] Liu, F., B. Zhou, Y. Mao, C. Huang, Y. Chen, and Z. Wang. Microstructure and mechanical properties of laser welded joints between 2198/2060 Al–Li alloys. Materials Science and Technology, Vol. 34, No. 1, 2018, pp. 111–122.10.1080/02670836.2017.1365221Search in Google Scholar
[10] Liu, T., Y. Q. Zhao, K. Yue, and X. H. Zhan. Effect of micro morphology in different zones on mechanical properties of 2060 Al–Li alloy laser welded joints. Journal of Manufacturing Processes, Vol. 50, 2020, pp. 336–344.10.1016/j.jmapro.2019.12.032Search in Google Scholar
[11] Meng, X., Y. Huang, J. Cao, J. Shen, and J. F. Santos. Recent progress on control strategies for inherent issues in friction stir welding. Progress in Materials Science, Vol. 115, 2021, id. 100706.10.1016/j.pmatsci.2020.100706Search in Google Scholar
[12] Andres, J., A. Wrońska, T. Gałaczyński, G. Luty, and R. Burek. Effect of process parameters on microstructure and mechanical properties of RFSSW lap joints of thin Al7075-T6 sheets. Archives of Metallurgy and Materials, Vol. 63, No. 6, 2018, pp. 39–43.10.24425/118906Search in Google Scholar
[13] Wang, L. W., Y. C. Suo, C. F. Wu, D. L. Wang, Z. M. Liang, and S. Q. Dong. Effect of high-frequency pulse current coupled with DC TIG on microstructure and hardness of 6N01 aluminum alloy welded joints. Journal of Shenyang University (Natural Science), Vol. 31, No. 3, 2019, pp. 173–177.Search in Google Scholar
[14] Hao, H., G. Q. Lu, F. Y. Wang, and X. G. Zhang. Effects of electromagnetic stirring on solidification of metal foams. Journal of Iron and Steel Research(International), Vol. 19, No. Part S1, 2012, pp. 195–198.Search in Google Scholar
[15] Kumar, R., U. Dilthey, D. K. Dwivedi, and P. K. Ghosh. Thin sheet welding of Al 6082 alloy by AC pulse-GMA and AC wave pulse-GMA welding. Materials and Design, Vol. 30, No. 2, 2008, pp. 306–313.10.1016/j.matdes.2008.04.073Search in Google Scholar
[16] Wang, D. L., C. F. Wu, Y. C. Suo, L. W. Wang, and Z. M. Liang. Effect of pulse frequency on arc behavior and droplet transfer of 2198 Al–Li alloy by ultrahigh-frequency pulse AC CMT welding. Journal of Materials Research and Technology, Vol. 8, No. 5, 2019, pp. 3950–3958.10.1016/j.jmrt.2019.07.003Search in Google Scholar
[17] Feng, J. C., H. T. Zhang, and P. He. The CMT short-circuiting metal transfer process and its use in thin aluminum sheets welding. Materials and Design, Vol. 30, No. 5, 2009, pp. 1850–1852.10.1016/j.matdes.2008.07.015Search in Google Scholar
[18] Qi, Z. W., B. Q. Cong, B. J. Qi, H. Y. Sun, G. Zhao, and J. L. Ding. Microstructure and mechanical properties of double-wire + arc additively manufactured Al–Cu–Mg alloys. Journal of Materials Processing Technology, Vol. 255, 2017, pp. 347–353.10.1016/j.jmatprotec.2017.12.019Search in Google Scholar
[19] Liu, S. Q. Effect of AC CMT compound high frequency pulse on the structure and properties of aluminum-lithium alloy welded joints. Master’s thesis. Hebei University of Science and Technology, Hebei, 2015.Search in Google Scholar
[20] Zhang, B. W. and L. Q. Zhang. Research progress of AC cold metal transfer technology (Advanced CMT) and its application in additive manufacturing. The Journal of New Industrialization, Vol. 7, No. 11, 2017, pp. 82–88.Search in Google Scholar
[21] Park, H. J., S. Rhee, M. J. Kang, and D. C. Kim. Joining of steel to aluminum alloy by AC pulse MIG welding. Materials Transactions, Vol. 50, No. 9, 2009, pp. 2314–2317.10.2320/matertrans.M2009105Search in Google Scholar
[22] Ying, S. Y. Basic and application of gas shielded welding process, China Machine Press, Beijing, 2012.Search in Google Scholar
[23] Song, Y. L., S. B. Yan, T. J. Xiao, X. H. Yang. A study on the macro-micro physical properties in pulsed arc plasma. Transactions of Japanese Welding Research Institute, Vol. 39, No. 2, 2010, pp. 17–18.Search in Google Scholar
