Next Article in Journal
Effect of Magnetic Field on Maskless Localized Electrodepositing Three-Dimensional Microstructure of Nano Nickel Crystals
Next Article in Special Issue
Stress Concentration Factors for Non-Load-Carrying Welded Cruciform Joints Subjected to Tension, Bending, and Shear
Previous Article in Journal
Effects of Adsorption and Desorption of Low-Boiling-Point Total Hydrocarbon Gases on Activated Carbon
Previous Article in Special Issue
The Mutual Influence of Thermal Contact Conductivity and Convective Cooling on the Temperature Field in a Tribosystem with a Functionally Graded Strip
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Residual Stress and Microstructure on the Fatigue Crack Growth Behavior of Aluminum Friction Stir Welded Joints

1
School of Civil Engineering, Xi’an University of Architecture and Technology, Xi’an 710055, China
2
School of Aeronautics, Northwestern Polytechnical University, Xi’an 710072, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(2), 385; https://doi.org/10.3390/ma17020385
Submission received: 18 December 2023 / Revised: 4 January 2024 / Accepted: 10 January 2024 / Published: 12 January 2024

Abstract

:
Friction stir welding (FSW) has been adopted in the aerospace industry for fabricating structural alloys due to the low melting point and high thermal conductivity of aviation aluminum alloys. However, welding residual stresses can lead to secondary deformation in friction stir welded (FSWed) structures. Additionally, microstructural characteristics impact the crack growth rates and directions in these structures. Therefore, it is necessary to investigate the effects of residual stress and microstructure on the fatigue responses of FSWed joints. In this paper, we studied the fatigue crack growth behavior of two homogeneous and dissimilar FSWed joints with varying welding parameters, namely 2024-T3 and 7075-T6. The residual stresses were measured with the X-ray diffraction method. The dislocations and precipitates in different zones of the FSWed joints were analyzed via transmission electron microscopy (TEM). The results demonstrated that the residual stress significantly affected the fatigue crack growth rate and direction; the tensile residual stress promoted fatigue crack growth and offset the decrease in the fatigue crack growth rate that occurred due to grain refinement. The results of the microstructural analysis indicated that dislocation density and sliding resistance increased with the decrease in rotational speed and led to a decreased rate of fatigue crack propagation.

1. Introduction

AA 2xxx and AA 7xxx aluminum alloys are widely used in aircraft structures due to their low density, high strength, and high stiffness [1]. Specifically, the 2024-T3 alloy is primarily employed in the fuselage for the rivets, skin, wings, skeleton, etc. The 7075-T6 alloy serves as the essential material in wing spars, baffle frames, and wing panels. However, traditional welding methods cause several defects in the 2024-T3 and 7075-T6 alloys [2]. On the other hand, friction stir welding (FSW) stands out as a typical solid-state joining technology, which is capable of fabricating high-quality welded joints with minimal welding deformation [3,4]. Therefore, FSW is being increasingly adopted in industries such as aerospace, shipping, railway, electronics, and others [5,6].
During the welding process, changes in local temperature and deformation induce plastic deformation, which occurs in the welded zone (WZ) and thereby leads to the formation of residual stress, which is a self-balancing system that is not dependent on external loads [3,7]. The residual stress and the external loads can result in secondary deformations, as well as the redistribution of residual stress, which significantly influences the integrity and safety of welded structures [8,9,10]. Previous studies have shown that the maximum residual stress in the longitudinal direction (LD) is generated in the heat-affected zone (HAZ), while the minimum compressive residual stress originates on the advancing side just beyond the WZ [11,12]. Furthermore, another study indicated that in the 2024/7075 dissimilar welded joints, the longitudinal residual stress on the 7075 side was higher than that on the 2024 side [13]. This is related to the position within the welded joint, as the heat input on the advancing side is relatively higher than that on the retreating side. Additionally, compared to the 2024 alloy, 7075 aluminum exhibits higher mechanical properties, including yield strength and hardness.
Due to the strong local thermal coupling effect in the FSW process, the microstructure of the welded joint becomes diversified and non-uniform [14,15]. The WZ, which is composed of fine equiaxed grains experiences the highest thermal cycle and the most intense plastic deformation during the welding process, causing dynamic recrystallization in this region [16]. In the thermo-mechanically affected zone (TMAZ), plastic deformation induces the formation of several dislocations within the crystal structure. Additionally, the increase in temperature leads to dynamic recovery, thereby resulting in the formation of a large number of sub-grains [17]. A previous study demonstrated that the HAZ exclusively underwent the thermal cycle that results in a partial coarsening of the grains, which typically leads to a morphology that closely resembles that of the base metal [18].
Previous studies have also demonstrated that the fatigue crack growth rate in FSWed joints is influenced by the interaction between residual stress distribution and the microstructure [19,20]. The high level of compressive residual stress near the welded zone can induce crack closure effects, while tensile stress increases the effective stress intensity factor range [21,22]. Fratini et al. [23,24] noted that stress relief did not alter the hardness and microstructural properties of FSWed structures, but the closure phenomena resulting from residual stress did affect the growth rates. Ilman et al. [25] stated that the effect of the microstructure on the fatigue crack growth rate was not significant when compared to residual stress. However, in the study by Anderson-Wedge et al. [26], the changes in crack growth were attributed to the depletion of strengthening precipitates in thermally affected zones rather than to the residual stresses.
Considering the varying conclusions from previous studies, this paper aims to investigate the coupling effect between residual stress and microstructure on the fatigue crack growth behavior of 2024 and 7075 aluminum alloys. The residual stresses were measured with the X-ray diffraction method, and the microstructure was observed with TEM. The finite element model was built using ANSYS and FRANC3D to calculate the fatigue crack growth rate, and the results were compared with those of the subsequent experiments.

2. Experimental Section

Rectangular plates of 2024-T3 and 7075-T6 alloys that were 87.5 mm × 400 mm × 8 mm in size were joined using the FSW-RL31-010 machine (Beijing FSW Technology Co., Ltd., Beijing, China) at Northwestern Polytechnical University in Xi’an, China. The shoulder diameter was 20 mm, and the probe diameters were 5 mm and 7.5 mm. Table 1 displays the chemical compositions of the 2024-T3 and 7075-T6 alloys, which were produced by Shanghai Miandi Metal Group Co., Ltd. (Shanghai, China) The welding parameters are shown in Table 2.

3. Residual Stress Measurement

The residual stress distribution of the 2024 specimen (400 rpm-150 mm/min) was analyzed using the X-350A X-ray diffraction setup at Xi’an Jiaotong University in Xi’an, China (Figure 1a). The stress distributions in both the longitudinal direction (LD) and the transverse direction (TD) were measured separately. The scan line is illustrated in Figure 1b.
The residual stress profile of the 2024 FSWed specimen (400 rpm-150 mm/min) is shown in Figure 2. The LD stress profile exhibits a “double peak” near the welded zone, with a peak stress of 103.2 MPa on the advancing side and a maximum stress of 93.5 MPa on the retreating side. This difference arises due to the comparatively higher heat input on the advancing side. In the TD, the residual stress values are relatively uniform, with a peak stress of 27.65 MPa on the advancing side.

4. Fatigue Crack Growth Experiment

The fatigue crack growth experiments in this study were conducted using an INSTRON 810 testing machine at Northwestern Polytechnical University, Xi’an, China, with a maximum loading force of 26 kN (Figure 3a). This setup ensured symmetrical load distribution with a load error of less than ±1% and minimal variation in the indicating value. The testing machine was equipped with an accurate counting device to meet the testing requirements. The compact tension (C(T)) specimens were machined according to the ASTM-647 standard [27]; the size of the specimen is shown in Figure 3b. It can be seen that the cracks in the C(T) specimens were parallel to the welded joints. To facilitate fixture fixation and crack growth observation, the surfaces of the specimen were polished before the experiment (Figure 3c). Fatigue tests were performed at R = 0.1 for all the C(T) specimens, with a fatigue load frequency of 20 Hz. Load ranges of 6.21 kN for the 2024 and 7075 homogeneous specimens, 8.04 kN for the 2024/7075 dissimilar specimens, and 3.26 kN for the base metal were applied.
The results of the fatigue crack growth rate experiment of the 2024 homogeneous FSWed specimens with different rotational speeds are presented in Figure 4. It can be seen that the crack growth rate of the 2024 base metal is higher than that of the welded specimen. Under the same welding speed, the fatigue crack growth rate of the 400 rpm specimen is higher than that the of 200 rpm specimen. As seen in Figure 4b, the crack path of 200 rpm was deflected towards the loading hole. With further crack propagation, the crack crossed through the WZ into the thermo-mechanically affected zone (TMAZ). Figure 4c shows that the crack growth path of the 400 rpm specimen remained within the WZ. According to previous studies on the dynamic tensile properties of the WZ, the increase in rational speed reduces the size of the particles in the WZ and makes them more uniform [28]. This enhances the mechanical properties of the WZ and reduces the fatigue crack growth rate in this region. Therefore, it can be concluded that tensile longitudinal residual stress promotes the crack growth and that it offsets the decrease in fatigue crack growth rate by grain refinement.
Figure 5a displays the fatigue crack growth results of the 7075 homogeneous FSWed specimens with different welding speeds. It is evident that welding speed exerts a significant influence on the crack growth rate. At the initial stage, the crack growth rate of the 7075 base metal was much higher than that of the FSWed specimen. Notably, the fatigue crack growth rate of the specimen with a welding speed of 100 mm/min was higher than that of the specimen with a speed of 150 mm/min. In a previous study, it was demonstrated that the mechanical properties of the 100 mm/min and 150 mm/min specimens were similar [29]; however, the longitudinal residual stress of the 100 mm/min specimen was higher than that of the 150 mm/min specimen. Consequently, the reduction in longitudinal residual stress contributed to the lower fatigue crack growth rate.
Figure 6a shows the fatigue crack growth rate curve of the 2024/7075 dissimilar FSWed specimen with a welding speed of 150 mm/min and rational speed of 400 rpm. It was compared with the 2024 and 7075 homogeneous FSWed specimens under the same welding parameters. The results indicated that, during the initial stage of crack growth, the fatigue crack growth rate of the 2024/7075 dissimilar specimens was between that of the 2024 and 7075 homogenous specimens. With further crack propagation, the dissimilar crack growth rate gradually approached that of the 2024 homogeneous specimen. This could be attributed to the lower hardness of the 2024 material, which facilitates crack propagation and ultimately allows the crack to extend into the 2024 side.