[24] Ma, M. Research on high frequency gas shielded welding equipment and arc behavior. Master’s thesis. Beijing University of Technology, Beijing, 2015.Search in Google Scholar
[25] Yang, M. X., Z. Yang, B. Q. Cong, Q. Wang, and B. J. Qi. Metal molten behavior with ultrahigh frequency pulsed arc welding. Journal of Beijing University of Aeronautics and Astronautics, Vol. 41, No. 4, 2015, pp. 589–593.Search in Google Scholar
[26] Nguyen, A. V., S. Tashiro, M. H. Ngo, A. H. Le, H. V. Bui, and M. Tanaka. Influence of shielding gas composition on molten metal flow behavior during plasma keyhole arc welding process. Journal of Manufacturing Processes, Vol. 53, 2020, pp. 431–437.10.1016/j.jmapro.2020.03.031Search in Google Scholar
[27] Han, B., Y. B. Chen, T. Wang, H. Li, and L. Q. Li. Microstructural evolution and interfacial crack corrosion behavior of double-sided laser beam welded 2060/2099 Al–Li alloys T-joints. Materials and Design, Vol. 135, 2017, pp. 353–365.10.1016/j.matdes.2017.09.042Search in Google Scholar
[28] Zou, W. Q., Z. G. Zhang, H. Yang, and W. Li. Effect of vibration frequency on microstructure and performance of high chromium cast iron prepared by lost foam casting. China Foundry, Vol. 13, No. 4, 2016, pp. 248–255.10.1007/s41230-016-6037-3Search in Google Scholar
[29] Cross, C. E., Ø. Grong, and M. Mousavi. A model for equiaxed grain formation along the weld metal fusion line. Scripta Materialia, Vol. 40, No. 10, 1999, pp. 1139–1144.10.1016/S1359-6462(99)00023-8Search in Google Scholar
[30] Zhang, B. J., J. Z. Cui, and G. M. Lu. Effects of low-frequency electromagnetic field on microstructures and macro segregation of continuous casting 7075 aluminum alloy. Materials Science and Engineering A, Vol. 355, No. 1, 2003, pp. 325–330.10.1016/S0921-5093(03)00105-9Search in Google Scholar
[31] Zhang, B. J., J. Z. Cui, and G. M. Lu. Effect of low-frequency magnetic field on macrosegregation of continuous casting aluminum alloys. Materials Letters, Vol. 57, No. 11, 2003, pp. 1707–1711.10.1016/S0167-577X(02)01055-8Search in Google Scholar
[32] Nakada, M., Y. Shiohara, and M. C. Flemings. Modification of solidification structures by pulse electric discharging. ISIJ International, Vol. 30, No. 1, 1990, pp. 27–33.10.2355/isijinternational.30.27Search in Google Scholar
[33] Li, J. Mechanism and application of pulse current solidification technology. Doctoral dissertation. Shanghai University, Shanghai, 2009.Search in Google Scholar
[34] Mousavi, S. H., M. Sabzib, M. Ghobeiti-Hasabc, and A. Roshan-Ghiyasd. Optimization of spot welding process parameters in dissimilar joint of dual phase steel DP600 and AISI 304 stainless steel to achieve the highest level of shear-tensile strength. Materials Science and Engineering A, Vol. 726, 2018, pp. 120–125.10.1016/j.msea.2018.04.072Search in Google Scholar
[35] Hall, E. O. The deformation and ageing of mild steel: Ⅲ discussion of results. Proceedings of the Physical Society. B, Vol. 64, No. 9, 1951, pp. 747–753.10.1088/0370-1301/64/9/303Search in Google Scholar
[36] Petch, N. J. The cleavage strength of polycrystals. Journal of the Iron and Steel Institute, Vol. 174, 1953, pp. 25–28.Search in Google Scholar
[37] Sabzi, M. and S. M. Dezfuli. Drastic improvement in mechanical properties and weldability of 316L stainless steel weld joints by using electromagnetic vibration during GTAW process. Journal of Manufacturing Processes, Vol. 33, 2018, pp. 74–85.10.1016/j.jmapro.2018.05.002Search in Google Scholar
[38] Pandey, C., M. M. Mahapatra, P. Kumar, F. Daniel, and B. Adhithan. Softening mechanism of P91 steel weldments using heat treatments. Archives of Civil and Mechanical Engineering, Vol. 19, No. 2, 2019, pp. 297–310.10.1016/j.acme.2018.10.005Search in Google Scholar