5. Numerical Analysis of Fatigue Crack Growth

The residual stress distribution of the FSWed specimens was calculated using ANSYS software, and the stress results were subsequently imported into FRANC3D for crack growth and post-processing analysis. The calculation process is illustrated in Figure 7.

5.1. Residual Stress Distribution of the Fatigue Specimen

The coupled thermo-mechanical model was employed to calculate the residual stress distribution of the welded plates [28]. Considering the influence of the mechanical properties on the fatigue crack growth rate, the model was divided into seven zones, which represent the base metal (BM), the advancing HAZ, the advancing TMAZ, the WZ, the retreating TMAZ, and the retreating HAZ, respectively. The thermo-mechanical properties of the alloys are shown in Table 3, Table 4 and Table 5 [13]. For the 2024/7075 dissimilar welded zone, the parameters were derived as the mean of those for the 2024 and 7075 specimens. Figure 8a shows the size and position of the C(T) specimen. The “birth and death element” method was used to calculate the residual stress of the C(T) specimen (Figure 8b). The FSWed plate model includes 105,992 elements, while the C(T) specimen contains 40,768 elements (Figure 8c).
The rate of heat generation can be obtained using the following equation [30]:
q ˙ = r 0 R 0 2 π ω r 2 μ ( T ) p ( T ) d r = 2 3 π ω μ ( T ) p ( T ) ( R 0 3 r 0 3 )
The boundary conditions are illustrated in Figure 9. The surface heat input from the shoulder was applied to the top surface of the specimen with a radius of 10 mm. The volume heat from the pin was generated as a body heat source with dimensions of 5 mm in diameter and 7.5 mm in length. As the temperature varied over time during welding, the coefficient of friction (μ) was assumed to be 0.6, and the pressure of the welding tool (p) was 10 kN. The ambient temperature was set at 25 °C. The heat transfer coefficient of the top and side surfaces was assumed to be 20 W/(m2 °C), while that of the bottom surface was 300 W/(m2 °C).
A comparative analysis of the numerical and measured residual stress profiles is shown in Figure 10. The numerical simulation of residual stress also exhibits a “double-peak” pattern. In the advancing TMAZ, the maximum LD stress reaches 126 MPa, which is 22.5 MPa higher than that of the measurements. Moreover, for the TD, the numerical and measured residual stress values are close, with a difference of 5.3 MPa in the retreating TMAZ. It can be seen that the numerical simulation results are in agreement with the X-ray diffraction measurements. Thus, the residual stress model presented in this paper is deemed reliable.
The longitudinal residual stress distribution of the 2024 C(T) specimen is shown in Figure 11a. Path 1 represents the normal direction in the middle of the specimen, while Path 2 highlights its crossing path. Upon comparison with the welded plate, it is evident that the residual stress in the C(T) specimen was redistributed after being removed from the welded plate. Figure 11b illustrates that the residual stress along Path 1 decreased from top to bottom. Figure 11c shows the longitudinal residual stress of the 2024 C(T) specimen with different rotational speeds in Path 2. It can be seen that for the same welding speed (150 mm/min), the maximum longitudinal residual stress in Path 2 increased with the increasing rotational speed. The impact of the welding speed on the residual stress of the 7075 was investigated, and the result is displayed in Figure 11d. The maximum longitudinal residual stress in Path 2 decreased with increasing welding speed.

5.2. Calculation Results of Fatigue Crack Growth

In this paper, fatigue crack propagation in the FSW joints was designed using the Walker model, which incorporates the effects of residual stress and stress relaxation. The Walker model has a simple form, requires few parameters, and is capable of predicting crack growth during stable propagation. The results of the study in Ref. [24] demonstrated that the Walker model exhibited good consistency with the fatigue crack growth tests.
d a d N = C [ ( 1 R ) m Δ K ] n
where a is the crack length; N is the number of propagation cycles; d a d N is the crack growth rate; C is the coefficient of the Walker model; m and n are the exponents; and ΔK is the stress intensity factor range.
According to the crack growth rates of the 2024 and 7075 base metal specimens, the material constants C, m, and n were determined via data analysis and linear fitting (Table 6). For the dissimilar welded specimens, the average values of the material constants from the 2024 and 7075 base metals were utilized. Figure 12 presents the comparison of the fatigue specimens and the results of the numerical calculation, respectively. It is shown that the crack growth predictions were in agreement with the experimental results.
The residual stress with a rotational speed of 400 rpm and a welding speed of 150 mm/min was applied to the C(T) specimen. The fatigue crack growth model was designed with the same load conditions as those in the experiments. To assess the effects of the mechanical properties and residual stress on fatigue crack growth behavior, the fatigue crack growth rates under the following five conditions were compared. These conditions, respectively, represent 2024 “base metal”, “Mechanical properties” (only considering the mechanical properties of the FSWed specimen), “Longitudinal direction” (only considering the residual stress of the longitudinal direction), “Transverse direction” (only considering the residual stress of the transverse direction), and “Normal direction” (only considering the residual stress of the normal direction). The results shown in Figure 13a demonstrate that the 2024 base metal exhibited the highest fatigue crack growth rate. However, the crack growth rate decreased when the residual stresses of the transverse and normal directions were considered in the calculations. This decrease occurred because of the presence of compressive stress in these two directions, although the stress level was quite low. On the other hand, the tensile stress in the longitudinal direction led to an increase in the crack growth rate.
The comparison of the fatigue crack growth paths is presented in Figure 13b. The crack growth path of the “Mechanical properties” condition aligns with that of the base metal, indicating that the mechanical properties of the FSWed specimen did not influence the fatigue crack deflection. When residual stress was taken into account, all the cracks in the specimen deflected upward. However, the crack deflection angle that resulted from the longitudinal residual stress was larger than that observed in the other directions. This difference can be attributed to the higher residual stress value in the longitudinal direction and the greater stress gradient along the thickness of the specimen as compared to the transverse and normal directions.
The comparison of the fatigue crack growth rates for the 2024 specimen under different rotational speeds is shown in Figure 14a. The results demonstrate that in the 2024 FSWed specimen, the fatigue crack growth rate increased with the increasing rotational speed. This occurred because the residual stress in the longitudinal direction in the 600 rpm specimen was higher than that in the other directions. Thus, the changes in mechanical properties that occurred due to grain refinement in the welded joint were counteracted, which thereby led to an increased rate of crack growth. In Figure 14b, the fatigue crack growth rate of the 7075 specimens decreased with the increasing welding speed. Hence, for the 7075 fatigue specimens, the fatigue crack growth rate was also predominantly influenced by the residual stress.

6. Microstructure Analysis

The fracture morphologies of the 2024 (400 rpm-150 mm/min) specimens in the fatigue propagation zone are presented in Figure 15. These specimens exhibited fatigue striations. Notably, the crack at the front exhibited an inclination rather than expanding vertically. This phenomenon occurred as the residual stress in the specimen gradually decreased from the top to the bottom, and because the grain size at the bottom of the WZ was smaller than that at the top [13]. As a result, the rate of fatigue crack propagation at the top was faster than that at the bottom.
In Figure 16, the TEM diagrams of the 2024 welded joints with different rotational speeds are shown using the Talos F200X at Xi’an University of Architecture and Technology in Xi’an, China. The TEM images of the WZ confirm the presence of needle-shaped S-phase (Al2CuMg) precipitates [31]. In the WZ (Figure 16a,d), most of the dislocations appeared in the form of distortion and entanglement, which resulted from the significant distortion caused by the rotation of the stirring probe. In the TMAZ, the S-phase was also present, with more dislocation entanglement observed around the grain boundaries (Figure 16b,e). This corresponds to the residual stress measurements shown in Figure 2. In the HAZ of the 400 rpm-150 mm/min specimen (Figure 16c), the precipitated S-phase increased due to the influence of the welding thermal cycling, while the extent of dislocation entanglement decreased relatively. In Figure 16f, fewer dislocation cell substructures and AlCu3 precipitates (remnants from the aging treatment process) were observed in the base material.
A comparison between the 200 rpm and 400 rpm specimens revealed that the 200 rpm specimen exhibited a higher dislocation density and a greater presence of precipitated phases in the WZ and TMAZ. During the process of fatigue crack propagation, these precipitated phases not only pinned dislocations but also facilitated the generation of additional dislocations, thereby leading to dislocation entanglement. As a result, the dislocation density and sliding resistance increased within the welding joint, which thus resulted in a decreased rate of fatigue crack propagation. This explains why the fatigue crack growth rate of the 200 rpm specimen is lower than that of the 400 rpm from a microstructural perspective, as shown in Figure 5a. Additionally, when compared to the WZ, the S-phase precipitates in the TMAZ and HAZ were coarser; this is a phenomenon that is well-documented in the literature [32,33]. Therefore, the crack predominantly propagated toward the TMAZ, as shown in Figure 4 and Figure 5.

7. Conclusions

In this study, the effect of residual stress and microstructure on the fatigue crack growth behavior of 2024-T3 and 7075-T6 friction stir welded joints with different welding parameters was analyzed. The key findings are as follows:
(1)
The welding parameters significantly influenced the fatigue crack growth rate. For the 2024 specimens, the fatigue crack growth rate increased with the higher rotational speed, while for the 7075 specimens, it decreased with the higher welding speed.
(2)
The longitudinal residual stress had the most significant impact on crack deflection. Additionally, the tensile residual stress in the welded joints accelerated crack growth, partially counteracting the reduction in the fatigue crack growth rate due to grain refinement. Conversely, the compressive residual stress reduced the stress intensity factor, thereby resulting in a diminished crack driving force.
(3)
The microstructural characteristic of the welded joints aligns with the distribution of residual stress. It was revealed that the precipitated phase both pinned dislocations and induced the generation of additional dislocations, thereby resulting in dislocation plugging. Consequently, as the rotational speed decreased, the dislocation density and sliding resistance increased, which thus led to a decreased rate of fatigue crack propagation. Furthermore, the TMAZ and HAZ contained coarser precipitates compared to the WZ, causing the crack to deflect towards the TMAZ.