[39] Pandey, C., M. M. Mahapatra, P. Kumar, N. Saini, and J. G. Thakre. Study on effect of double austenitization treatment on fracture morphology tensile tested nuclear grade P92 steel. Engineering Failure Analysis, Vol. 96, 2018, pp. 158–167.10.1016/j.engfailanal.2018.09.036Search in Google Scholar
[40] Saini, N., C. Pandey, M. M. Mahapatra, and R. S. Mulik. On study of effect of varying tempering temperature and notch geometry on fracture surface morphology of P911 (9Cr-1Mo-1W-V-Nb) steel. Engineering Failure Analysis, Vol. 85, 2018, pp. 104–115.10.1016/j.engfailanal.2017.12.013Search in Google Scholar
[41] Pandey, C. C., N. Pandey, M. M. Saini, and P. K. Mahapatra. Study of the fracture surface morphology of impact and tensile tested cast and forged (C&F) Grade 91 steel at room temperature for different heat treatment regimes. Engineering Failure Analysis, Vol. 71, 2017, pp. 131–147.10.1016/j.engfailanal.2016.06.012Search in Google Scholar
[42] Sabzi, M., A. Obeydavi, and S. H. Mousavi. The effect of joint shape geometry on the microstructural evolution, fracture toughness, and corrosion behavior of the welded joints of a hadfield steel. Mechanics of Advanced Materials and Structures, Vol. 26, No. 12, 2019, pp. 1053–1063.10.1080/15376494.2018.1430268Search in Google Scholar
[43] Sabzi, M. and S. M. Dezfuli. Post weld heat treatment of hypereutectoid hadfield steel: Characterization and control of microstructure, phase equilibrium, mechanical properties and fracture mode of welding joint. Journal of Manufacturing Processes, Vol. 34, 2018, pp. 313–328.10.1016/j.jmapro.2018.06.009Search in Google Scholar
[44] Jafarian, H. R., M. Sabzi, S. H. Mousavi, A. R. Eivani, and N. Park. The influence of austenitization temperature on microstructural developments, mechanical properties, fracture mode and wear mechanism of Hadfield high manganese steel. Journal of Materials Research and Technology, Vol. 10, 2021, pp. 819–831.10.1016/j.jmrt.2020.12.003Search in Google Scholar
[45] Mousavi, S. H. and M. Sabzi. The effect of heat treatment process parameters on mechanical properties, precipitation, fatigue life, and fracture mode of an austenitic Mn hadfield steel. Journal of Materials Engineering and Performance, Vol. 27, No. 10, 2018, pp. 5246–5253.10.1007/s11665-018-3625-ySearch in Google Scholar
© 2021 Liwei Wang et al., published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Research Articles
- Fused deposition modeling of poly(ether ether ketone) scaffolds
- Investigation of the microstructure evolution in TP347HFG austenitic steel at 700°C and its characterization method
- Hot deformation behavior and processing maps of 9Cr3W3Co oxide dispersion-strengthened steel
- Evolution of physicochemical properties of quick lime at converter-smelting temperature
- Influence of phase distribution of converter slag microzones on the occurrence of P
- Investigation on ultrasonic assisted friction stir welding of aluminum/steel dissimilar alloys
- Analysis of oxide scale thickness and pores position of HCM12A steel in supercritical water
- Behavior of MnS inclusions during homogenization process in low-alloyed steel FAS3420H
- Preparation and cutting performance of nano-scaled Al2O3-coated micro-textured cutting tool prepared by atomic layer deposition
- Prediction of hot metal temperature based on data mining
- Effect of TiO2 content in slag on Ti content in molten steel
- Performance evaluation of titanium-based metal nitride coatings and die lifetime prediction in a cold extrusion process
- Effect of different drilling techniques on high-cycle fatigue behavior of nickel-based single-crystal superalloy with film cooling hole
- Effect of CO2 injection into blast furnace tuyeres on the pulverized coal combustion
- Microstructure and properties of Co–Al porous intermetallics fabricated by thermal explosion reaction
- Evolution regularity of temperature field of active heat insulation roadway considering thermal insulation spraying and grouting: A case study of Zhujidong Coal Mine, China