Author Contributions

Conceptualization, Y.G.; Methodology, Y.G.; Software, Y.G.; Resources, P.L.; Data curation, P.L.; Writing—original draft, Y.G.; Funding acquisition, Y.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received external funding from the Post-doctoral Research Project in Shaanxi Province (No. 2023BSHEDZZ266).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rajendran, C.; Kumar, M.V.; Sonar, T.; Mallieswaran, K. Investigating the Effect of PWHT on microstructural features and fatigue crack growth behavior of friction stir welded AA2024-T6 aluminum alloy joints. Forces Mech. 2022, 8, 100107. [Google Scholar] [CrossRef]
  2. Thomas, W.M.; Nicholas, E.D.; Needham, J.C.; Murch, M.G.; Temple-Smith, P.; Dawes, C.J. Friction Stir Butt Welding. International Patent Application No. PCT/GB92102203; Great Britain Patent Application No. 9125978.8, 6 December 1991. [Google Scholar]
  3. Oliveira, J.P.; Duarte, J.F.; Inácio, P.; Schell, N.; Miranda, R.M.; Santos, T.G. Production of Al/NiTi composites by friction stir welding assisted by electrical current. Mater. Des. 2017, 113, 311–318. [Google Scholar] [CrossRef]
  4. Ke, W.C.; Oliveira, J.P.; Ao, S.S.; Teshome, F.B.; Chen, L.; Peng, B.; Zeng, Z. Thermal process and material flow during dissimilar double-sided friction stir spot welding of AZ31/ZK60 magnesium alloys. J. Mater. Res. Technol. 2022, 17, 1942–1954. [Google Scholar] [CrossRef]
  5. Zhang, X.; Ma, Y.E.; Wang, Z.; Guo, Y. The effect of residual stress on mixed-mode crack propagation behaviour in friction stir welded 7075-T6 aluminium alloy panel under biaxial loading. Fatigue Fract. Eng. Mater. Struct. 2021, 44, 805–821. [Google Scholar] [CrossRef]
  6. Das, J.; Banik, S.R.; Reddy, S.R.S.K.; Sankar, M.R.; Robi, P.S. Review on process parameters effect on fatigue crack growth rate in friction stir welding. Mater. Today Proc. 2019, 18, 3061–3070. [Google Scholar] [CrossRef]
  7. Luan, G.H.; North, T.H.; Guo, D.L. Characterizations of frictions stir welding on aluminum alloy. Trans. China Weld. Inst. 2002, 23, 62–66. [Google Scholar]
  8. Milan, M.T.; Bose Filho, W.W.; Tarpani, J.R.; Malafaia, A.M.D.S.; Silva, C.P.O.; Pellizer, B.C.; Pereira, L.E. Residual stress evaluation of AA2024-T3 friction stir welded joints. J. Mater. Eng. Perform. 2007, 16, 86–92. [Google Scholar] [CrossRef]
  9. Dattoma, V.; Giorgi, M.D.; Nobile, R. An experimental study on residual stress relaxation in low-cycle fatigue of Inconel 718 Plus. Fatigue Fract. Eng. Mater. Struct. 2022, 46, 310–324. [Google Scholar] [CrossRef]
  10. Liu, R.F.; Young, M.C. Influence on structure and fracture mechanics evaluation of a BWR feedwater nozzle following weld overlay repair. J. Mech. 2023, 39, 113–134. [Google Scholar] [CrossRef]
  11. He, X.C.; Gu, F.S.; Ball, A. A review of numerical analysis of friction stir welding. Prog. Mater. Sci. 2014, 65, 1–66. [Google Scholar] [CrossRef]
  12. Mishra, R.S.; Ma, Z.Y. Friction stir welding and processing. Mater. Sci. Eng. R Rep. 2005, 50, 1–78. [Google Scholar] [CrossRef]
  13. Guo, Y.N.; Ma, Y.E.; Zhang, X.S.; Li, J. Study on residual stress distribution of 2024-T3 and 7075-T6 aluminum dissimilar friction stir welded joints. Eng. Fail. Anal. 2020, 118, 104911. [Google Scholar] [CrossRef]
  14. Cai, B.; Zheng, Z.Q.; He, D.Q.; Li, S.C.; Li, H.P. Friction stir weld of 2060 Al-Cu-Li alloy: Microstructure and mechanical properties. J. Alloys Compd. 2015, 649, 19–27. [Google Scholar] [CrossRef]
  15. Liu, M.; Tao, X.; Di, Z.; Qin, M.; Liu, Z.; Bai, S. Effect of Pre-Rolling on Microstructure and Fatigue Crack Propagation Resistance of a Third-Generation Al-Li Alloy. Materials 2023, 16, 7540. [Google Scholar] [CrossRef]
  16. Wang, F.F.; Li, W.Y.; Shen, J.; Hu, S.Y.; dos Santos, J.F. Effect of tool rotational speed on the microstructure and mechanical properties of bobbin tool friction stir welding of Al-Li alloy. Mater. Des. 2015, 86, 933–940. [Google Scholar] [CrossRef]