- Evolution of reduction process from tungsten oxide to ultrafine tungsten powder via hydrogen
- A thermodynamic assessment of precipitation, growth, and control of MnS inclusion in U75V heavy rail steel
- Effect of basicity on the reduction swelling properties of iron ore briquettes
- Effect of Cr and Al alloying on the oxidation resistance of a Ti5Si3-incorporated MoSiBTiC alloy
- Microstructure and mechanical properties of 2060 Al–Li alloy welded by alternating current cold metal transfer with high-frequency pulse current
- Effects of composition and strain rate on hot ductility of Cr–Mo-alloy steel in the two-phase region
- Effect of K and Na on reduction swelling performance of oxidized roasted briquettes
- Dephosphorization mechanism and phase change in the reduction of converter slag
- Parametric investigation and optimization for CO2 laser cladding of AlFeCoCrNiCu powder on AISI 316
- Optimization of heat transfer and pressure drop of the channel flow with baffle
- Quantitative analysis of microstructure and mechanical properties of Nb–V microalloyed high-strength seismic reinforcement with different Nb additions
- Visualization of the damage evolution for Ti–3Al–2Mo–2Zr alloy during a uniaxial tensile process using a microvoids proliferation damage model
- Research on high-temperature mechanical properties of wellhead and downhole tool steel in offshore multi-round thermal recovery
- Dephosphorization behavior of reduced iron and the properties of high-P-containing slag
- Jet characteristics of CO2–O2 mixed injection using a dual-parameter oxygen lance nozzle for different smelting periods
- Effects of ball milling on powder particle boundaries and properties of ODS copper
- Heat transfer behavior in ultrahigh-speed continuous casting mold
- Solidification microstructure characteristics of Cu–Pb alloy by ECP treatment
- Luminescence properties of Eu2+ and Sm3+ co-doped in KBaPO4
- Research on high-temperature oxidation resistance, hot forming ability, and microstructure of Al–Si–Cu coating for 22MnB5 steel
- The differential analysis for temperature distribution diagnostics of arc current-carrying region in sheet slanting tungsten electrode inert gas welding with the electrostatic probe
- Reactions at the molten flux-weld pool interface in submerged arc welding
- The effect of liquid crystalline graphene oxide compared with non-liquid crystalline graphene oxide on the rheological properties of polyacrylonitrile solution
- Study on manganese volatilization behavior of Fe–Mn–C–Al twinning-induced plasticity steel
- Physical modeling of bubble behaviors in molten steel under high pressure
- Rapid Communication
- The new concept of thermal barrier coatings with Pt + Pd/Zr/Hf-modified aluminide bond coat and ceramic layer formed by PS-PVD method
- Topical Issue on Science and Technology of Solar Energy
- Solution growth of chalcopyrite Cu(In1−xGax)Se2 single crystals for high open-circuit voltage photovoltaic device
- Copper-based kesterite thin films for photoelectrochemical water splitting
Articles in the same Issue
- Research Articles
- Fused deposition modeling of poly(ether ether ketone) scaffolds
- Investigation of the microstructure evolution in TP347HFG austenitic steel at 700°C and its characterization method
- Hot deformation behavior and processing maps of 9Cr3W3Co oxide dispersion-strengthened steel
- Evolution of physicochemical properties of quick lime at converter-smelting temperature
- Influence of phase distribution of converter slag microzones on the occurrence of P
- Investigation on ultrasonic assisted friction stir welding of aluminum/steel dissimilar alloys
- Analysis of oxide scale thickness and pores position of HCM12A steel in supercritical water
- Behavior of MnS inclusions during homogenization process in low-alloyed steel FAS3420H