  17. Prangnell, P.B.; Heason, C.P. Grain structure formation during friction stir welding observed by the stop action technique. Acta Mater. 2005, 53, 3179–3192. [Google Scholar] [CrossRef]
  18. Threadgill, P.L.; Leonard, A.J.; Shercliff, H.R.; Withers, P.J. Friction stir welding of aluminium alloys. Int. Mater. Rev. 2009, 54, 49–93. [Google Scholar] [CrossRef]
  19. Vuherer, T.; Milčić, M.; Glodež, S.; Milčić, D.; Radović, L.; Kramberger, J. Fatigue and fracture behavior of Friction Stir Welded AA-2024-T351 joints. Theor. Appl. Fract. Mech. 2021, 114, 103027. [Google Scholar] [CrossRef]
  20. Lan, Z.; Zhong, H.L.; Li, S.C.; Zhao, H.; Chen, J.; Qi, L. Microstructure, mechanical properties and fatigue crack growth behavior of friction stir welded joint of 6061-T6 aluminum alloy. Int. J. Fatigue 2020, 135, 105556. [Google Scholar]
  21. Pasta, S.; Reynolds, A.P. Residual stress effects on fatigue crack growth in a Ti-6Al-4V friction stir weld. Fatigue Fract. Eng. Mater. Struct. 2008, 31, 569–580. [Google Scholar] [CrossRef]
  22. Bussu, G.; Irving, P.E. The role of residual stress and heat affected zone properties on fatigue crack propagation in friction stir welded 2024-T351 aluminium joints. Int. J. Fatigue 2003, 25, 77–88. [Google Scholar] [CrossRef]
  23. Fratini, L.; Zuccarello, B. An analysis of through-thickness residual stresses in aluminium FSW butt joints. Int. J. Mach. Tools Manuf. 2006, 46, 611–619. [Google Scholar] [CrossRef]
  24. Fratini, L.; Pasta, S.; Reynolds, A.P. Fatigue crack growth in 2024-T351 friction stir welded joints: Longitudinal residual stress and microstructural effects. Int. J. Fatigue 2009, 31, 495–500. [Google Scholar] [CrossRef]
  25. Ilman, M.N.; Kusmono, P.; Iswanto, T. Fatigue crack growth rate behaviour of friction-stir aluminium alloy AA2024-T3 welds under transient thermal tensioning. Mater. Des. 2013, 50, 235–243. [Google Scholar] [CrossRef]
  26. Anderson-Wedge, K.; Stubblefield, G.; Zhu, N.; Long, B.; Daniewicz, S.R.; Allison, P.; Sowards, J.; Rodriguez, O.; Amaro, R. Characterization of the evolution of 2219-T87 aluminum as a function of the friction stir welding process. Int. J. Fatigue 2021, 142, 105954. [Google Scholar] [CrossRef]
  27. ASTM-647-2011; Standard Test Method for Measurement of Fatigue Crack Growth Rates 1 [S]. ASTM International: West Conshohocken, PA, USA, 2008.
  28. Buglioni, L.; Tufaro, L.N.; Svoboda, H.G. Thermal cycles and residual stresses in FSW of aluminum alloys: Experimental measurements and numerical models. Procedia Mater. Sci. 2015, 9, 87–96. [Google Scholar] [CrossRef]
  29. Guo, Y.N.; Ma, Y.E.; Wang, F. Dynamic fracture properties of 2024-T3 and 7075-T6 aluminum friction stir welded joints with different welding parameters. Theor. Appl. Fract. Mech. 2019, 104, 102372. [Google Scholar] [CrossRef]
  30. Chen, C.M.; Kovacevic, R. Finite element modeling of friction stir welding-thermal and thermo mechanical analysis. Int. J. Mach. Tools Manuf. 2003, 43, 1319–1326. [Google Scholar] [CrossRef]
  31. Niu, P.; Li, W.; Yang, C.; Chen, Y.; Chen, D. Low cycle fatigue properties of friction stir welded dissimilar 2024-to-7075 aluminum alloy joints. Mater. Sci. Eng. A 2022, 832, 142423. [Google Scholar] [CrossRef]
  32. Genevois, C.; Deschamps, A.; Denquin, A.; Doisneau-Cottignies, B. Quantitative investigation of precipitation and mechanical behaviour for AA2024 friction stir welds. Acta Mater. 2005, 53, 2447–2458. [Google Scholar] [CrossRef]
  33. Zhang, Z.; Xiao, B.L.; Ma, Z.Y. Hardness recovery mechanism in the heat-affected zone during long-term natural aging and its influence on the mechanical properties and fracture behavior of friction stir welded 2024Al-T351 joints. Acta Mater. 2014, 73, 227–239. [Google Scholar] [CrossRef]
Figure 1. Residual stress measurement: (a) X-ray diffraction method; (b) scan line (mm).
Figure 1. Residual stress measurement: (a) X-ray diffraction method; (b) scan line (mm).
Materials 17 00385 g001
Figure 2. Results of residual stress measurement.
Figure 2. Results of residual stress measurement.
Materials 17 00385 g002
Figure 3. (a) Fatigue crack growth experiment; (b) size of the C(T) specimens (mm); (c) C(T) specimen.
Figure 3. (a) Fatigue crack growth experiment; (b) size of the C(T) specimens (mm); (c) C(T) specimen.
Materials 17 00385 g003
Figure 4. Fatigue crack growth results of the 2024 specimen: (a) crack growth rates; (b) crack growth path at 200 rpm; (c) crack growth path at 400 rpm.
Figure 4. Fatigue crack growth results of the 2024 specimen: (a) crack growth rates; (b) crack growth path at 200 rpm; (c) crack growth path at 400 rpm.
Materials 17 00385 g004
Figure 5. Fatigue crack growth results of the 7075 specimen: (a) crack growth rates; (b) crack growth path at 100 mm/min; (c) crack growth path at 150 mm/min.
Figure 5. Fatigue crack growth results of the 7075 specimen: (a) crack growth rates; (b) crack growth path at 100 mm/min; (c) crack growth path at 150 mm/min.
Materials 17 00385 g005
Figure 6. Fatigue crack growth results of the 2024/7075 dissimilar specimen: (a) crack growth rates; (b) crack growth path.
Figure 6. Fatigue crack growth results of the 2024/7075 dissimilar specimen: (a) crack growth rates; (b) crack growth path.
Materials 17 00385 g006
Figure 7. Flow chart of the process for calculating the fatigue crack growth behavior.
Figure 7. Flow chart of the process for calculating the fatigue crack growth behavior.
Materials 17 00385 g007
Figure 8. Residual stress model of the C(T) specimen: (a) size of the specimen (mm); (b) “birth and death element” method (Pa); (c) finite element meshing.
Figure 8. Residual stress model of the C(T) specimen: (a) size of the specimen (mm); (b) “birth and death element” method (Pa); (c) finite element meshing.
Materials 17 00385 g008
Figure 9. Boundary conditions.
Figure 9. Boundary conditions.
Materials 17 00385 g009
Figure 10. Comparison of the residual stress profiles of the numerical and X-ray diffraction results.
Figure 10. Comparison of the residual stress profiles of the numerical and X-ray diffraction results.
Materials 17 00385 g010
Figure 11. Residual stress of the C(T) specimen: (a) longitudinal residual stress distribution of 2024 (400 rpm-150 mm/min) (Pa); (b) residual stress in Path1; (c) effect of rotational speed on LD residual stress in Path 2; (d) effect of welding speed on LD residual stress in Path 2.
Figure 11. Residual stress of the C(T) specimen: (a) longitudinal residual stress distribution of 2024 (400 rpm-150 mm/min) (Pa); (b) residual stress in Path1; (c) effect of rotational speed on LD residual stress in Path 2; (d) effect of welding speed on LD residual stress in Path 2.
Materials 17 00385 g011
Figure 12. Comparison of the experimental results and numerical analysis: (a) 2024 specimen; (b) 7075 specimen; (c) 2024/7075 dissimilar welded specimen.
Figure 12. Comparison of the experimental results and numerical analysis: (a) 2024 specimen; (b) 7075 specimen; (c) 2024/7075 dissimilar welded specimen.
Materials 17 00385 g012
Figure 13. Comparison of the fatigue crack growth rate in the 2024 specimen: (a) curve of the fatigue crack growth rate; (b) path of the fatigue crack growth.
Figure 13. Comparison of the fatigue crack growth rate in the 2024 specimen: (a) curve of the fatigue crack growth rate; (b) path of the fatigue crack growth.
Materials 17 00385 g013
Figure 14. Effect of the welding parameters on fatigue crack growth: (a) effect of rotational speed (2024 specimen); (b) effect of welding speed (7075 specimen).
Figure 14. Effect of the welding parameters on fatigue crack growth: (a) effect of rotational speed (2024 specimen); (b) effect of welding speed (7075 specimen).
Materials 17 00385 g014
Figure 15. (a) Fracture morphology of the 2024 400 rpm-150 mm/min specimen; (b) local enlarged image.
Figure 15. (a) Fracture morphology of the 2024 400 rpm-150 mm/min specimen; (b) local enlarged image.
Materials 17 00385 g015
Figure 16. TEM diagrams of the welded joints of the 2024 specimen: (a) 400 rpm-150 mm/min WZ; (b) 400 rpm-150 mm/min advancing TMAZ; (c) 400 rpm-150 mm/min HAZ; (d) 200 rpm-150 mm/min WZ; (e) 200 rpm-150 mm/min advancing TMAZ; (f) base metal.