- Preparation and cutting performance of nano-scaled Al2O3-coated micro-textured cutting tool prepared by atomic layer deposition
- Prediction of hot metal temperature based on data mining
- Effect of TiO2 content in slag on Ti content in molten steel
- Performance evaluation of titanium-based metal nitride coatings and die lifetime prediction in a cold extrusion process
- Effect of different drilling techniques on high-cycle fatigue behavior of nickel-based single-crystal superalloy with film cooling hole
- Effect of CO2 injection into blast furnace tuyeres on the pulverized coal combustion
- Microstructure and properties of Co–Al porous intermetallics fabricated by thermal explosion reaction
- Evolution regularity of temperature field of active heat insulation roadway considering thermal insulation spraying and grouting: A case study of Zhujidong Coal Mine, China
- Evolution of reduction process from tungsten oxide to ultrafine tungsten powder via hydrogen
- A thermodynamic assessment of precipitation, growth, and control of MnS inclusion in U75V heavy rail steel
- Effect of basicity on the reduction swelling properties of iron ore briquettes
- Effect of Cr and Al alloying on the oxidation resistance of a Ti5Si3-incorporated MoSiBTiC alloy
- Microstructure and mechanical properties of 2060 Al–Li alloy welded by alternating current cold metal transfer with high-frequency pulse current
- Effects of composition and strain rate on hot ductility of Cr–Mo-alloy steel in the two-phase region
- Effect of K and Na on reduction swelling performance of oxidized roasted briquettes
- Dephosphorization mechanism and phase change in the reduction of converter slag
- Parametric investigation and optimization for CO2 laser cladding of AlFeCoCrNiCu powder on AISI 316
- Optimization of heat transfer and pressure drop of the channel flow with baffle
- Quantitative analysis of microstructure and mechanical properties of Nb–V microalloyed high-strength seismic reinforcement with different Nb additions
- Visualization of the damage evolution for Ti–3Al–2Mo–2Zr alloy during a uniaxial tensile process using a microvoids proliferation damage model
- Research on high-temperature mechanical properties of wellhead and downhole tool steel in offshore multi-round thermal recovery
- Dephosphorization behavior of reduced iron and the properties of high-P-containing slag
- Jet characteristics of CO2–O2 mixed injection using a dual-parameter oxygen lance nozzle for different smelting periods
- Effects of ball milling on powder particle boundaries and properties of ODS copper
- Heat transfer behavior in ultrahigh-speed continuous casting mold
- Solidification microstructure characteristics of Cu–Pb alloy by ECP treatment
- Luminescence properties of Eu2+ and Sm3+ co-doped in KBaPO4
- Research on high-temperature oxidation resistance, hot forming ability, and microstructure of Al–Si–Cu coating for 22MnB5 steel
- The differential analysis for temperature distribution diagnostics of arc current-carrying region in sheet slanting tungsten electrode inert gas welding with the electrostatic probe
- Reactions at the molten flux-weld pool interface in submerged arc welding
- The effect of liquid crystalline graphene oxide compared with non-liquid crystalline graphene oxide on the rheological properties of polyacrylonitrile solution
- Study on manganese volatilization behavior of Fe–Mn–C–Al twinning-induced plasticity steel
- Physical modeling of bubble behaviors in molten steel under high pressure
- Rapid Communication
- The new concept of thermal barrier coatings with Pt + Pd/Zr/Hf-modified aluminide bond coat and ceramic layer formed by PS-PVD method
- Topical Issue on Science and Technology of Solar Energy
- Solution growth of chalcopyrite Cu(In1−xGax)Se2 single crystals for high open-circuit voltage photovoltaic device
- Copper-based kesterite thin films for photoelectrochemical water splitting