Figure 16. TEM diagrams of the welded joints of the 2024 specimen: (a) 400 rpm-150 mm/min WZ; (b) 400 rpm-150 mm/min advancing TMAZ; (c) 400 rpm-150 mm/min HAZ; (d) 200 rpm-150 mm/min WZ; (e) 200 rpm-150 mm/min advancing TMAZ; (f) base metal.
Materials 17 00385 g016
Table 1. Chemical compositions of the 2024-T3 and 7075-T6 alloys.
Table 1. Chemical compositions of the 2024-T3 and 7075-T6 alloys.
MaterialChemical Composition (%)
2024-T3CuSiFeMnMgZnCrTiAl
3.8–4.90.50.50.3–0.91.2–1.80.250.10.15Base
7075-T6SiFeMnCuMgCrZnTiAl
0.40.50.31.2–2.02.1–2.90.18–0.285.1–6.10.2Base
Table 2. Welding parameters.
Table 2. Welding parameters.
SpecimenMaterialRotational Speed (rpm)Welding Speed (mm/min)
12024 homogeneous400150
22024 homogeneous200150
37075 homogeneous400150
47075 homogeneous400100
52024/7075 dissimilar400150
Table 3. Thermo-mechanical properties of the 2024 and 7075 aluminum alloys.
Table 3. Thermo-mechanical properties of the 2024 and 7075 aluminum alloys.
Temperature (°C)25100200300400500
2024Specific heat (J/(kg·K))900910950102011001100
Density (kg/m3)277027602730271026902670
Thermal conductivity (W/(m·K))126129153178187194
Poisson’s ratio0.330.330.330.330.330.33
Young’s modulus (MPa)7.2 × 1047.1 × 1046.3 × 1045.3 × 1044.5 × 1044.3 × 104
Coefficient of thermal expansion (×10−5/K)2.112.292.382.472.652.87
7075Specific heat (J/(kg·K))835.48979741012.511281205
Density (kg/m3)280028002800280028002800
Thermal conductivity (W/(m·K))114.8128.4142.2152.7160.8166.7
Poisson’s ratio0.330.330.330.330.330.33
Young’s modulus (MPa)7.3 × 1047.2 × 1046.4 × 1045.6 × 1044.8 × 1044.8 × 104
Coefficient of thermal expansion (×10−5/K)2.32.442.662.873.093.2
Table 4. Mechanical parameters of the 2024 FSWed specimen.
Table 4. Mechanical parameters of the 2024 FSWed specimen.
Temperature (°C)25100200300400500
2024 base metalYield stress (MPa)30626115257135
Plastic modulus (MPa)5304003302339565
2024 advancing HAZYield stress (MPa)32627816260135
Plastic modulus (MPa)424289103633515
2024 advancing TMAZYield stress (MPa)2051741013893
Plastic modulus (MPa)82075060037021090
WZYield stress (MPa)2291981134293
Plastic modulus (MPa)96087072027010570
2024 retreating TMAZYield stress (MPa)2111801053993
Yield stress (MPa)82075060037021090
7075 retreating HAZYield stress (MPa)31326715558135
Plastic modulus (MPa)424289103633515
Table 5. Mechanical parameters of the 7075 FSWed specimen.
Table 5. Mechanical parameters of the 7075 FSWed specimen.
Temperature (°C)25100200300400500
7075 base metalYield stress (MPa)455389278473320
Plastic modulus (MPa)250210150501510
7075 advancing HAZYield stress (MPa)381326233392716
Plastic modulus (MPa)300260190904525
7075 advancing TMAZYield stress (MPa)252215154261811
Plastic modulus (MPa)99087077530014590
WZYield stress (MPa)259221158261811
Plastic modulus (MPa)9408306802307550
7075 retreating TMAZYield stress (MPa)264226161791811
Plastic modulus (MPa)99087077530014590
7075 retreating HAZYield stress (MPa)371317226382616
Plastic modulus (MPa)300260190904525
Table 6. Parameters of the Walker model.
Table 6. Parameters of the Walker model.
MaterialCmn
20242.71 × 10−104.3603.375
70756.79 × 10−103.9573.108
2024/70754.75 × 10−104.1593.242
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Guo, Y.; Li, P. Effect of Residual Stress and Microstructure on the Fatigue Crack Growth Behavior of Aluminum Friction Stir Welded Joints. Materials 2024, 17, 385. https://doi.org/10.3390/ma17020385

AMA Style

Guo Y, Li P. Effect of Residual Stress and Microstructure on the Fatigue Crack Growth Behavior of Aluminum Friction Stir Welded Joints. Materials. 2024; 17(2):385. https://doi.org/10.3390/ma17020385

Chicago/Turabian Style

Guo, Yanning, and Peiyao Li. 2024. "Effect of Residual Stress and Microstructure on the Fatigue Crack Growth Behavior of Aluminum Friction Stir Welded Joints" Materials 17, no. 2: 385. https://doi.org/10.3390/ma17020385

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop