4.8.1.1. Positioning of Alloy
The placement of the alloy strongly affected material stirring and mixing. Particularly, it can determine the final microstructure of joints made between Al alloys with significantly different mechanical properties [
257,
258]. The higher hot-strength material should be placed on the advancing side to enhance mechanical properties of the dissimilar joint [
259,
260,
261]. Local weld temperatures were highest on the AS where the highest shear rates appeared.
Welded joints of components made: the first of 2024-T351/5083-H112 alloys and the second of 7075-T651/2024-T351 alloys (
Table 3) studied by Niu et al. [
262] contained 2024 and 7075 alloys placed on the AS while 5083 and 2024 alloys placed on the RS, respectively.
Studying an 2024/7075 joint Niu et al. [
263] found that the top section of the SZ was composed of the BM of RS, whereas the middle and bottom sections of the BM of AS. Studying FSWed joints of 7075-T651/2024-T351 alloys (
Table 3),) Hasan et al. [
264] found that materials fixed location on the AS and RS of weld influenced the quality of joint, which was better in case when the softer base material was positioned on AS.
The proper FSWed joints of 7075-T6/2024-T3 alloys studied by Ge et al. [
61] were obtained the 7075 and 2024 sheets were placed as upper sheet RS and lower sheet AS, respectively.
Studying FSWed joints of components made of 7075-T651/5083-H111 alloys (
Table 3) Kalemba-Rec et al. [
265] reported that at higher TRSs and in the configuration with 5083 on the AS and 7075 on the RS was accompanied with occurrence of porosity, voids, or wormholes in the stir zone. The highest tensile strength defect-free joint was obtained with 5083 on the advancing side, 7075 on the retreating side, a TRS of 280 rpm, and the Triflute pin. Then the WE reached above 100%. However, the effect of alloy location on the WE was rather small.
The proper FSWed joints of 2024-T4/7075-T6 alloys, studied by Safarbali et al. [
266] were obtained when 2024 alloy was placed on AS while 7075 alloy was placed on RS of joint.
During studies of FSWed joints of plates of dissimilar 6351-T6/5083-H111 alloys Palanivel et al. [
267] placed 6351 alloy on the AS and 5083 alloy on the RS. They observed that the tool shoulder increased the material transport at the top surface, from the RS to the AS, pushing it downward within the tool pin diameter.
During studies on FSWed joints of components made of dissimilar alloys 2017A-T451 and 7075-T651 Hamilton et al. [
268] found that the 7075-alloy exhibited longer positron lifetimes than the 2017A alloy. The positron lifetime profiles across the weld comprised many local maxima and minima on the AS and RS, corresponding to the hardness behavior. Weld temperatures on the advancing side were greater compared to those on the retreating side promoting more precipitation on the AS away from the weld center. This behavior related to the higher positron lifetime on the advancing side compared to the retreating side, at the same distances from the weld center.
Studying FSWed joints of dissimilar 5052/Al-Mg
2Si alloys (
Table 3), Huang et al. [
269] utilized 5052 alloy on RS and Al-Mg
2Si one on AS. They observed that on the top of the RS and at the bottom of the AS, the weld nugget WN comprised no banded structure. Contrary, on the top of the AS and at the bottom of the RS, and at the center of the WN, a banded structure occurred. Such a band structure partly covering weld width extended from the AS toward the RS. A rich Al-Mg
2Si layer was formed at the weld top surface also of the RS. The interface between the 5052 alloy and the rich Al-Mg
2Si region at the bottom of the RS appeared as transitional layer with a thickness of 50 μm. The interface began from the RS at the top surface due to the materials on the RS were not driven to the AS. With enhancing distance from the top surface, the interface was located on the AS due to the materials on the RS were dragged to the AS.
Studying FSWed joints of components made of dissimilar 2024/6061 alloys (
Table 3) Moradi et al. [
270] found that the fraction of precipitates in the stirred zone of the retreating side exceeded that of the advancing side. The extent of continuous dynamic recrystallization in the thermo-mechanically affected zone TMAZ of the advancing side was less than that of the retreating side and the recrystallized grains seldom occurred on the advancing side. The initial texture components became asymmetric after FSW process. The overall texture intensity was weaker on the advancing side and stronger on the retreating side than that in the starting materials. The discontinuous static recrystallization and/or meta-dynamic recrystallization occurred on the advancing side. The microhardness profile on the advancing side was almost identical, while it comprised three distinguishable regions on the retreating side.
The FSWed joints of 6061-T6/6351-T6 alloys (
Table 3) studied by Prasanth and Raj [
271] were obtained for the cases where each of dissimilar 6061-T6 and 6351-T6 alloys was placed separately on the AS and the on the RS of joints.
Studying double-sided FSWed joints of components made of dissimilar 6082-T6/7075-T6 alloys (
Table 3), Azeez and Akinlabi [
178] reported that little abnormalities at the retreating side were caused by the pre-heating of the plates during the initial welding process.
In case of single-sided FSWed joints of components made of the same alloys Azeez and Akinlabi [
272] reported that some microstructure imperfection occurred at the weld nugget when 6082 Al plates were clamped on the RS to the backing plate. However, deviation in the positioning of the Al plates prevented the fabrication of good bonding and quality welds despite the material flow and mixing occurrence.
The proper FSWed joints of rolled plates made of dissimilar 6061-T651/5A06-H112 alloys (
Table 3), studied by Peng et al. [
273] were obtained when 6061 alloy was placed on AS while 5A06 alloy was placed on RS of joint.
The correct FSWed joints of dissimilar 6101-T6/6351 alloys (
Table 3), studied by Das and Toppo [
274] were obtained when 6101 alloy was placed on AS while 6351 alloy was placed on RS of joint.
The proper FSWed joints of dissimilar 2024-T3/6063-T6 alloys (
Table 3), studied by Sarsilmaz [
275] were obtained when 2024 alloy was placed on AS while 6063 alloy was placed on RS of joint.
Studying FSWed joints of components made of dissimilar 2219-T87/2195-T8 alloys (
Table 3), No et al. [
276] found that the best joining properties were obtained for conditions including 600 rpm and 180-240 mm/min when the 2219-T8 alloy was on the RS.
During studies on FSWed joints of components made of dissimilar wrought 2017A/ cast AlSi9Mg alloys (
Table 3), Kopyscianski et al. [
277] reported that the AlSi9Mg alloy on the AS dominated the weld center. The local maximum on the AS was on the nugget side with a high density of the bands of the 2017A alloy.
The proper FSWed joints of dissimilar 5083-H12/6061-T6 alloys studied by Ghaffarpour et al. [
278] were obtained when 6061 alloy was placed on AS while 5083 alloy was placed on RS of joints. The FSW joints possessed a much higher quality and improved mechanical properties than those obtained with TIG welding.
Studying the UFSWed joints of dissimilar alloys 6061 and 7075 Bijanrostrami et al. [
279] placed 6061 alloy on the AS and 7075 alloy on the RS on top of a steel backing plate. The FSWed joints of dissimilar Al 6082-T6/5083-H111 alloys (
Table 3), studied by Kasman et al. [
280] were obtained when 6082 alloy was positioned on AS while 5083 alloy was positioned on RS of joints. The small cavity- and tunnel-type defects occurred at the nugget zone and were located on the advancing side of the pin. These defects lowered the strength and elongation of the weld joint.
The FSWed joints of the 6 mm thick sheets made of dissimilar 5083-H111/6351-T6 alloys (
Table 3), studied by Palanivel et al. [[
144] were obtained when 6351-T6 alloy was placed on AS, while 5083-H111 alloy was placed on TS of joint. Similarly, during studies presented in [
183,
187] 6351 alloy was placed on AS, while 5083-H111 alloy was placed on the RS of joint. The grain size within the friction stir processed (FSPed) region was much smaller than in the parent material due to the higher temperature and extensive plastic deformation. The grain size in the TMAZ clearly differed from that in the FSPed region.
FSWed joints of dissimilar 5052-H32/6061-T6 alloys (
Table 3), studied by Doley and Kore [
281] were obtained when 5052 alloy was placed on RS while 6061 alloy was placed on AS of the joints. The microhardness values of the dissimilar welds were lower at heat-affected zones HAZ on both the sides of weld line, whereas the lowest one occurred at HAZ of 5052 alloy.
During studies on the effect of shoulder diameter to pin diameter ratio on microstructure and mechanical properties of FSWed joints of dissimilar 2024-T6/7075-T6 alloys (
Table 3), Saravanan et al. [
282] placed 2024-T6 alloy on the AS, and 7075-T6 on the RS. They reported that the joints fabricated with the ratios of 2 and 2.5 fractured in heat-affected zone HAZ region of the advancing side, and joints fabricated with the ratios of 3, 3.5, and 4 fractured at stir zone SZ. For all the D/d ratios, minimum hardness was seen at HAZ region in the advancing side and was maximum in the SZ and again decreases at HAZ in the retreating side.
Studying FSWed joints of sheets made of dissimilar Al-Mg-Si/Al-Zn-Mg alloys (
Table 3), Yan et al. [
283] and [
284]found that for the Al-Zn-Mg alloy positioned at the advancing side AS, the joints exhibited better fatigue properties caused by the bridging effect of the big secondary phase particles. For the Al-Zn-Mg alloy placed at the AS, there was limited movement of the Al-Mg-Si alloy material to the AS due to its easier flow. For the Al-Mg-Si placed at the RS, there was no RS material Al-Zn-Mg flow to AS due to the high resistance to flow of such a material.
The proper FSWed joints of dissimilar 2024-T3/6061-T6 alloys (
Table 3), studied by Zapata et al. [
285] were obtained when 2024 alloy was placed on AS while 6061 alloy was positioned on RS of joints.
Studying FSWed joints of the dissimilar UFGed 1050/ 6061-T6 alloys (
Table 3), Sun et al. [
286] reported that sound welds were performed at the wide revolutionary pitches from 0.5 to 1.25 mm/min, only when the 6061-T6 alloy was put on the AS. Otherwise, welds comprised the large defects formed in the softened 1050 Al side RS.
The proper FSWed joints of dissimilar 2024-T3/2198-T3 alloys (
Table 3), studied by Texier et al. [
287] were obtained when 2024-T3 and 2198-T3 sheets were on RS and AS of joints, respectively.
The correct FSWed joints of dissimilar 6061/7050 alloys (
Table 3), studied by Rodriguez et al. [
288] and [
289] were obtained when 7050 alloy was positioned on AS while 6061 alloy was positioned on RS of joints.
The proper lap FSWed joints of dissimilar 6111-T4/5023-T4 alloys (
Table 3), studied by Yoon et al. [
47]. Two different joints, one with 6111 as the top plate RS and the other with 5023 as the top plate, were used.
Studying FSWed joints of the dissimilar 6061/5086 alloys (
Table 3), Ilangovan et al. [
290] placed 6061 alloy on the AS and 5086 alloy on the RS. They found that the AS of the thermo-mechanically affected zone AS-TMAZ was the softest region in the microhardness plot for both pin profiles including straight cylindrical STC one, threaded cylindrical THC one and tapered cylindrical TAC one. It was due to the dissolution of precipitates in the AS-TMAZ region because of the heating and cooling cycles prevailing during welding. Under tensile loading the strain localization occurred in such a region causing the failure therein. Only slight hardness variations appeared at the RS.
During studies on butt FSWed joints of components made of dissimilar 2050/6061 alloys (
Table 3), Reza-E-Rabby et al. [
291] found that joint quality, process parameters and welding temperature depended on material orientation in FSW. Defect-free welded joints with effective material transportation in the weld nugget zone were formed for 2050 on the advancing side. In the latter case, the tool was also less loaded by in-plane reaction force.
The proper FSWed joints of dissimilar 5083-O/6082-T6 alloys (
Table 3), studied by Donatus et al. [
292] were obtained when 5083 alloy was positioned on AS while 6082 alloy was positioned on RS of joints.
Studying FSWed joints of components made of dissimilar cast Al–Si alloys A319/A413 (
Table 3), Karam et al. [
293] obtained sound joints between the A319 and A413 plates, when A413 alloy was placed on AS while A319 alloy was placed on RS of joints. Each tensed welded specimen was fractured outside the welded regions at the A413 base metal placed on the AS.
The proper butt FSWed joints of dissimilar 7075-O/6061-O and 7075-T6/6061-T6 alloys (
Table 3), studied by Ipekoglu and Cam [
294] were obtained when 6061 alloy was placed on AS while 7075 alloy was positioned on RS of joints.
Studying the FSWed joints of dissimilar 6061-T6/ 7075-T6 alloys (
Table 3), Cole et al. [
295] reported that quality of welds was sensitive to alloy placement, tool offset, and tool–workpiece interface temperature. Under tensile loading, the welds failed in the heat-affected zone of 6061 on the AS of the weld. Weld tool offsets into the retreating side 7075 enhanced the tensile strength of the joint. The weld AS was hotter than the RS at both the tool shoulder and pin. There was a 20 °C enhancement in advancing-side shoulder interface temperature when offsetting from −2 to +2 mm, while a lesser enhancement appeared at the pin interface (∼3 °C). The strongest welds (−2 mm offset) corresponded to the lowest temperatures on the AS.
During studies on lap FSWed joints of 5 mm thick sheets made of dissimilar 2024-T3/7075-T6 alloys (
Table 3), Song et al. [
296] found that the WS and joint combination affected the hook geometry which in return affected the lap shear strength. In all 2024/7075 joints, voids occurred, and the joints fractured from the tip of hook on AS along the SZ/TMAZ interface during lap shear test inducing tensile fracture mode. In 7075/2024 joints, the hook-on RS horizontally extended a long distance into the bottom stir zone at higher WSs. 7075/2024 joints exhibited greater failure load than 2024/7075 joints at lower WSs while the opposite trend occurred at higher WSs.
The proper FSWed joints of dissimilar 6061-T6/5083 alloys (
Table 3), studied by Jannet and Mathews [
297] were obtained when 6061 alloy was placed on the AS while 5083 alloy was placed on the RS of joints.
The correct FSWed joints of dissimilar 5083-H111/ 6351-T6 alloys (
Table 3), Palanivel et al. [
298] and [
299] obtained when 6351 alloy was placed on the AS while 5083 alloy was placed on the RS of joint.
During studies on butt FSWed joints of plates made of dissimilar 2014-T6/ 6061-T6 alloys (
Table 3), Jonckheere et al. [
300] found that the alloys placement or tool lateral shift affected the welds hardness as they influence the precipitate radius and volume fraction. The 2014 alloy was successively placed on the RS and on the AS. More 2014 occurred in nugget zone if tool was shifted towards 2014 alloy on the AS.
Similarly, Jonckheere et al. [
301] reported that material flow and joint quality, regardless of material placement, were affected by the welding conditions and their influences on heat input and weld nugget temperatures. If the 2014 alloy is placed on the AS of the weld, an abrupt transition between the weld nugget and the 6061-alloy appeared leading to premature fracture in tension.
The FSWed joints of the dissimilar A356/6061-T6 alloys (
Table 3), studied by Ghosh et al. [
302] and [
303] were obtained when 6061 alloy was placed on the AS while A356 alloy was placed on the RS of joints. According to [
302] low hardness of A356 alloy appeared at retreating side. The increment of hardness at AS was correlated to the higher strength of 6061 with respect to A356 alloy. It was due to a composite microstructure where both the alloys contributed appeared near weld line. As reported in [
303] during welding, in front of tool, material got plasticized and transported from the RS to the AS.
During studies on FSWed joints of dissimilar Al alloys Koilraj et al. [
304] found that the microstructures of the weld TMAZ on the AS exhibited highly deformed grains, with a clearly discernible SZ/TMAZ and TMAZ/HAZ boundaries. On the RS these boundaries were diffused, especially the latter. On the AS, there was a significant drop in hardness from the 2219 base material to the weld nugget boundary. On the RS, only slight drop in hardness appeared from the 5083-base material to the weld nugget boundary.
Studying the FSWed dissimilar cast and wrought 6061 alloy (
Table 3), Dinaharan et al. [
305] reported that the material location strongly affected the material flow behavior. The material in the advancing side occupied the major portion of the weld zone under enhanced TRS. The joint exhibited maximum tensile strength when cast Al alloy was positioned on the advancing side.
During studies on FSWed joints of dissimilar 5083-H111/6351-T6 alloys (
Table 3), Palanivel et al. [
306] reported that the transportation of plasticized material from AS to RS was uniform from top to bottom of the joint when straight pin profile tool was utilized.
The proper FSWed joints of dissimilar 5052-H34/5023-T4 alloys (
Table 3), studied by Song et al. [
127] were obtained when 5052 alloy was placed on the AS while 5053 alloy was placed on the RS of joints.
Studying FSWed joints for dissimilar 5052/A5J32 alloy sheets (
Table 3), Kim et al. [
307] obtained defect-free welds under all welding conditions with fixing the A5J32 alloy on the retreating side. However, for fixing the 5052 alloy on the retreating side, some welding defects occurred at the joint under certain welding conditions with a weakened heat input. Placing the high-strengthened Al alloy on the AS led to excessive agglomerations and defects due to limited material flow. Therefore, placing such a kind of Al alloy should be at the RS to limit the resistance to material flow.
The proper FSWed butt joints of dissimilar 7050-T7451/ 2024-T351 alloys (
Table 3), studied by Prime et al. [
308] were obtained when 2024 alloy was placed on the AS while 7050 was placed on the RS of joints.
Gérard and Ehrström [
259] suggested that the material with the higher solidus temperature should be on the AS to improve joint quality and to eliminate internal defects/porosity.
During studies on the butt FSWed joints of dissimilar 2024-T351/ 6056-T4 alloys (
Table 3), Amancio-Filho et al. [
260] placed the 2024-T351 alloy on the AS as the stronger one from the alloys joined.
In case of FSWed joints of wrought 6061/ Al and A356 Al alloys (
Table 3), Lee et al. [
261] found that the joint properties are strongly affected by the alloy on the RS. The mechanical properties of the stir zone were greater when 6061 Al alloys were fixed at the retreating side. The WE was of 80% for similar A356 joint, and for the cases of dissimilar alloys of 83% for A356 on AS, and of 87% for A356 on the RS, respectively.
Liu et al. [
57] studied the relations between welding parameters and tensile properties of the FSWed joints of components made of the 2017-T351 alloy (
Table 3). The voids-less joints fractured near or at the interface between the weld nugget and the thermo-mechanically affected zone TMAZ on the advancing side.
Studying the FSWed joints of components made of 1050-H24 alloy (
Table 3), Liu et al. [
309] found that the location of the maximum strain gradually moved to the RS from the AS of the joint. Therefore, the fracture location of the joint gradually changed to the RS from AS of the joint as the WS gradually increased.
The FSWed joints of dissimilar cast AlSi9Mg (hypoeutectic silumin)/ 2017A alloys (
Table 3), studied by Mroczka [
310] were obtained when 2017A alloy was positioned on AS while AlSi9Mg alloy was positioned on RS of joints. During the process, the welding line was offset toward the advancing side, and an additional heat source was applied from the root side. Studying FSWed joints of components made of 2017A alloy (
Table 3) Mroczka et al. [
311] found that microhardness exhibited a tendency to grow on the advancing side of the joint.
The FSWed joints of dissimilar 7003/7046 alloys (
Table 3), studied by Yang et al. [
312] were obtained when 7003 alloy was placed on AS while 7046 alloy was placed on RS of joints. The hardness was much higher on the RS than that on the AS, and the average hardness difference between the two sides was about 30HV. After artificial ageing, the hardness enhanced significantly, while the hardness difference rose to about 50HV for the two sides.
Kasman and Ozan [
313] studied butt FSWed joints of 6013 Al plates (
Table 3) obtained via pin offset technique. The highest tensile strength equal to 206 MPa was obtained under 1.5 mm pin offset towards the advancing side and 500 min−1 TRS.
Zhao et al. [
314] studied the influence of exchanging AS and RS side material on microstructure, mechanical properties, and electrochemical corrosion resistance for FSWed joints of components made of dissimilar 6013-T4/7003 alloys (
Table 3). The joint with the 6013-T4 placed at the AS was called the A6R7 joint. Accordingly, the A7R6 referred to the joint with the Al7003 placed at the AS. The authors reported that various joint cross-sections appeared when exchanging AS and RS materials. The material on the AS was more deformed during the welding process. When the Al6013 was positioned on the AS, the plastic flow of the weld was enough. Independently on the AS or RS, the Al6013-T4 side was the weak region for both tensile strength and hardness.
For, lap FSW process when the rotation speed was low, and the WS was high, void-type defects appeared on the AS or center of the nugget [
315].
Studying FSWed joints of 6061-T6 alloy (
Table 3), Juarez et al. [
316] reported that the surface fracture surfaces of tensile specimens, for the welding without heat treatment BMW and solubilized heat treatment and partial aging before welding HTBW cases, most of the fractures occurred on the AS of the tool and in the heat-affected zone. Fracture appeared on the unaffected material zone UFM and RS for the solubilized heat treatment and aging after welding HTAW case.
Godhani et al. [
317] obtained proper butt joints of dissimilar 6061-T6/ 7075-T6 alloys (
Table 3) when 6061 alloy was placed on the AS, while 7075 on the RS of the joint.
Investigating FSWed joints of components made of dissimilar 5052/ and 6061 alloys obtained with various pin-eccentric stir tools (the pin eccentricities of 0, 0.4, and 0.8 mm, respectively), Chen et al. [
318] found that sound joints were obtained for the 6061 alloy on the AS.
Zhang et al. [
319] studied the FSWed joints of 5 mm thick 7075AA7075-T651/2024 and AA2024-T351 similar and dissimilar alloys obtained using tool with cylindrical taper threaded pin, shoulder diameter of 15 mm, pin diameter 3.76 mm on the insertion side and of 6.66 mm on the shoulder side, pin length of 5 mm under TRS of 600, 950, 1300 and 1650 rpm, WS of 100 mm/min, tilt angle of 2.5 degrees. They found that the width of the TMAZ on the RS is greater than that of the AS.
Material flow under specific welding conditions was the common thread among the authors, with material placement close, but distinguishably secondary. Various configurations of FSW joints of various Al alloys were presented by Patel et al. [
256].
Table 3.
Configurations of FSW joints of various Al alloys.
Table 3.
Configurations of FSW joints of various Al alloys.
Refs. |
|
Alloy Combinations |
Thick (mm) |
Alloy Positioning |
Configuration |
AS |
RS |
262 |
Butt |
2024-T351/5083-H112 |
6.35 |
2024 |
5083 |
262 |
Butt |
7075-T651/2024-T351 |
6.35 |
7075 |
2024 |
264 |
Butt |
7075-T651/2024-T351 |
6 |
Both |
Both |
265 |
Butt |
7075-T651/5083-H111 |
6 |
Both |
Both |
265 |
Butt |
5052/AlMg2Si |
8 |
Al- Mg2Si |
5052 |
270 |
Butt |
2024-T351 /6061-T6 |
6 |
2024 |
6061 |
271 |
Butt |
6061-T6/6351T6 |
6.35 |
Both |
Both |
272,328 |
Butt |
6082-T6/7075-T6 |
10 |
7075 |
6082 |
273 |
Butt |
6061-T651and 5A06-H112 |
5 |
6061 |
5A06 |
274 |
Butt |
6101-T6/6351-T6 |
12 |
6101 |
6351 |
275 |
Butt |
2024-T3/6063-T6 |
8 |
2024 |
6063 |
276 |
Butt |
2219-T87/2195-T8 |
7.2 |
Both |
Both |
277 |
Butt |
2017A-T451/cast AlSi9Mg |
6 |
2017A |
AlSi9Mg |
280 |
Butt |
5083-H111/6082-T6 |
5 |
6082NR |
5083NR |
144 |
Butt |
5083-H111/6351-T6 |
6 |
6351 |
5083 |
144 |
Butt |
5083-H111/6351 |
6 |
6351 |
5083 |
282 |
Butt |
2024-T6/7075-T6 |
5 |
2024 |
7075 |
283,284 |
Butt |
Al-Mg-Si/Al-Zn-Mg |
15 |
Both |
Both |
286 |
Butt |
UFG 1050/6061-T6 |
2 |
Both |
Both |
287 |
Butt |
2024-T3/2198-T3 |
3.18 |
2198 |
2024 |
288,289 |
Butt |
6061-T6/7050-T7451 |
5 |
7050 |
6061 |
290 |
Butt |
5086-O/6061-T6 |
6 |
6061 |
5086 |
291 |
Butt |
2050-T4/6061-T651 |
20 |
Both |
Both |
292 |
Butt |
5083-O/6082-T6 |
NR(~7) |
5083 |
6082 |
293 |
Butt |
A319/A413 cast |
10 |
A413 |
A319 |
294 |
Butt |
7075-O/6061-O 7075-T6/6061-T6 |
3.17 |
6061 |
7075 |
295 |
Butt |
6061-T6/7075-T6 |
4.6 |
Both |
Both |
297 |
Butt |
5083-O/6061-T6 |
6 |
6061 |
5083 |
302,303 |
Butt |
A356/6061-T6 |
3 |
6061 |
A356 |
304 |
Butt |
2219-T87/5083-H321 |
6 |
2219 |
5083 |
305 |
Butt |
6061 cast/6061 rolled |
6 |
Both |
Both |
306 |
Butt |
6351-T6/5083-H111 |
6 |
6351 |
5083 |
127 |
Butt |
5052-H34/5023-T4 |
~1.5 |
5052 |
5023 |
307 |
Butt |
5052-H34/5023-T4 |
1.5 & 1.6 |
Both |
Both |
308 |
Butt |
7050-T7451/2024-T351 |
25.4 |
2024 |
7050 |
260 |
Butt |
2024-T351/6056-T4 |
4 |
2024 |
6056 |
261 |
Butt |
cast A 356/6061 |
4 |
Both |
Both |
57 |
Butt |
2017-T351 |
5 |
Both |
Both |
309 |
Butt |
1050-H24 |
5 |
Both |
Both |
310 |
Butt |
2017A-T451/AlSi9Mg |
6 |
2017A |
AlSi9Mg |
311 |
Butt |
2017A |
6 |
Both |
Both |
312 |
Butt |
7003-T4/7046-T4 |
3 |
7003 |
7046 |
314 |
Butt |
6013-T4/7003 |
2.8 |
Both |
Both |
313 |
Butt |
6013-T6 |
5 |
Both |
Both |
316 |
Butt |
6061-T6 |
9.5 |
Both |
Both |
317 |
Butt |
6061-T6/7075-T6 |
6 |
6061 |
7075 |
279 |
Underwater Butt |
6061-T6/7075-T6 |
5 |
6061 |
7075 |
46,268, 326 |
Butt NA Butt |
2017A-T451/ 7075-T651 |
6 |
Both |
Both |
298,299 |
NA Butt |
6351-T6/5083-H111 |
6 |
6351 |
5083 |
300,301 |
NA Butt |
2014-T6/6061-T6 |
4.7 |
Both |
Both |
278 |
NA |
5083-H12/6061-T6 |
1.5 |
6061 |
5083 |
285 |
NA |
2024-T3/6061-T6 |
4.8 |
2024 |
6061 |
281 |
NA |
5052/6061 |
1, 1.5 |
6061 |
5052 |
266 |
NA |
2024-T4/7075-T6 |
4 |
2024 |
7075 |
267 |
NA |
6351-T6/5083-H111 |
6 |
6351 |
5083 |
315 |
Lap |
6111-T4/5023-T4 |
1 |
Both |
Both |
296 |
Lap |
2024-T3/7075-T6 |
5 |
Both |
Both |
47 |
Lap |
6111-T4/5023-T4 |
1 |
Both |
Both |
61 |
Lap |
7075-T6/2024-T3 7075-upper; 2024-lower |
3 |
2024 |
7075 |
It can be notice that the material position (AS/RS) plays significant role in FSW process, particularly in the case of dissimilar Al alloys. The placement of harder material in the AS, for both butt and to a lesser extent lap configuration, provides better joints’ quality. It is in agreement with observations from [
320].
4.8.1.2. Tool Rotation and Welding Speeds
The tool rotation speed influenced the intensity of plastic deformation and thus material mixing. Kalemba-Rec et al. [
265] found that material mixing was proportional to the tool rotation speed for a dissimilar 7075-5083 joint. Under high tool rotation speeds numerous imperfections including poor surface (flash), voids, porosity, tunneling or formation of wormholes occurred due to the excessive heat input [
322,
323] Low WSs enhanced heat input and are often accompanied by defects such as tunneling [
297,
302,
321,
324,
325].
Obtaining defect-free joint with a good metallurgical bond and mechanical properties needs the selection of the appropriate/optimized combination of tool rotation and WS, particularly for combinations of dissimilar Al alloys [
271,
276,
278,
279,
281,
285,
286,
293,
297,
298,
302,
303,
304]. Welded joints of components made: the first of 2024-T351/ 5083-H112 alloys and the second of 7075-T651/ 2024-T351 alloys (
Table 4), studied by Niu et al. [
262] were obtained under TRS of 600 rpm and WS of 150 mm/min, respectively.
The proper FSWed joints of 7075-T651/2024-T351 alloys (
Table 4), studied by Hasan et al. [
264] were obtained under TRS of 900 rpm and WS of 150 mm/min.
Ge et al. [
61] studied how EST affects the shear failure load of lap joints. Shear fracture mode occurred in lap joints obtained with a small (3-mm) pin at all WSs. The higher TRS, lower WS, and bigger plunge depth improve the diffusion bonding strength of lap joint. The lap shear failure load decreased with the increase of WS, due to worsened diffusion bonding induced by lower heat input. The highest lap shear failure load with a small pin was obtained at WS of 60 mm/min.
Studying FSWed joints of components made of 7075-T651/ 5083-H111 alloys (
Table 4), Kalemba-Rec et al. [
265] reported that at higher TRSs and in the configuration with 5083 on the advancing side and 7075 on the retreating side was accompanied with occurrence of porosity, voids, or wormholes in the stir zone. The highest tensile strength defect-free joint was obtained for a TRS of 280 rpm. The higher TRS at a constant traverse speed caused the lower joint efficiency.
Saeidi [
324] found that an enhancement in TRS from 450 to 800 rpm, at a selected WS (30, 41.5, or 50 mm/min) initially lowered and then enhanced the joint efficiency.
The proper FSWed joints of 2024-T4/ 7075-T6 alloys (
Table 4), studied by Safarbali et al. [
266] were obtained under TRS of 1140 rpm, and WS of 32 mm/min.
Palanivel et al. [
267] conducted studies on FSWed joints of components made of dissimilar 6351-T6/ 5083-H111 alloys (
Table 4) focusing on optimization of, inter alia, TRS and WS. The WE reached up to 78.7% for the full impeller shoulder tool, WS of 60 mm/min and TRS of 1000 rpm.
For study on FSWed sheets made of dissimilar alloys 2017A-T451 and 7075-T651 Hamilton et al. [
326] obtained qualitied welds using a tool positioned with the tool tilt angle of 1.5 deg. The TRS and WS were 355 rpm and 112 mm/min, respectively, and the applied force during processing was of 32.8 kN.
Gupta et al. [
327] conducted studies on FSWed joints of components made of dissimilar 5083-O and 6063-T6 alloys focusing on optimization of tool geometry, TRS, and WS. The multi-optimal set of weld properties comprising tensile strength, average hardness at weld nugget zone, the set of process parameters and average grain size at weld nugget zone was obtained for 900 rpm of TRS, 60 mm/min of WS. The WE reached up to 76.4%.
Huang et al. [
269] quoted that for Al metal matrix composites MMCs, the material flow depended on the rotation speed and the reinforcing phases. The welding of Al-Mg/ Al-Mg
2Si alloys (
Table 4) also depended on the hard and brittle intermetallic compounds of the primary Mg
2Si phases in Al-Mg
2Si alloys.
The proper FSWed joints of dissimilar 2024/6061 alloys (
Table 4), studied by Moradi et al. [
270] were obtained under a TRS of 800 rpm and WS of 31.5 mm/min.
The FSWed joints of dissimilar 6351-T6/ 6061-T6 alloys (
Table 4), studied by Prasanth and Raj [
271] were obtained under TRS values of 600, 900, and 1200 rpm, and WS values of 30, 60, and 90 mm/min. The highest WE was obtained for TRS of 900 rpm, WS of 60 mm/min and axial force of 6 kN.
The proper FSWed joints of dissimilar 6082-T6/ 7075-T6 alloys (
Table 4), studied by Azeez et al. [
272,
328] were obtained under TRS values of 950, and 1000 rpm, and WS values of 80 and 100 mm/min.
Peng et al. [
273] studied FSWed joints of rolled plates made of dissimilar 6061-T651 and 5A06-H112 alloys (
Table 4), obtained for various rotation speeds and transverse speeds. With the increase of rotation speed, more heat was generated during FSW. The increasing of heat input could enlarge the size of HAZ and reduce the slant angle of HAZ and thus lead the fracture angle to decrease and cause the dimples change from inclined ones to normal ones.
Studying FSWed joints of plates made of dissimilar 6101-T6/ 6351-T6 alloys (
Table 4), Das and Toppo [
274] found that with an enhancement of the TRS the impact strength behavior pointed to a high change in mechanical properties. The impact test samples exhibited a ductile fibrous fracture.
Investigating FSWed joints of components made of dissimilar 2024-T3/6063-T6 alloys (
Table 4), Sarsilmaz [
275] found that micro-structural and mechanical properties were strongly affected by variations in welding parameters within the chosen range of welding conditions. Under lower rotational and higher traverse speed in all welding conditions, the Wohler curves exhibited maximum fatigue strength.
Studying FSWed joints of components made of dissimilar 2219-T87/2195-T8 alloys (
Table 4), No et al. [
276] found that the WS only slightly affected the properties of the joint, but the latter strongly depended on the TRS.
Kopyscianski et al. [
277] obtained high weld quality for the process parameters including WS equal to 112 mm/min, TRS equal to 355 rpm and vertical force equal to 32.8 kN (
Table 4).
The proper FSWed joints of 5083-H12/ 6061-T6 alloys (
Table 4), studied by Ghaffarpour et al. [
278] were obtained under TRS values in the range of 700–2500 rpm, and WS values in the range of 25–400 mm/min.
Studying the underwater FSWed joints of dissimilar 6061/7075 alloys (
Table 4), Bijanrostami et al. [
279] found that the maximum tensile strength of 237.3 MPa and elongation of 41.2% were reached under a TRS 1853 rpm and a traverse speed of 50 mm/min. Thus, the WE reached up to 76.5%.
Studying the effect of the TRS to WS ratio (υ ratio) on the strength of the FSWed joints of dissimilar 6082-T6/5083-H111 alloys (
Table 4), Kasman et al. [
280] found that nugget zone profile containing onion rings of the shape depended on the value of the TRS and WS. These speeds varied also the effect of a constant υ ratio on the profile and structure of the nugget zone. At a lower TRS and WS lower UTS values occurred.
Studying FSWed joints of dissimilar 5083-H111/6351-T6 alloys (
Table 4), Palanivel et al. [
144] reported that a low WS and a high TRS enhanced the frictional heat because of the enhanced residing time of tool. TRS caused stirring and mixing of the material about the rotating pin, which in turn enhanced the temperature of the metal. Thus, the TRS strongly affected the WS. A low TRS providing a low heat input led to a lack of stirring and yielded defects. Contrary, during FSW, the enhancement in the TRS caused an increase in the heat input. More heat input destroyed the regular flow of plasticized material, and an enhanced TRS induced the excessive release of stirred materials to the upper surface, which left voids in the weld zone. The lowest and highest WS produced defects due to the increased frictional heat and insufficient frictional heat generated, respectively. The FSW at higher WSs caused a short exposure time in the weld area with insufficient heat and poor plastic flow of the metal and produced defects in the joints. Higher WSs causing low heat inputs provided faster cooling rates of the welded joint and hence yields defects.
The proper FSWed joints of 6 thick dissimilar alloys 6351/ and 5083-H111 alloys (
Table 4), studied by Palanivel et al. [
281] were obtained under TRS of 950 rpm, WS of 1.0 5mm/s and axial force 10kN.
Studying FSWed joints of components made of 5052-H32/6061-T6 blanks (
Table 4), Doley and Kore [
282] found that for all thicknesses, weld produced at 63 mm/min speed exhibited more ductility compared to that produced at 98 mm/min.
Saravanan et al. [
283] reported that the maximum tensile strength of 356 MPa appeared with the D/d ratio of 3, TRS of 1200 rpm, WS of 12 mm/min, and axial load of 8 kN (
Table 4).
Yan et al. [
283,
284] obtained proper FSWed joints of sheets made of dissimilar Al-Mg-Si/Al-Zn-Mg alloys (
Table 4) under TRS equal to 800 rpm and WS equal to 180 mm/min.
Studying FSWed joints of components made of dissimilar 2024-T3/6061-T6 alloys (
Table 4), Zapata et al. [
285] found that the enhancement of the TRS decreased the magnitude of the longitudinal residual stresses. This was due to the rise of heat input and the weakening of thermal mismatch between the different zones of the weld. The effect of the WS on the residual stress was small in comparison to the effect of the TRS, generating only a small rise in the profile of the retreating side when it was enhanced.
Studying butt FSWed joints of 2 mm thick plates including the one rolled from ultrafine-grained UFGed 1050 alloy and the one made of the 6061-T6 alloy (
Table 4), Sun et al. [
193] found that at various WSs, two fracture modes occurred for the tensed specimens depending on their revolutionary pitches. The FSWed joints were obtained under TRS of 800 rpm and WS values of 400, 600, 800, and 1000 mm/min.
From a second source of information, studying the butt FSWed joints of the ultrafine grained UFGed 1050 Al plates with a thickness of 2 mm with the 2 mm thick 6061-T6 alloy plates (
Table 4), Sun et al. [
286] reported that after welding process under a revolutionary pitch varying in range from 0.5 to 1.25 mm/rev, in the joint stir zone, the initial nano-sized lamellar structure of the UFGed 1050 Al alloy plate changed into an equiaxial grain structure with a greater average grain size as a result of the dynamic recrystallization and subsequent grain growth. An equiaxial grain structure with a lower grain size simultaneously appeared in the 6061 alloy plates, together with coarsening of the precipitates.
The proper FSWed joints of 2024-T6/ 6061-T6 alloys (
Table 4), studied by Sun et al. [
329] were obtained under TRS of 1000 rpm, and the WS of 500 mm/min.
During studies on butt FSWed joints of components made of dissimilar 6061 and 7050 alloys Rodriguez et al. [
289] found in joints microstructure occurred distinct lamellar bands and various degrees of intermixing affected by TRS. The joints consistently fractured on the 6061-alloy side. Two modes of failure existed, one through the stir zone and the other through the heat-affected zone. The inadequate material intermixing produced at low TRSs induced low mechanical strength and failure through the stir zone. The failure through the heat-affected zone at high TRSs occurred due to the material softening. Studying FSWed joints of components made of dissimilar 6061 to 7050 high strength Al alloys Rodriguez et al. [
288] found that the cyclic strain hardening and the number of cycles to failure enhanced with increasing the TRS.
The proper butt FSWed joints of dissimilar 6061/7050 alloys (
Table 4), studied by Rodriguez et al. [
288] and [
289] were obtained under TRS values of 270, 340, 310 rpm, and WS of 114 mm/min.
The lap FSWed joints of plates made of dissimilar 6111-T4/ 5023-T4 alloys (
Table 4), studied by Yoon et al. [
47] were obtained with a revolutionary pitch of 0.067 mm/rev, with an onion ring nugget with a rotation speed of 1500 rpm and a WS of 100 mm/min, and a revolutionary pitch of 0.7 mm/rev, with a void-defect nugget with a TRS of 1000 rpm and a WS of 700 mm/min.
Ilangovan et al. [
290] reported that all three pin profiles yielded the surface defect free joints with TRS of 1100 rpm and WS of 22 mm/min (
Table 4), as then the heat generation was almost the same for those tool pin profiles.
During studies on butt FSWed joints of components made of dissimilar 2050/ 6061 alloys (
Table 4). Reza-E-Rabby et al. [
291] found that quality welds can be produced at low rotational and travel speed. Flats could not produce defect free welds at the highest WS.
The proper FSWed joints of dissimilar 5083-O and 6082-T6 alloys (
Table 4), studied by Donatus et al. [
292] were obtained under TRS of 400 rpm, and WS of 400 mm/min.
The FSWed joints of dissimilar cast Al–Si alloys A319, A413 (
Table 4), studied by Karam et al. [
293] were obtained under rotational values of 630, 800, 1000 rpm, and WS values of 20, 40, 63 mm/min. The average size of the Si particles and of α-Al grains enhanced with a rise of the TRS and/or lowering of the WS. At the center of stirred zone, the Si particles were more uniformly distributed at low welding or high TRSs than in the case at higher welding or lower TRSs. The average hardness of the welded regions enhanced with the rise of the WS and/or lowering the TRS.
The butt FSWed joints of dissimilar 7075-O/6061-O and 7075-T6/6061-T6 alloys (
Table 4), studied by Ipekoglu and Cam [
294] where obtained under TRS values of 1000 and 1500 rpm, and WS values of 150 and 400 mm/min. The enhancing rotational rate raised the amount of the RS base material BM in the DXZ microstructure.
Cole et al. [
295] found that the highest joint strength was achieved at 700 rev/min spindle speed and 100 mm/min weld speed with 7075-T6 on the retreating side (
Table 4). The highest value of weld interface temperatures was obtained for the low tool travel speed value equal to 100 mm/mm.
During studies on lap FSWed joints of 5 mm-thick sheets made of dissimilar 2024-T3/ 7075-T6 alloys (
Table 4), Song et al. [
296] found that the hook deflects significantly upwards into the stir zone SZ at lower WSs in both combinations. The WS and joint combination affected the hook geometry which in return affected the lap shear strength. In both joint combinations, the lap shear strength raised with the enhancement of WS. 7075/2024 joints exhibited greater failure load than 2024/7075 joints at lower WSs while the opposite trend occurred at higher WSs. In case of 2024/7075 joints the WE varied in the range of 15-39% dependent on the WS in the range of 30-300 mm/min. The WE reached up to 57% under the WS of 150 mm/min for the case of 7075/2024 joint.
The proper FSWed joints of dissimilar 6061-T6/5083-0 alloys (
Table 4), studied by Jannet and Mathews [
297] were obtained under TRS values of 600, 750, 900, and WS of 60 mm/min.
The FSWed joints of 5083-H111 and 6351-T6 alloys studied by Palanivel et al. [
298] were obtained under TRS of 950 rpm, and for three values of WS including 36, 63, and 90 mm/min. The WS of 63 mm/min provided the best quality of welds.
The butt FSWed joints of dissimilar 2014-T6 and the 6061-T6 alloys studied by Jonckheere et al. [
300] were obtained under TRS values of 500, 1500 rpm, and WS of 90 mm/min. Welds obtained under a TRS of 500 rpm containing more 2014 alloy in their stirred zone or in contact with the tool shoulder were cooler and presented a narrower softened zone. Welds obtained under a TRS of 1500 rpm exhibited no effect of the tool shift or the alloys placement on their hardness profile.
Palantivel et al. [
299] reported that the joints exhibited better tensile strength using straight square pin profiled tool with TRS of 950 rpm, WS of 63 mm/min and axial force of 14.7 kN (
Table 4). The axial force acting on the tool most contributed on the UTS, and it was followed by tool pin profile, WS and TRS, for the range considered. The WE can reach up to 88.6%.
Studying FSWed joints of components made of dissimilar A356/ 6061 alloys (
Table 4), Ghosh et al. [
302] reported that with an increase in WS, matrix grain size became finer, without incessant limiting of Si-rich particles' size affected by interaction time between tool and substrate. The maximum joint efficiency of 116 % with respect to that of 6061 alloy occurred at an intermediate tool-traversing speed, providing fine matrix grain size and a small size of Si-rich particles.
The lap FSWed joints of 7075-T6/ 2198-T351 alloys (
Table 4), studied by Velotti et al. [
330] were obtained under a TRS of 830 rpm, and a WS of 40 mm/min. The WE was quite low in comparison to that obtainable the butt FSWed joints of the same alloys’ pair.
Studying FSWed joints of plates made of dissimilar 2219-T87/ 5083-H321 alloys (
Table 4), Koilraj et al. [
304] found WS strongly affected the joint soundness. The welds were obtained under TRS values in range of 400-800 rpm, and WS in range of 15-60 mm/min. The WE reached up to about 90%.
Investigating the FSWed dissimilar cast and wrought 6061 alloy (
Table 4), Dinaharan et al. reported [
305] that the material location prior to welding and TRS strongly affected the material flow behavior. The material in the advancing side occupied the major portion of the weld zone under enhanced TRS. The joint exhibited maximum tensile strength when cast Al alloy was positioned on the advancing side at all TRSs.
Studying of FSWed joints of components made of dissimilar 5083-H111/ 6351-T6 alloys (
Table 4), Palanivel et al. [
306] reported that the TRS and pin profile influenced the joint strength because of varying material flow, loss of cold work in the HAZ of 5083 side, dissolution and over aging of precipitates of 6351 side and formation of macroscopic defects in the weld zone. They found that an increase in TRS (from 600 to 1300 rpm) at a constant traverse speed of 60 mm/s, for various pin geometry, initially enhanced and then lowered weld effectiveness. The weld fabricated using TRS of 950 rpm and straight square pin profile reached its efficiency up to 88.6%.
Sivachidambaram et al. [
341] reported three various relations between TRS and joint efficiency for various WSs (40, 60, and 80 mm/min). The FSWed joints of dissimilar 5052-H34/ 5023-T4 alloys (
Table 4), studied by Song et al. [
127] were obtained under TRS of 1500 rpm, and WS values in range of 100-700 mm/min.
Ghosh et al. [
303] reported that tool rotation and traversing speed significantly affected the microstructure of welds. Welding at low TRS and WS caused the generation of fine grain size of 6061 alloy near the interface, limited residual thermal stress, lowered the extent of recovery–recrystallization, enhanced defect density, promoted the finer distribution of Si-rich particles and increased consolidation of transported material at the back of the tool to eliminate discontinuities within weld nugget. The welds fabricated at the lowest tool rotational and traversing speed exhibited the best mechanical properties. The 80 mm/min tool-traversing speed was optimal to achieve joint efficiency of ~116% with respect to that of 6061 Al alloy.
The FSWed joints of dissimilar 5052/ A5J32 alloys (
Table 4), studied by Kim et al. [
307] were obtained under TRS values of 1000, and 1500rpm, and WS values of 100, 200, 300, and 400 mm/min. In the case where A5J32 was fixed on the RS, the highest strength of the welded joints appeared under conditions of 1000 rpm and 300 mm/min. The WE reached up to about 94 %.
The proper FSWed butt joints of 7050-T7451/ 2024-T351 alloys (
Table 4), studied by Prime et al. [
308] were obtained under WS of 50.8 mm/min.
The correct FSWed joints of dissimilar 5182-O, 5754-O, and 6022-T4 alloys (
Table 4), studied by Miles et al. [
332] were obtained under rotation speed values in range of 500 to 1500 rpm, and WS values in range of 130 to 400 mm/min.
The proper butt FSWed joints of 6061-Al as itself, and of dissimilar 6061-T6/ 2024-T3 alloys (
Table 4), studied by Ouyang and Kovacevic [
333] were obtained under TRS values in range of 151-914 rpm, and WS values in range of 57-330 mm/min.
The butt FSWed joints of dissimilar 2024-T351/ 6056-T4 alloys (
Table 4), studied by Amancio-Filho et al. [
260] were obtained under TRS values in range of 500 to 1200 rpm and the WS values in range of 150 to 400 mm/min. Sound joints were obtained at TRS equal to 800 rpm and WS of 150 mm/min.
The proper FSWed joints of cast A356/wrought 6061 alloys (
Table 4), studied by Lee et al. [
261] were obtained under TRS of 1600 rpm, and WS values in range of 87 to 267 mm/min.
The correct FSWed joints of dissimilar 7003/7046 alloys (
Table 4), studied by Yang et al. [
312] were obtained under TRS of 2000 rpm and WS of 400 mm/min.
Mastenaiah et al. [
321] reported that defect free joints were produced with TRS 400 to 2000 rpm, tool offset position from -2mm to +2mm at WS of 30 mm/min (
Table 4) The joints produced with TRS in the range of 400-800 rpm, WS in the range of 30-390 mm/min and tool offset position from -2 mm to +1 mm also resulted in sound joints. The WE reached up to 97 %. Only welds obtained at the lowest TRS and highest WS and tool offset towards 219 alloy side exhibited defects. The extent of intermixing was affected by the TRS and WS. Intimate mixing of dissimilar alloys appeared at higher TRSs and lower WSs. The dissimilar joints FSWed under the conditions of high heat input (TRS varying from 400 to 2000 rpm) and lowest WS of 30 mm/min contained extensive intermixing in the nugget zone.
Studying the FSWed joints of rolled sheets made of dissimilar 2024/ 5056 alloys, Ivanov et al. [
334] found that the proper joints were obtained using lower linear WSs and high tool rotation speeds.
Investigating butt FSWed joints of components made of alloy 6063 and 5083 obtained for various TRSs in the range of 600-1200 rpm, 4 kN axial load and WS of 40 mm/min (
Table 4) Kumar et al. [
335] reported that the joints of higher tensile strength, lower flexural strength and lower impact strength with maximum hardness were fabricated at the TRS of 1000 rpm with a cylindrical profile. The flexural strength and impact strength lowered whereas the tensile strength and hardness enhanced with the rise of the tool's TRS.
Studying FSWed joints of 2618-T87/ 5086-H321 alloys (
Table 4), Sasikala et al. [
336] reported that the obtaining of the sound joints was affected, inter alia, by WS. The WE reached up to 90%.
Investigating the FSWed joints of components made of 2014-T6 alloy (
Table 4), Aydin et al. [
337] found that the hardness in the softened weld region lowered with a decrease in the WS. Independently of the TRSs, the best tensile and fatigue properties of the joints occurred under the WS of 80 mm/min. The WE varied in the range of 93-97 %
Studying the FSWed joints of components made of 3003-H12 alloy (
Table 4), Aydin et al. [
338] found that the tensile weld strength enhanced with an increase in the WS or a decrease in the rotation speed. The tensile fractures of the joints were in base metal under welding parameter combinations of 1070 rpm and 40 mm/min or 2140 rpm and 224 mm/min. All other joints failed at heat affected zone. The yield and ultimate tensile strengths of the joints lowered almost linearly with an enhancement of TRS at a constant WS, while such strengths of the joints enhanced almost linearly with a rise of WS at a constant TRS. The elongation values of FSW joints were smaller at higher TRS or lower WS.
Investigating the single-sided butt FSWed joints of 3 mm thick 3003-O non-heat-treatable Al alloy (
Table 4), Aydin et al. [
339] reported that the welding parameters strongly affected the fatigue behaviours of the 3003-O FS welds. The fatigue life of FS welds obtained under the WS of 40 mm/min at various rotating speeds were about 2 – 3 times longer compared to those of FS welds with the WSs of 80 mm/min and 112 mm/min at various TRSs at a fixed stress amplitude under the stress ratio R = −1. At a much lower WS and a higher TRS, the fatigue life of the joints was increased due to the enhanced amount of heat supplied to the weld per unit length.
Studying the FSWed joints of 4 mm thick plates made of 2024 alloy (
Table 4) without post-process heat treatment Weglowski et al. [
340] reported that the weldability of Al alloys used for FSW process was good and provided good quality of welded joints for a wide range of welding parameters. The kind of tool had no effect on joint properties at the same welding parameters.
Nejad et al. [
341] studied the structure and mechanical properties of FSWed joints of plates made of 2024-T4 alloy (
Table 4) with cylindrical outer and concave end surface shoulder and varied depth. Joints were obtained for two different tool designs including a threaded one and an unfeatured one. Obtaining a defect-free weld structure with both probe tools needed well different rotation and traverse speeds. TRS of 500 rpm, WS of 55 mm/min and plunge depth of 2.7 mm for threaded tool, and TRS of 1300 rpm, WS of 115 mm/min and plunge depth of 2.9 mm for unthreaded tool allowed obtaining the finest grain in stir zone, the best visual quality and smoothness, the highest tensile strength, elongation, and micro-hardness.
Studying the FSWed joints of 6 mm thick plates made of 2024-T351 alloy (
Table 4), Milčić et al. [
342] obtained the compounds without errors and with an acceptable flat surface under the constant TRS of 750 rpm, and the WS varied in range from 73 to 150 mm / min. The relation between the TRS and WS directly affected the fracture toughness and energy necessary for the initiation and propagation of the crack in the joint. The weld joint obtained under 750/116 rpm/(mm/min) exhibited better properties and microstructure than the joints obtained under conditions of 750/73 and 750/150 rpm/(mm/min), respectively. The WE of 97% was achieved under rotation speed of 750 rpm and WS of 116 mm/min.
Investigating the FSWed joints of 8 mm-thick plates made of 2014-T6 alloy (
Table 4), Lin et al. [
343]reported that the weld tensile strength was affected by welding parameters. The maximum ultimate tensile strength of 360 MPa equal to 78 % appeared at a TRS of 400 rpm and WS of 100 mm/min.
Studying the FSWed joints of components made of 2014 alloy (
Table 4), Sinhmar and Dwivedi [
17] reported that after welding process realized at TRS of 931 rpm and WS of 41 mm/min, the mechanical performance of 2014 lowered. Simultaneously the corrosion resistance of the weld joint was higher than that of the base metal.
Investigating the FSWed joints of components made of the 2014-T6 alloy (
Table 4) Ugender et al. [
344] reported that the defect-free welds were obtained at a TRS of 900 rpm, taper cylindrical tool pin profile and traverse speed of 40 mm/min, respectively.
Studying the FSWed joints of components made of the 2017-T351 alloy (
Table 4), Liu et al. [
57] found that the tensile properties and fracture locations of the joints strongly depended on the welding process parameters. Under the optimum revolutionary pitch of 0.07 mm/rev corresponding to the TRS of 1500 rpm and the WS of 100 mm/min, the maximum ultimate strength of the joints corresponded to 82% of that of the base material.
Investigating the FSWed joints of components made of 1050-H24 alloy (
Table 4), Liu et al. [
309] reported that a softened region located at the weld and heat affected zones appeared in the joints. The degree of softening and tensile properties of the joints depended strongly on WS and TRS. The optimum FSW parameters were affected both by the tensile properties and the welding parameters. They were obtained for WS of 200 mm/min and the TRS of 1000 rpm.
The FSWed joints of cast AlSi9Mg/ 2017A alloys (
Table 4), studied by Mroczka [
310] were obtained under TRS of 560 rpm, and WS of 1120 mm/min. The FSWed joints of 2017A alloy (
Table 4), studied by Mroczka et al. [
311] were obtained under TRS of 355 rpm, and WS of 280 mm/min. Studying the FSWed joints of sheets made of 2017A alloy Mroczka et al. [
345] found that a higher TRS (900 rpm then 355 rpm) the properties of the joint lowered. Cracks along grain boundaries and separation of grains at welds occurred at higher TRSs. They were due to the grain boundaries within joint nugget lost cohesion during the welding process at the high rate.
Takhakh and Abdullah [
346],compared fatigue properties of welded joints of plates made of 3003-H14 alloy (
Table 4) obtained by FSW (at TRS of 1500 rpm and WS of 80 mm/min) and TIG welding. They found that the fatigue properties of FSW joints were slightly lower than the base metal and higher than these of TIG welding.
Studying the FSWed joints of components made of 3003-alloy (
Table 4), Chekalil et al. [
347] reported that the joint mechanical properties were affected in the order by the TRS, feed rate and tool tilt angle. The best mechanical properties of a welded joint were obtained under the TRS of 1423.9 rpm, feed rate of 400 mm/min and tool tilt angle of 1.28.
Investigating the butt FSWed joints of plates made of 3003-H24 alloy (
Table 4), Kasman and Ozan [
348] reported that at a WS of 50 mm/min, tunnel-type defects with large size appeared in joints welded with the TRS of 500 and 1000 rpm. The tunnel-type defects also occurred under WS of 80 mm/min and TRS of 500 and 800 rpm. However, cavity-type defects occurred at both WSs. All welded joints fractured between the base metal and the heat-affected zone, except for the joints welded under WS of 50 mm/min WS, and TRS of 500 and 1000 rpm. The highest ultimate tensile strength among all the welded joints equal to 128 MPa was obtained under WS of 50 mm/min and TRS of 800 rpm. The welded joints were fractured in a ductile manner except the joint produced under WS of 50 mm/min and TRS of 500 rpm.
Studying the butt FSWed joints of 6013 Al plates (
Table 4) obtained via pin offset technique Kasman and Ozan [
313] found that the highest tensile strength equal to 206 MPa was obtained under 1.5 mm pin offset towards the advancing side and 500 min
−1 TRS, leading to the ratio of tensile strength of the joint to the ultimate tensile strength of the base metal (joint efficiency) equal to 74 %.
Kasman and Yenier [
322] reported that a defect-free joint was obtained under TRS of 1000 rpm, WS of 80 mm/min, and a 22-mm tool shoulder diameter (
Table 4). The UTS lowered with enhancement of WSWS or TRSTRS. The WE for 5754 alloy lowered with an enhancement in WS, however, can exceed 100% for some weld cases. As for 7075 alloy, the efficiency varied in the range of 23.3-41.9 %.
Xu [
349] studied FSWed joints of 5 mm thick plates made of 3003-H17 alloy (
Table 4) under WS of 1500 and 3000 mm/min and a constant TRS. They found that the UTS of weld joint lowered with an enhancement in the WS from 1500 mm/min to 3000 mm/min at a constant TRS of 2000 rpm and shoulder plunge depth of 0.2 mm. The WE reached 87 % at WS of 1500 mm/min.
Studying the FSWed joints made of 3003 alloy (
Table 4), Goyal et al. [
350] reported that the best UTS was obtained for the process parameters including a WS of 74.64 mm/min, a TRS of 971.77 rpm and a tool tilt angle of 1.52, respectively.
Janeczek et al. [
351] studied the effect of the shape of a tool and welding parameters on the quality of FSWed joints of components made of alloy 3004 (AlMn1Mg1) (
Table 4). Various butt joints were made with a cylindrical and tapered threaded tool with a TRS of 475 rpm. The other joints were obtained with TRS of 475 rpm and WS of 300 mm/min with the use of a cylindrical threaded pin. Most of the specimens were properly joined for TRS of 475 rpm. In the joints obtained under WS of 300 mm/min, the material was not stirred properly. The best joint quality appeared for a TRS of 475 rpm and various WS values between 150 and 475 mm/min. The WE widely varied in the range of 61.9-87.6%, however the individual cases of below 27% were also noticed.
Studying butt FSWed joints of components made of dissimilar 7020-T651/ 5083-H111 alloys (
Table 4), Torzewski et al. [
352] found that the FSWed samples obtained under TRS of 800 rpm and WS of 200 mm/min exhibited the best strength properties: UTS = 303 MPa, YS = 157 MPa, and A = 11.6 %. All joints obtained at WS of 100 mm/min reached the efficiency of 95%.
Choi et al. [
353] studied spot FSWed joints of sheets made of 5454 alloy (
Table 4) with the different thicknesses of 1.4 and 1.0 mm obtained under the TRS in the range from 500 to 2500 rpm, and plunging to the depth of 1.8 mm under a constant tool plunge speed of 100 mm/min. The rotating tool was maintained at the plunge depth during the dwell time ranging from 0 to 7 sec. The pull-out speed of the rotating tool was 100 mm/min. The enhancement of TRS changed the macrostructure of the friction-stir-spot-welded zone, especially the geometry of the welding interface.
Studying the FSWed joints of components made of 1100 alloy (
Table 4), Selvarajan and Balasubramanian [
354] reported that a maximum tensile strength of 105 MPa, hardness value of 67 HV, and minimum corrosion rate of 0.69*10
-4 in the stir zone region was obtained under the optimized parameters of 893 rpm TRS, 100 mm/min WS, 6.5 kN axial force, shoulder diameter of 14.8 mm, pin diameter of 4.9 mm, and tool material hardness of 45.4 HRC.
Dong et al. [
355] studied microstructure and mechanical properties of welded joints of components made of dissimilar 7003-T4/ 6060-T4 alloys (
Table 4), obtained by underwater friction stir weld UFSW. They reported that sound and defect-free joints were obtained in the UFSW process, however, tunnel defects appeared with a high WS of 240 mm/min. With the enhancement of the WS, more η and η′ phases remained because of the lower heat input.
Sheikhi and dos Santos [
356] studied the effect of welding parameters and welding tools on the weld quality and mechanical properties of FSWed joints of tailor welded blanks TWBs made of 6181-T4 alloy in a thickness combination 1 to 2 mm as-produced (
Table 4). Changing the WS had the biggest effect on the measured temperature and the heat input.
Zhou et al. [
357] studied FSWed joints of 6061-T6 alloy sheet obtained under the TRS of 11,000 rpm, and the WS varying from 200 mm/min to 500 mm/min. The sound joints were obtained under the travel speed of 300 mm/min. Due to the stirring effect of high TRS, the proportion of low angle boundaries in all zones was lower than that of the traditional FSW, while the average grain size was like traditional FSW. The WE reached up to 87.2%. Generally, for lap FSW process when the rotation speed was high and the WS was low, the weld nugget occurred on an onion ring shape, and when the rotation speed was low, and the WS was high, void-type defects appeared on the AS or center of the nugget [
315].
The FSWed joints of 6061-T6 alloy (
Table 4), studied by Juarez et al. [
316] were obtained at constant TRS of 1000 rpm, WS of 90 mm/min, penetration speed of 9 mm/min, and holding time of 10 seconds. The last parameters allowed lowering the defects of welding.
Godhani et al. [
317] obtained FSWed joints of dissimilar 6061/7075 alloys (
Table 4) under the WS of 31.5 mm/s, TRS of 765 rpm, and tool tilt angle of 2° forward. Various FSW process parameters for FSW joints of various Al alloys were presented by Patel et al. [
256].
Studying FSWed 5 mm- thick plates made of 5086-O/ and 6061-T6 alloys, Aval et al. [
358] reported that the enhanced TRS and lowered WS provided weaker welds and coarser grain size in the weld nugget.
Investigating the butt FSWed plates of 2219-T62 alloy, Xu et al.[
359] found that with the enhanced TRS, the longitudinal residual stress lowered on the top surface, but raised on the bottom surface.
Studying the FSWed joints of 5 mm- thick 7075AA7075-T651 and 2024AA2024-T351 similar and dissimilar alloys, Zhang et al. [
319] found that enhancing the TRS caused the widened TMAZ on the AS and RS. The mixing degree in the joints is remarkably affected by the TRS. Low TRS limited material mixing, while the typical onion ring of mixing pattern appeared at the high TRS. Compared to the base materials, significant grain refinement (average grain size: 1.7 μm) appeared at a TRS of 600 rpm. The enhanced TRS caused grain coarsening. The nugget zone of all the joints is dominated by a simple shear texture and varied with the TRS.
Sivachidambaram et al. [
341] studied butt FSWed joint of 6mm- thick components made of dissimilar 5383/7075 alloys using tool with shoulder diameter of 24 mm, square pin with diameter of 8 mm and length of 5.7 mm under TRS in the range of 700-900 rpm, weld speed in the range of 40-80 mm/min, tilt angle of 0 degree, and the axial load of 10 kN. They found that varying WS affected yield stress and the lower WS caused maximum yield stress. The TRS of 700 rpm and WS of 40 mm/min provided very high tensile strength and hardness.
The influence of TRS on mechanical properties and thus on joint efficiency is not clear due to it is strongly affected by other parameters, e.g., traverse speed and the type of materials joined. The optimization of welds’ quality very often considers the effect of both TRS and WS.
Table 4.
FSW process parameters for FSW joints of various Al alloys.
Table 4.
FSW process parameters for FSW joints of various Al alloys.
Refs |
Alloy Combinations |
Thick |
Rotation Speed |
Welding Speed |
Plunge depth |
Tool tilt angle |
Downward force |
|
|
[mm] |
[rpm] |
[mm/min] |
[mm] |
[°] |
[kN] |
262 |
2024-T351/5083-H112 |
6.35 |
600 |
150 |
|
|
|
262 |
7075-T651/2024-T351 |
6.35 |
600 |
150 |
|
|
|
264 |
7075-T651/2024-T351 |
6 |
900 |
150 |
|
|
|
61 |
7075-T6/2024-T3 Lap joint: 7075-upper; 2024-lower |
3 |
600 |
30, 60, 90, 120 |
0.2 |
2.5 |
|
265 |
7075-T651/5083-H111 |
6 |
280,355, 450, 560 |
140 |
|
|
26.4 |
266 |
2024-T4/7075-T6 |
4 |
1140 |
32 |
|
|
|
267 |
6351-T6/5083-H111 |
6 |
800,1000 1200 |
45, 60, 75 |
|
|
|
46,265 326 |
2017A-T451/7075-T651 |
6 |
355 |
112 |
|
1.5 |
32.8 |
327 |
5083-O/6063-T6 |
6 |
900 |
60 |
|
|
|
269 |
5052andAlMg2Si |
8 |
1000 |
80 |
|
2.5 |
|
270 |
2024-T351/6061-T6 |
6 |
800 |
31.5 |
|
2 |
|
271 |
6061-T6/6351-T6 |
6.35 |
600,900, 1200 |
30, 60, 90 |
|
|
|
272,328 |
6082-T6/7075-T6 |
10 |
950, 1000 |
80, 100 |
|
|
|
273 |
6061-T651/5A06-H112 |
5 |
600, 900, 1200 |
100, 150 |
4.7 |
2 |
|
274 |
6101-T6/6351-T6 |
12 |
900,1100, 1300 |
16 |
|
|
|
275 |
2024-T3/6063-T6 |
8 |
900,1120 1400 |
125, 160, 200 |
|
2.5 |
|
276 |
2219-T87/2195-T8 |
7.2 |
400, 600, 800 |
120, 180, 240, 300 |
|
|
|
277 |
2017A-T451/cast AlSi9Mg |
6 |
355 |
112 |
|
|
|
278 |
5083-H12/6061-T6 |
1.5 |
700,1800 2500 |
25, 30, 212.5, 400 |
|
|
|
279 |
6061-T6/7075-T6 |
5 |
1000, 1375, 1750, 2125, 2500 |
50, 125, 200, 275, 350 |
0.2 |
3 |
|
280 |
5083-H111/6082-T6 |
5 |
400,500, 630, 800 |
40, 50, 63, 80 |
|
2 |
|
144 |
5083-H111/6351-T6 |
6 |
800-1200 |
45-85 |
|
1 |
15 |
281 |
5052/6061 |
1, 1.5 |
1500 |
63, 98 |
|
|
|
282 |
2024-T6/ 7075-T6 |
5 |
1200 |
12 |
|
|
8 |
283,284 |
Al-Mg-Si/Al-Zn-Mg |
15 |
800 |
180 |
0.2 |
2.5 |
|
285 |
2024-T3/6061-T6 |
4.8 |
500, 650, 840 |
45, 65 |
|
2 |
|
286 |
UFG 1050/6061-T6 |
2 |
800 |
400, 600, 800, 1000 |
|
3 |
8 |
329 |
2024-T6/ 6061-T6 |
4 |
1000 |
500 |
|
2.5 |
|
288,289 |
6061-T6/7050-T7451 |
5 |
270, 340, 310 |
114 |
|
|
|
47 |
6111-T4/5023-T4 Lap joint |
1 |
1500 1000 |
100 700 |
|
|
|
290 |
5086-O/6061-T6 |
6 |
1100 |
22 |
|
1 |
12 |
291 |
2050-T4/6061-T651 |
20 |
150, 300, 300 |
101, 203, 406 |
|
|
|
292 |
5083-O/6082-T6 |
NR(~7) |
400 |
400 |
|
|
|
293 |
A319/ A413 cast |
10 |
630, 800, 1000 |
20, 40, 63 |
1 |
3 |
|
294 |
7075-O/6061-O 7075-T6/6061-T6 |
3.17 |
1000 1500 |
150 400 |
|
|
|
295 |
6061-T6/7075-T6 |
4.6 |
700-1450 |
100 |
|
|
|
296 |
2024-T3/7075-T6 Lap joint |
5 |
1500 |
50, 150, 225, 300 |
0.2 |
2.5 |
|
297 |
5083-O/6061-T6 |
6 |
600, 750, 900 |
20, 40 |
|
|
|
298 |
6351-T6/5083-H111 |
6 |
950 |
36, 63, 90 |
|
|
|
300 |
2014-T6/6061-T6 |
4.7 |
500, 1500 |
90 |
|
|
|
299 |
6351-T6/5083-H111 |
6 |
600-1300 |
36-90 |
|
|
9.8, 12.25, 14.7, 17.18, 19.6 |
302 |
A356/6061-T6 |
3 |
1000 |
70-240 |
|
3 |
|
330 |
2198-T351/7075-T6 Lap joint |
3 & 1.9 |
830 |
40 |
|
2 |
|
304 |
2219-T87/5083-H321 |
6 |
400-800 |
15-60 |
|
|
|
305 |
6061 cast/6061 rolled |
6 |
800,1000 1200, 1400 |
50 |
|
|
8 |
306 |
6351-T6/5083-H111 |
6 |
600, 950, 1300 |
60 |
|
0 |
8 |
127 |
5052-H34/5023-T4 |
~1.5 |
1500 |
100-700 |
|
3 |
|
303 |
A356/6061-T6 |
3 |
1000, 1400 |
80, 240 |
|
|
|
307 |
5052-H34/5023-T4 |
1.5 & 1.6 |
1000, 1500 |
100, 200, 300, 400 |
|
3 |
|
308 |
7050-T7451/2024-T351 |
25.4 |
NA |
50.8 |
|
|
|
332 |
5182-O/5754-O 5182-O/6022-T4 5754-O/6022-T4 |
~2 |
500, 1000, 1500 |
130, 240, 400 |
|
|
|
333 |
6061-T6/2024-T3 |
12.7 |
151-914 |
57-330 |
|
|
|
260 |
2024-T351/6056-T4 |
4 |
500-1200 |
150-400 |
|
|
|
261 |
cast A 356/wrought 6061 |
4 |
1600 |
78-267 |
|
3 |
|
312 |
7003-T4/7046-T4 |
3 |
2000 |
400 |
0.3 |
2.5 |
|
321 |
2219-T6/5083-H116 |
5 |
400, 800, 1200, 1600, 2000 |
30, 210, 390, 570, 750 |
|
|
|
335 |
6063/5083 |
6 |
600,800, 1000 |
40 |
|
|
4 |
336 |
2618-T87/5086-H321 |
6 |
450,600, 750, 850 |
15, 35, 50, 65 |
|
|
|
337 |
2014-T6 |
3 |
1070, 1520, 2140 |
40, 80, 112 |
|
|
|
338 |
3003-H12 |
3 |
1070, 1520, 2140 |
40, 80, 112, 160, 224 |
|
|
|
339 |
3003-O |
3 |
1070, 1520, 2140 |
40, 80, 112 |
|
|
|
340 |
2024-T4 |
4 |
350 |
210 |
|
|
|
341 |
2024-T4 |
3 |
300-1300 |
40-145 |
2.7, 2.9 |
6 |
|
342 |
2024-T351 |
6 |
750 |
73, 116, 150 |
|
|
|
343 |
2014-T6 |
8 |
300-800 |
50-300 |
|
|
|
17 |
2014 |
NA |
931 |
41 |
|
|
|
344 |
2014-T6 |
5 |
900 |
40 |
|
2.5 |
5 |
57 |
2017-T351 |
5 |
1500 |
25-600 |
|
3 |
|
309 |
1050-H24 |
5 |
600-2000 |
100-800 |
|
3 |
|
310 |
2017A-T451 /AlSi9Mg |
6 |
560 |
1120 |
|
1.5 |
|
311 |
2017A |
6 |
355, 900 |
280 |
|
1.5 |
|
346 |
3003-H14 |
3 |
1500 |
80 |
|
|
|
347 |
3003 |
2 |
1000, 1500, 2000 |
200, 300, 400 |
|
|
|
348 |
3003-H24 |
3 |
500,800, 1000 |
50, 80 |
|
|
|
313 |
6013-T6 |
5 |
500, 630, 800 |
50 |
|
|
|
322 |
5754-H111/7075-T651 |
5 |
1000, 1250 |
80, 100, 125 |
|
|
|
349 |
3003-H17 |
5 |
2000 |
1500, 3000 |
0.2 |
|
|
350 |
3003 |
5 |
663,800, 1000, 1200, 1336 |
20, 40, 70, 100, 120 |
|
0.65, 1, 1.5, 2, 2.35 |
|
351 |
3004 |
5 |
95-600 |
115-925 |
|
|
|
352 |
5083-H111/7020-T651 |
5 |
400,800, 1200 |
100, 200, 300 |
|
|
|
353 |
5454-O |
1, 1.4 |
500-2500 |
100 |
|
|
|
354 |
AA1100 |
5 |
562,700, 800,900, 1037 |
40.54, 75, 100, 125, 159.5 |
|
|
3.62, 5, 6, 7, 8.38 |
355 |
7003-T4/6060-T4 |
4.5 |
1000 |
40, 120, 240 |
|
|
|
315 |
6111-T4/5023-T4 |
1 |
1000, 1250, 1500 |
100, 300, 500, 700 |
|
|
|
356 |
6181-T4 |
1, 2 |
1300, 1600, 2000 |
800, 1000, 1125, 1500 |
|
1.5, 3 |
4.5, 5.5 |
357 |
6061-T6 |
1 |
11000 |
200-500 |
0.05 |
|
|
316 |
6061-T6 |
9.5 |
1000 |
90 |
|
|
|
317 |
6061-T6/7075-T6 |
6 |
765 |
31.5 |
|
2 |
|
4.8.1.3. Tool Geometry
Zhou et al. [
357] reported that the geometry of the shoulder and the pin profile strongly affected heat generation and material flow during welding process. The high shoulder size governed a heat input. The common shoulder profiles included the flat, the concave and the convex. The pin features such as a spiral or a groove improved frictional behavior and material flow in joint. Threads guided material flow around the pin in a rotational and a translation direction [
264,
265,
290,
360]. The polygonal pin profiles provided pulses in the flow during material stirring and mixing, causing material adhering to the pin [
361,
362,
363,
364].This pulsating effect highly impeded material flow in joints between dissimilar Al alloys. Thus, the cylindrical or a conical pin profile with various features provided for good material flow leading to obtain sound joints between the dissimilar Al alloys.
Hasan et al. [
261] studied effect of welding tool pin flute radius during the FSW process of dissimilar 7075-T651/2024-T351 alloys (
Table 5) on the weld quality. Five pin tools with various flute radii (0, 2, 3, 6, and ∞ mm) were investigated. They found that the flute radius affects the material flow pattern and weld quality. The strongest joint was obtained for the flute radius equal to the pin radius. The joint efficiency reached up to 94.3%.
Ge et al. [
61] studied how EST affects the shear failure load of lap joints. Shear fracture mode occurred in lap joints obtained with a small (3 mm) pin at all WSs. The tensile fracture mode appeared for the lap joints obtained with greater (4 mm or 5 mm) pin. Studying FSWed joints of components made of alloys 7075-T651 and 5083-H111 Kalemba-Rec et al. [
265] reported that the use of the Triflute pin provided greater tensile strength and WE.
Palanivel et al. [
267] studied the effect of shoulder profiles on the 5083-6351 combination properties (
Table 5). They utilized three different shoulder features including the partial impeller, the full impeller, and the flat grove. The full impeller shoulder tool provided the optimum mechanical strength due to the increased material flow in joint. The pin profile strongly influences material stirring and mixing. Cylindrical or conical pin profiles without threads provided a smaller surface to the material. Such pin profiles with the threaded and flat features enhanced the contact area.
For study on FSWed sheets made of dissimilar 2017A-T451/7075-T651 alloys (
Table 5), Hamilton et al. [
326] obtained qualitied welds using a tool made of HS6-5-2 high speed steel with a scrolled shoulder with a 24 diameter. The pin diameter tapered linearly from 6 mm at the shoulder to 4.5 mm at the tip with an overall height of 5.7 mm. The pin was also threaded.
Gupta et al. [
327] conducted studies on FSWed joints of components made of dissimilar 5083-O/ 6063-T6 alloys (
Table 5) focusing on optimization of tool geometry, TRS, and WS. The multi-optimal set of weld properties comprising tensile strength, average hardness at weld nugget zone and average grain size at weld nugget zone was obtained for 18 mm of shoulder diameter and 5 mm of pin diameter.
The proper FSWed joints of Al-Mg2Si/Mg
2Si5052 alloys studied by Huang et al. [
269] were obtained using an H13 steel-made pin tool comprising a concave 18-mm-diameter shoulder and a conical pin (the end and root diameter are 4 and 6 mm, respectively) with a pin length of 5.7 mm.
The correct FSWed joints of dissimilar 2024/ 6061 alloys (
Table 5), studied by Moradi et al. [
270] were obtained using AISI H13 hot work steel tool possessing a conical geometry with 18 mm shoulder diameter, a 4° conical cavity, a square frustum probe measuring 3.5–7 mm in diameter, and 5.9 mm in length.
The proper FSWed joints of dissimilar 6351-T6/ 6061-T6 alloys (
Table 5), studied by Prasanth and Raj [
271] were obtained using cylindrical tool having a scroll with 0.75 mm taper at the tip of the pin, and it has 16 mm probe diameter, 14 mm shoulder diameter, 5 mm pin length and of 4 mm pin diameter, made of molybdenum M42 with HRC 63.
The correct FSWed joints of dissimilar 6061-T651/ 5A06-H112 alloys (
Table 5), studied by Peng et al. [
273] were obtained using tool with cylindrical shoulder diameter of 16 mm, and conical pin with diameter varying from 5 to 4.2 mm, and the length of 4.6 mm.
The proper FSWed joints of dissimilar 2024-T3/ 6063-T6 alloys (
Table 5), investigated by Sarsilmaz [
275] were obtained using D5 steel tool with conical triangular pin profile quenched-tempered to 60 HRC.
The sound FSWed joints of dissimilar 2219-T87/ 2195-T8 alloys (
Table 5), studied by No et al. [
276] were obtained using the tool made of austenitized H13, with a spiral shape and a shoulder diameter of 16 mm.
The proper FSWed joints of dissimilar wrought 2017A/ cast AlSi9Mg alloys (
Table 5), studied by Kopyscianski et al. [
277] were obtained using a modified Whorl-type tool made of HS6-5-2 high speed steel with a 24 mm diameter and scrolled shoulder. The threaded pin diameter tapered linearly from 6 mm at the shoulder to 4.5 mm at the tip with an overall height of 5.7 mm.
The correct FSWed joints of 5083-H12/ 6061-T6 alloys (
Table 5), studied by Ghaffarpour et al. [
278] were obtained using tool with shoulder diameter in range 10–14 mm, and pin diameter in range 2–4 mm. The FSWed joints of studied by Bijanrostami et al. [
279] were obtained using tool made from 2344 steel heat-treated to obtain hardness by 52 HRC, and comprised the shoulder with diameter of 15 mm, and a threaded conic pin with conic angle of 5°, the length of 4.7 mm, and a diameter of 5 mm.
Studying the FSWed joints of dissimilar 6082-T6/ 5083-H111 alloys (
Table 5), Kasman et al. [
280] found that the pin shape significantly affected the tensile properties and microstructure of weld joints. The strengths of the weld joint obtained with the pentagonal-shaped pin were lower than those with triangular-shaped pin. The pin shape influenced each nugget zone profile containing onion rings.
Palanivel et al. [
144] obtained proper FSWed joints of the 6-mm-thick sheets made of dissimilar 5083-H111/ 6351-T6 alloys (
Table 5) using the straight square tool pin with a shoulder diameter of 18 mm, a pin diameter of 6 mm and a pin length of 5.6 mm.
The proper FSWed joints of 6 mm-thick components made of dissimilar alloys 6351/ and 5083-H111 alloys (
Table 5), studied by Palanivel et al. [
281] were obtained using five tool pin profiles including straight cylinder, threaded cylinder, square, tapered square, and tapered octagon without draft. The ratio of shoulder diameter and pin diameter was of 3. The tool material was HCHCr steel oil hardened to obtain a hardness of 60–62 HRC. The joint obtained using a tapered square pin profiled tool provided the least tensile strength. Using straight cylinder, threaded cylindrical, tapered square and tapered octagon pin profiled tools such a strength varied insignificantly. It resulted from the difference in dynamic orbit created by the eccentricity of the rotating tool during the FSW process.
Studying FSWed joints of components made of the dissimilar 2024-T6/ 7075-T6 alloys (
Table 5), Saravanan et al. [
283] found that the joint fabricated under D/d ratio equal to 3 showed better mechanical properties in comparison to other joints.
Yan et al. [
284,
285] obtained proper FSWed joints of sheets made of dissimilar Al-Mg-Si/Al-Zn-Mg alloys (
Table 5) using tool with shoulder diameter of 35 mm, and pin with diameters of the pin root and pin bottom equal to 20 and 12 mm, respectively, while pin length was of 14.5 mm.
The correct FSWed joints of dissimilar 2024-T3/ 6061-T6 alloys (
Table 5), studied by Zapata et al. [
286] were obtained using a tool consisted of a 20 mm diameter concave shoulder with a 4 mm diameter tapered threated pin.
The proper butt FSWed joints of UFGed 1050/ 6061-T6 alloys (
Table 5), studied by Sun et al. [
287] were obtained using the rotating tools made of tool steel, containing a concave-shaped shoulder with a diameter of 12 mm and a threaded pin with a diameter of 4 mm and a length of 1.8 mm.
Sun et al. [
329] studied the influence of various shapes of tool pin including conical thread, deep groove thread, and conical cam thread on the plastic flow of 2024-T6/6061-T6 alloys (
Table 5) during FSW process. They found that the metal in the weld nugget zone WNZ came from the base metal of the advancing side, the thread was the driving force of the downward movement of the FSW plastic metal. The deep groove thread tool pin strongly drove the metal downward. The conical cam thread tool pin provided the strongest stirring of materials and the best metal fluidity. Welds were obtained using tool with shoulder with concentric circles and diameter of 18 mm, and pin with diameter varied from 7 to 5 mm and length of 3.7 mm.
The proper FSWed joints of dissimilar 6061/7050 high strength Al alloys (
Table 5), studied by Rodriguez et al. [
288] and [
289] were obtained using a tool consisted of a cylindrical threaded pin and a shoulder having a diameter of 10 mm and of 18 mm respectively.
The sound lap FSWed joints of plates made of dissimilar 6111-T4/5023-T4 alloys studied by Yoon et al. [
40] were obtained using a tool with a shoulder diameter of 8 mm, and a threaded pin with diameter of 3 mm, and length of 1.45 mm. During studies on FSWed components made of the heat treatable 6061 and non-heat treatable 5086 alloys (
Table 5), Ilagovan et al. [
290] found that the use of threaded tool pin profile provided better flow of materials between two alloys and the generation of defect-free stir zone. It also allowed obtaining higher hardness values of 83 HV in the stir zone and higher tensile strength of 169 MPa compared to those of the other two pin profiles.
During studies on butt FSWed joints of components made of dissimilar 2050/6061 alloys (
Table 5), Reza-E-Rabby et al. [
291] found that joint quality, process parameters and welding temperature depended on pin features. Pin with thread flats allowed production of quality welds in some cases.
The proper FSWed joints of dissimilar 5083-O/ 6082-T6 alloys (
Table 5), studied by Donatus et al. [
292] were using the tool with a diameter to length ratio was 1 : 0.8 with a 25 mm diameter scroll shoulder applicable at a tilt angle of 0°.
The correct FSWed joints of dissimilar cast Al–Si alloys A319/A413 (
Table 5), studied by Karam et al. [
293] were obtained using tool with shoulder diameter of 26 mm, and conical threaded pin with diameters varying from 10 to 6 mm, and length of 9 mm.
The proper butt FSWed joints of dissimilar 7075-O/6061-O and 7075-T6/6061-T6 alloys (
Table 5), studied by Ipekoglu and Cam [
294] where obtained using tool with a concave shoulder diameter of 15 mm, and a M4 threaded cylindrical pin with diameter of 4 and length of 3 mm.
The correct FSWed joints of dissimilar 6061-T6/7075-T6 alloys (
Table 5), studied by Cole et al. [
295] were obtained using a tool with a 4.4 ° concave shoulder with diameter of 15 mm, and a threaded, conical tapered pin with three flats with diameter varying from 7.0 mm to 5.2 mm, and the pin length of 4.7 mm.
The proper lap FSWed joints of dissimilar 2024-T3/7075-T6 alloys (
Table 5), studied by Song et al [
296] were obtained using a tool with a 15 mm diameter concave shoulder and a 6 mm long threaded taper cylindrical pin with the top and bottom diameter of 4 mm and 6 mm, respectively.
The sound butt FSWed joints of dissimilar 2014-T6/ 6061-T6 alloys (
Table 5), studied by Jonckheere et al. [
300] were obtained using tool with a 15 mm diameter scrolled shoulder and threaded pin with three flats with a diameter of 5 mm and length of 4.4 mm.
Palantivel et al. [
299] for welding utilized tools with shoulder to work piece interference surface with 3 concentric circular equally spaced slots of 2 mm in depth on all tools. The tools utilized also five pin configurations including straight square, tapered square, straight hexagon, straight octagon, and tapered octagon without draft. The most helpful was straight square pin profiled tool.
The proper FSWed joints of dissimilar A356/6061 alloys (
Table 5), studied by Ghosh et al. [
302] were obtained using a tool made of high-speed steel with concave shoulder diameter of 15 mm, and cylindrical pin with diameter of 5 mm and length of 2.6 mm.
The correct FSWed lap joints of 7075-T6/2198-T351 alloys (
Table 5), studied by Velotti et al. [
330] were obtained using tool with shoulder diameter of 15.5 mm, and conical pin with maximum diameter of 4 mm and length of 3.1 mm.
Studying FSWed joints of plates made of dissimilar 2219-T87/5083-H321 alloys (
Table 5), Koilray et al. [
304] found the ratio between tool shoulder diameter and pin diameter the most dominated the joint soundness while pin geometry also strongly influenced it. Welds were obtained using tools with a pin with length of 5.7 mm and diameter of 6 mm. The ratio between tool shoulder diameter and pin diameter took values 1.5, 2, 2.5, 3 increasing with TRS and WS.
The proper FSWed joints of dissimilar cast and wrought 6061 alloy (
Table 5), investigated by Dinaharan et al. [
305] were obtained using a tool with a shoulder diameter of 19.2 mm, and a hexagonal pin profile with diameter of 6 mm and length of 5.8 mm.
Studying of FSWed joints of components made of dissimilar 5083-H111/6351-T6 alloys (
Table 5), Palanivel et al. [
306] reported that the TRS and pin profile influenced the joint strength because of varying material flow, loss of cold work in the HAZ of 5083 side, dissolution and over aging of precipitates of 6351 side and formation of macroscopic defects in the weld zone. Weld were obtained using tool with flat shoulder diameter of 18 mm, and pins with diameter of 6 mm, length of 5.7 mm, and various profiles including straight square, straight hexagon, straight octagon, tapered square, and tapered octagon. Square pins produced highly intense pulses which long last compared to those for hexagon and octagon pins causing severe and random layer-by-layer material movement.
The proper FSWed joints of dissimilar 5052-H34/5023-T4 alloys (
Table 5), studied by Song et al. [
127] were produced using a left-handed threaded tool with shoulder diameter of 12 mm, and a pin with diameter of 3.8 mm, and length of 1.45 mm.
The correct FSWed joints of dissimilar A356/6061 alloys (
Table 5), studied by Ghosh et al. [
303] were obtained using a tool with concave shoulder diameter of ~15 mm, and a cylindrical pin with diameter of ~5 mm, and length of ~2.6 mm.
The proper FSWed joints of dissimilar 5052 and A5J32 alloys (
Table 5), studied by Kim et al. [
307] were produced using a tool with shoulder diameter of 8 mm, and the threaded cylindrical pin with diameter of 3 mm and length of 1.45 mm.
The correct FSWed butt joints of 7050-T7451/2024-T351 alloys studied by Prime et al. [
308] were obtained using tool with threaded pin.
The sound FSWed joints of dissimilar 5182-O, 5754-O, and 6022-T4 alloys (
Table 5), studied by Miles et al. [
332] were produced using tool with a concave shoulder with diameter of 10.2 mm, and the cylindrical threaded pin with diameter of 3.18 mm, and length of 1.95 mm.
The proper butt FSWed joints of 6061-Al as itself, and of dissimilar 6061-T6/2024-T3 alloys (
Table 5), studied by Ouyang and Kovacevic [
333] were obtained using tool with threaded pin.
The sound butt FSWed joints of dissimilar 2024-T351/ 6056-T4 alloys (
Table 5), investigating by Amancio-Filho et al. [
260] were produced using tool with 5 mm diameter threaded cylindrical pin and 15 mm concave shoulder.
The correct FSWed joints of cast A356/wrought 6061 alloys (
Table 5), studied by Lee et al. [
261] were obtained using tool with screw-like pin.
The proper FSWed joints of the 2017-T351 alloy studied by Liu et al. [
57] and the FSWed joints of 1050 - H24 alloy (
Table 5), studied by Liu et al. [
309] were produced using tool with shoulder diameter of 15 mm, and pin with diameter of 5 mm and length 4.7 mm.
The correct FSWed joints of cast AlSi9Mg/ 2017A alloys (
Table 5), studied by Mroczka and these [
310] were of 2017A alloy studied by Mroczka et al. [
311] were obtained using tool with shoulder diameter of 22 mm, and cylindrical threaded pin with diameter of 8 mm.
The various FSWed joints of a 6061-T6 alloy (
Table 5), studied by Zhou et al. [
357] were obtained using tools with three configurations of pin including a quadrangular prism, quadrangular frustum pyramid, and frustum one. When the shape of the pin was a quadrangular frustum pyramid, sound joints are obtained.mm.
Studying butt FSWed joints of components made of 6063/5083 alloys (
Table 5), Kumar and Kumar [
335] reported that the joints of higher tensile strength, lower flexural strength and lower impact strength with maximum hardness were fabricated at the tool with a cylindrical profile.
Sheikhi and dos Santos [
356] studied the effect of welding parameters and welding tools on the weld quality and mechanical properties of FSWed joints of tailor welded blanks TWBs made of 6181-T4 alloy (
Table 5) in a thickness combination 1 to 2 mm. The peak temperature during welding slightly increased with increasing pin diameter. The effect of shoulder type on such peak temperature was negligible .
Studying FSWed joints of 2618-T87/ 5086-H321 alloys (
Table 5), Sasikala et al. [
336] reported that the obtaining of the sound joints was affected by the fraction of tool contact area to pin diameter, and to a lower extent by pin shape.
Investigating the FSWed of 4 mm thick plates made of alloy 2024 Weglowski et al. [
340] reported that the joints welded with the different tools and under various conditions exhibited a characteristic shape of a nugget zone, heat-affected zone and thermo-mechanically affected zone. The kind of tool had no effect on joint properties at the same welding parameters. The WE was of 91.1% for the Triflat tool with the flat bottom pin, while it was of 95,7% for the Triflute tool with the round bottom pin.
Nejad et al. [
341] studied the structure and mechanical properties of FSWed joints of plates made of 2024-T4 alloy (
Table 5) with cylindrical outer and concave end surface shoulder and varied depth. Joints were obtained for two different tool designs including a threaded one and an unfeatured one. Obtaining a defect-free weld structure with both probe tools needed well different rotation and traverse speeds. Despite increasing the elongation and strength properties of joints obtained with the threaded tool, they exhibited elevated average hardness and less uniform properties over various welding zones in comparison to the joints prepared by unfeatured tool.
Studying the FSWed joints of components made of the 2014-T6 alloy (
Table 5), Ugender et al. [
344] reported that the defect-free welds were obtained using taper cylindrical tool pin profile. The joints fabricated at a taper cylindrical tool profile with a 3 mm radius of curvature exhibited better mechanical properties compared to the straight cylindrical tool profile. The WE was of 69.5% for the taper cylindrical tool, while it was of 63.4% for the straight cylindrical one.
Burek et al. [
365] studied tool wear effect on the quality of lap FSWed joints of Al7075-T6 alloy sheets for two thicknesses (
Table 5). They explained that due to the small diameter of the pin and the great forces occurring in the process, this element was most susceptible to tool wear. The welding process caused the tool to undergo friction wear, resulting in lowered tool dive depth in the jointed material. After creating 200m of joints, the strength of the joints lowered and the changes in the stirring conditions in the material became more intensive. The degradation of the tool led to lowering the characteristic sizes of the thermoplastic zone strongly affecting the joint strength.
The FSWed joints of 3003 alloy (
Table 5), studied by Chekalil et al. [
347] were obtained using tool with flat shoulder with diameter of 19.5 mm, and conical pin with diameter varying from 6.8 to 5 mm, and length of 1.7 mm. The FSWed joints of 3003-H17 alloy studied by Xu [
349] were obtained using tool with shoulder diameter of 16 mm, and conical threated pin with maximal diameter of 6 mm, length of 4.7 mm, and taper angle of 2.5°. Similar joints studied by Goyal et al. [
350] were obtained using tool with concave shoulder with diameter of 18 mm, and square pin with diameter of 6 mm, and length of 4.75 mm.
Janeczek et al. [
351] studied the effect of the shape of a tool and welding parameters on the quality of FSWed joints of components made of 3004 alloy (
Table 5).Various butt joints were made with a self-developed tool with cylindrical threated and tapered threaded pin. They found that the material outflow for the joints made with the cylindrical threaded pin was higher than that for the joints made with the tapered threaded pin. However, voids-like defects appeared in the joints made with the tapered threaded tool. The use of the cylindrical tool provided higher for about 37% mechanical properties compared with those for the tapered threaded joint.
Studying the FSWed joints of components made of 1100 alloy (
Table 5), Selvarajan and Balasubramanian [
354] found that optimized parameters of welding process comprised shoulder diameter of 14.8 mm and, pin diameter of 4.9 mm, and tool material hardness of 45.4 HRC .
The proper FSWed joints of 6061-T6 alloy (
Table 5), studied by Juarez et al. [
316] were obtained using tool with flat shoulder diameter of 25.4 mm, composite pin with a hexagonal shank with a maximum diameter of 8 mm and a length of 9 mm and a cylindrical collar with a diameter of 11 mm and a height of 3 mm made of H13 steel.
Khan et al. [
366] studied the influence of tool pin offset and tool plunge depth on the formation of defects such as tunnel (tunneling defect) and kissing bond (KB) in 4.75 mm thick FSWed plates made of AA5083-H116 and AA6063-T6 alloys. The joints were obtained using tool made of Tungsten carbide, with tapered conical pin, shoulder diameter of 20 mm, pin length of 4.4 mm under TRS of 450 rpm, WS of 100 mm/min, tilt angle of 2 degrees, tool offset in range of 0.5-1.5 mm (AS to RS) and plunge depth of 0.3/0.4 mm. They found that the tunneling defects appeared at all offset (including zero offset) values towards stronger material (AS). The cross-section of the tunnel varied with the amount of offset. KBs appeared at the interface for all pin offset values except 0.5 mm towards softer material and high plunge depth causing the poor mechanical properties. Therefore, both plunge depth and tool pin offset strongly affected the weld quality. The plunge depth provided heat generation and the control over forging action and welding thrust. Tool pin offset distributed the heat and mixed the joined alloys.
Table 5.
Tools parameters used for FSW process for joining various Al alloys.
Table 5.
Tools parameters used for FSW process for joining various Al alloys.
Refs. |
Alloy Combinations |
Thick (mm) |
Tool profile (-) |
Shoulder diameter |
Pin diameter/ length/ taper angle |
Tool material hardness |
Shoulder |
Pin |
[mm] |
[mm]/ [mm]/ [°] |
[HRC] |
264 |
7075-T651/2024-T351 |
6 |
concave |
conical threaded and with flute radius (0, 2, 3, 6, and ∞mm)) |
18 |
6/5.7 |
AISI H13 |
61 |
7075-T6/ 2024-T3 Lap joint: 7075-upper; 2024-lower |
3 |
concentric-circles- flute |
tapered |
13.5 |
6/3,4,5/16.7 |
|
265 |
7075-T651/5083-H111 |
6 |
spiral (convex scrolled) |
triflute, tapered with a thread |
24 |
10/5.8; 10 (6 on tip)/5.8 |
HS 6-5-2 |
267 |
6351-T6/ 5083-H111 |
6 |
partial impeller, full impeller, flat grove |
cylindrical or conical with and without threads |
|
|
|
46 268,, 326 |
2017A-T451/7075-T651 |
6 |
scrolled |
tapered threaded |
24 |
6-4.5/5.7 |
HS 6-5-2 |
327 |
5083-O/6063- T6 |
6 |
|
|
18 |
5 |
|
269 |
5052/Al-Mg2Si |
8 |
concave |
conical |
18 |
6-4/5.7 |
H13 steel |
270 |
2024-T351/6061-T6 |
6 |
conical with 4° cavity |
square frustum |
18 |
7-3.5/5.9 |
H13 steel |
271 |
6061-T6/ 6351-T6 |
6.35 |
cylindrical scrolled |
cylindrical |
14 |
4/5 |
molybdenum M42/ HRC 63 |
273 |
6061-T651/5A06-H112 |
5 |
cylindrical |
conical |
16 |
5-4.2/4.6 |
|
275 |
2024-T3/ 6063-T6 |
8 |
|
conical triangular |
|
|
D5 steel/60 |
276 |
2219-T87/2195-T8 |
7.2 |
spiral |
|
16 |
|
H13 steel |
277 |
2017A-T451/cast AlSi9Mg |
6 |
scrolled |
tapered threaded |
24 |
6-4.5/5.7 |
HS 6-5-2 |
278 |
5083-H12/ 6061-T6 |
1.5 |
|
|
10-14 |
2-4 |
|
279 |
6061-T6/ 7075-T6 |
5 |
|
conic threated |
15 |
5/4.7/5° |
2344 steel/52 |
280 |
5083-H111/6082-T6 |
5 |
|
triangular, pentagonal |
20 |
5-6 |
DIN EN 1.7131 steel |
144 |
5083-H111/6351-T6 |
6 |
|
straight square |
18 |
6/5.6 |
|
282 |
2024-T6/ 7075-T6 |
5 |
flat |
smooth cylindrical |
15-16 |
3-8/4.7 |
high carbon steel |
283 284 |
Al-Mg-Si/Al-Zn-Mg |
15 |
|
|
35 |
20-12/14.5 |
|
285 |
2024-T3/ 6061-T6 |
4.8 |
concaved |
tapered threaded |
20 |
4 |
|
286 |
UFG 1050/6061-T6 |
2 |
concave |
thread |
12 |
4/1.8 |
steel |
329 |
2024-T6/ 6061-T6 |
4 |
concentric circles |
conical thread, deep groove thread, conical cam thread |
18 |
7-5/3.7 |
|
288289 |
6061-T6/7050-T7451 |
5 |
|
cylindrical threaded |
18 |
10 |
|
47 |
6111-T4/ 5023-T4 Lap joint |
1 |
|
threaded |
8 |
3/1.45 |
|
290 |
5086-O/ 6061-T6 |
6 |
|
straight cylindrical, threaded cylindrical, tapered cylindrical |
18 |
6-5/5.7 |
steel HSS |
291 |
2050-T4/ 6061-T651 |
20 |
single scroll |
conical threaded |
25.4 |
15.9/12.7/8° |
steel H13 |
292 |
5083-O/ 6082-T6 |
NR(~7) |
scroll |
triflute |
25 |
8/6.4 |
|
293 |
A319/A413 cast |
10 |
|
conical threated |
26 |
10-6/9 |
steel H13 |
294 |
7075-O/ 6061-O 7075-T6/ 6061-T6 |
3.17 |
concave |
cylindrical threaded |
15 |
4/3 |
steel H13/52 |
295 |
6061-T6/ 7075-T6 |
4.6 |
concave |
conical threaded |
15 |
7-5.2/4.7 |
steel H13 |
296 |
2024-T3/ 7075-T6 Lap joint |
5 |
concave |
cylindrical threaded |
15 |
6-4/6 |
steel H13/52 |
300 |
2014-T6/ 6061-T6 |
4.7 |
scrolled |
cylindrical threaded |
15 |
5/4.4 |
|
299 |
6351-T6/ 5083-H111 |
6 |
with concentric circular slots |
straight square, tapered square, straight hexagon, straight octagon, tapered octagon without draft |
18 |
6/5.6 |
high carbon high chromium steel |
302 |
A356/6061-T6 |
3 |
concave |
cylindrical |
15 |
5/2.6 |
high speed steel |
330 |
2198-T351/7075-T6 Lap joint |
3 and 1.9 |
flat |
conical |
15.5 |
max 4/3.1 |
|
304 |
2219-T87/5083-H321 |
6 |
|
straight cylinder, tapered cylinder, cylindrical threaded tapered threaded |
9, 12, 15, 18 |
6/5.7 |
Steel H13/50-55 VHN |
305 |
6061 cast/6061 rolled |
6 |
with concentric circular slots |
hexagonal |
19.2 |
6/5.8 |
HCHCr steel/ 62 |
306 |
6351-T6/5083-H111 |
6 |
flat |
straight square, straight hexagon, straight octagon, tapered square, tapered octagon |
18 |
6/5.7 |
High carbon high chromium steel/63 HRC |
127 |
5052-H34/ 5023-T4 |
~1.5 |
|
cylindrical threaded |
12 |
3.8/1.45 |
|
303 |
A356/6061-T6 |
3 |
|
cylindrical |
15 |
5/2.6 |
HSS steel |
307 |
5052-H34/ 5023-T4 |
1.5 & 1.6 |
|
cylindrical threaded |
8 |
3/1.45 |
|
308 |
7050-T7451/2024-T351 |
25.4 |
|
threaded |
|
|
|
332 |
5182-O/ 5754-O 5182-O/ 6022-T4 5754-O/ 6022-T4 |
~2 |
concave |
cylindrical threaded |
10.2 |
3.18/1.95 |
H13 steel |
333 |
6061-T6/ 2024-T3 |
12.7 |
|
threaded |
|
|
|
260 |
2024-T351/6056-T4 |
4 |
concave |
cylindrical threaded |
15 |
5 |
|
261 |
cast A356/ wrought 6061 |
4 |
|
screw-like |
|
|
|
57 |
2017-T351 |
5 |
|
|
15 |
6/4.7 |
|
309 |
1050-H24 |
5 |
|
|
15 |
6/4.7 |
|
26 |
2017A-T451/AlSi9Mg |
6 |
|
cylindrical threaded |
22 |
8 |
|
311 |
2017A |
6 |
|
|
25 |
8 |
|
357 |
6061-T6 |
1 |
flat |
quadrangular prism, quadrangular frustum pyramid, frustum |
7 |
2-1.5/0.9 |
|
335 |
6063/5083 |
6 |
|
Straight cylindrical |
20 |
5/5 |
steel HSS |
356 |
6181-T4 |
1, 2 |
concave, scroll |
cylindrical and threaded |
13 |
5, 6.5, 7 |
|
336 |
2618-T87/5086-H321 |
6 |
|
Straight cylinder, tapered cylinder, cylindrical threaded tapered threaded |
24, 30, 33, 36 |
12/5.7 |
steel H13 |
340 |
2024-T4 |
4 |
|
triflute with round bottom pin, triflat with round bottom pin, triflute with flat bottom pin, triflat with flat bottom pin |
|
|
high speed steel SW7M |
341 |
2024-T4 |
3 |
cylindrical, concave |
tapered unthreaded, tapered threaded |
20 |
6/3 |
|
344 |
2014-T6 |
5 |
|
straight cylindrical, tapered cylindrical |
18 |
6/4.8 |
stainless steel |
365 |
7075-T6 |
1, 0.8 |
concave |
cylindrical threaded |
10 |
4/1.2 |
Schilling 10S4ZGO/54-56 |
347 |
3003 |
2 |
flat |
conical |
19.5 |
6.8-5/1.7 |
X210Cr12 steel |
349 |
3003-H17 |
5 |
|
conical threated |
16 |
6/4.7/2.5° |
|
350 |
3003 |
5 |
concave |
square |
18 |
6/4.75 |
steel H13/45 |
351 |
3004 |
5 |
flat |
cylindrical threated tapered threated |
21 |
10/4.5/10° |
|
354 |
1100 |
5 |
|
|
7.86, 12, 15, 18, 22.13 |
2.6, 4, 5, 6, 7.37 |
high carbon steel 33, 40, 45, 50, 56 HRC |
316 |
6061-T6 |
9.5 |
flat |
composite (cylindrical shaft and cylindrical |
25.4 |
11/9(3) |
H13 steel |
317 |
6061-T67075-T6 |
6 |
flat |
cylindrical |
21 |
6/6 |
|
Summarizing, it can be notice that the pin length should be a little less than the plate thickness for butt joints arrangement to prevent the tool damaging or the backing plate of FSW machine used for FSW process. Additionally, the tool pin profile and diameter, shoulder shape and diameter, TRS and WS the most highly affect FSW process, particularly in case of dissimilar aluminium alloys. It is in agreement with observations from [
320].
4.8.1.4. Microstructure Evolution
The typical microstructure of a FSW joint comprised three zones including HAZ, TMAZ and SZ [
367,
368] The shapes of such zones were affected by the thermal and mechanical deformation induced by the tool during the welding process. The SZ exhibited fine-grain microstructures due to extensive grain refinement, while the TMAZ exhibited an elongated grain structure [
369,
370] The welding parameters affected the microstructure evolution, due to the higher influence of the material movement or flow in joints between dissimilar Al alloys compared to joints between the same Al alloys. The appropriate selection of all process parameters provided an intensive material mixing on both sides AS and RS of the joint and thus a sound weld. The electron backscatter diffraction EBSD - based orientation maps for the 5083-2024 joint [
371] revealed tilted and elongated grains in the TMAZ and refined grains d in the SZ resulting from dynamic recrystallization. Grain boundary orientations also varied in all three zones. The SZ comprised a higher amount of large (>10°) angular grain boundaries, while more of low (2–10°) angular grain boundaries appeared in HAZ. The SZ also exhibited a more intense texture compared to other zones.
Studying the FSWed joints of 7075-T651 and 2024-T351 alloys Hasan et al. [
130] found the difference in materials flow and mixing relating to the tool pin design. The grain size and shape of onion rings appearing in nugget zone were affected by material placement and tool pin geometry. The mixing stir zone became more homogeneous when the flute radius reached that of the tool pin. Three different sub-layers appeared in the weld nugget, two of them were close to each base welding material, and the other was a mix of both materials. The non-recrystallized heat-affected zone HAZ and thermo-mechanically affected zone TMAZ were similar in chemical composition to their corresponding base materials.
Ge et al. [
61] reported that four typical zones including base material BM, heat-affected zone HAZ, thermo-mechanically affected zone TMAZ, and stir zone SZ appeared in the lap weld cross-section. Stir zone SZs presented a bowl-like shape due to the tool geometry effect and comprised nonuniform grain size along the joint thickness due to the differences in material flow and temperature during the welding process. The dynamic recrystallization also occurred therein caused by the strong stirring of the tool and the elevated temperature effect. The material concentrated zone MCZ was also formed under the material plastic flow towards the tool zone during welding process. Simultaneously, the lath-like microstructure of the BM transformed into fine equiaxed grains. TMAZ comprised severely deformed and elongated grains induced by strong plastic deformation of SZ during FSW. The grains of 2024 alloys appeared in the upper sheet near the lap interface due to the laminar material flow. Also, the microcracks can occur on the tip of the cold lap on RS and gradually fade away in SZ.
Studying FSWed joints of components made of alloys 7075-T651 and 5083-H111 Kalemba-Rec et al. [
265] found that the weld centres comprised fine, equiaxed grains resulting from dynamic recrystallization. For a particular TRS, no differences in grain refinement appeared. The insufficient heat input had a neglected effect on grain size changes. However, the microstructure consisted of regions in the form of bands differing in grain size and in chemical composition. The bands came from both base alloys: one formed by the 7075 alloy and the other by the 5083 alloy. The dominant region of the stir zone comprised elements from the alloy from the advancing side. Studying the post-weld heat treated dissimilar FSWed 7075 and 2024 joints Safarbali et al. [
266] reported that fracture existed at the interface between thermo-mechanical affected zone TMAZ and heat affected zone HAZ on the retreating side 7075 of as-welded joint, while by applying post-weld heat treatment fracture was shifted towards the stir zone SZ of the welded joint. In post-weld heat treated joints, fracture surface was inter-granular, while in as-weld joint, fracture surface was mostly trans-granular. This was due to dissolution and coarsening of precipitates within grains in post-weld heat treated joints.
Studying FSWed sheets made of dissimilar alloys 2017A-T451 and 7075-T651 Hamilton et al. [
326] reported that because of the flow of surface material into the welded sheet thickness, the weld nugget comprised alternating layers of 7075 and 2017A. Such layers exhibited distinctive precipitate distributions because of their unique temperature histories affected by the material’s initial position. Supersaturated surface material flew into the process zone and formed a core comprising GP zones re-precipitating upon cooling. Mid-plane and bottom-plane material flew toward the sheet surface and embraced the surface material core. Within such region, the weld temperatures exceeded the equilibrium θ phase in 2017A, lowering the hardness, and simultaneously dissolved the equilibrium η/T phase in the 7075, causing re-precipitation of GP zones upon cooling and a hardness recovery.
During studies on FSWed joints of components made of dissimilar alloys 2017A-T451 and 7075-T651 Hamilton et al. [
170] reported that the microstructures comprised many dislocations due to remnants from the extrusion process and post-solution treatment stretching. The 7075 base alloy contained a much higher number of second phase particles due to the T6 temper than the 2017A alloy, being in the T4 temper. The dislocation density and the average grain size in both base alloys extruded and stretched under similar conditions were similar. The microstructure of weld nuggets was composed of interleaving bands of material from each alloy. The material from one side predominated on the other side of the nugget. On the RS of the nugget of the AS 2017A – RS 7075 weld the number of second-phase particles in 7075 was much higher than that in 2017A, while the density of residual dislocations between the two alloys was comparable. In the TMAZ and HAZ changes in the type and/or concentration of secondary phase particles can appear, particularly distinct in the 2017A alloy placed on the advancing side during welding.
Studying FSWed joints of components made of Al-Mg
2Si alloy and 5052 Al alloy Huang et al. [
269] identified three distinct zones in FSW joint: the base material zone BMZ, the transitional zone, and the weld nugget WN. The primary Mg
2Si phases comprised coarse equiaxed crystals for Al-Mg
2Si alloys in the BMZ. The WN was a mixture of rich Al-Mg
2Si and rich 5052 alloy forming a banded structure. In the WN, the equiaxed crystals varied into polygonal particles with lowered sizes in the rich Al-Mg
2Si zone. In addition to the white-rich Mg phase appearing in the rich 5052 zone near the interface, the 5052 alloy was unchanged.
Studying FSWed joints of components made of dissimilar 2024 and 6061 alloys Moradi et al. [
270] found grain refinement in the stirred zone via continuous and discontinuous recrystallization. The fraction of precipitates in the stirred zone of the retreating side exceeded that of the advancing side. The extent of continuous dynamic recrystallization in the TMAZ of the advancing side was less than that of the retreating side and the recrystallized grains seldom occurred on the advancing side. The initial texture components became asymmetric after FSW process. The overall texture intensity was weaker on the advancing side and stronger on the retreating side than that in the starting materials. The discontinuous static recrystallization and/or meta-dynamic recrystallization occurred on the advancing side.
Studying double-sided FSWed joints of components made of 6082-T6 and 7075-T6 dissimilar alloys Azeez and Akinlabi [
272] reported that the microstructure deviated from the conventional trend. The weld nugget exhibited no onion ring with a long flow arm. The worm hole defect occurred at the heat affected region of the 6082-T6 alloy. Little abnormalities at the retreating side were caused by the pre-heating of the plates during the initial welding process.
Studying single-sided FSWed joints of components made of dissimilar 6082-T6 and 7075-T6 alloys Azeez et al. [
328] reported that the equiaxed grain structures resulted from the dynamic recrystallization mechanism at the weld nugget. Some microstructure imperfection occurred at the weld nugget when 6082 Al plates were clamped on the retreating side to the backing plate. However, deviation in the positioning of the Al plates prevented the fabrication of good bonding and quality welds despite the material flow and mixing occurrence.
Studying FSWed joints of dissimilar alloy 6061-T651 and 5A06-H112 Peng et al. [
273] noticed that the grain structure evolution in the stir zone was dominated by continuous dynamic recrystallization. The grain size in the HAZ and TMAZ was refined. Fractures in all tensile specimens were of ductile nature due to the presence of dimples. The enhancement of heat input enlarged the size of HAZ and lowered the slant angle of HAZ leading the fracture angle to decrease and changing the dimples from inclined ones to normal ones. Shear stress formed shallow and inclined dimples, whilst equiaxed and normal dimples resulted from normal stress.
The FSWed joints of dissimilar alloys 2024-T3 and 6063-T6 studied by Sarsilmaz [
275] comprised several weld zones including BM, SZ, TMAZ, and HAZ formed in close relation to plastic flow and frictional heat generation during the welding process. Such joints were sound without any micro cracks, micro voids, and unbounded regions in the welded interface. There were different morphologies of the micro-structure at interface zone of joint. All welds exhibited the formation of the elliptical onion structure in the weld center. The SZ also included onion rings where the tool pin contacted the welded parts. Onion ring patterns exhibited lamellar-like structures of stacked two materials. Under low values of traverse speed and high TRS, when higher frictional heat was generated, generated weld nugget was wider than under other parameters. The higher temperature and severe forging deformation resulted in grains smaller than those of the base metal. The SZ had a fine equiaxed grain structure. Under low values of TRS and WS, the clear vase-like boundary line appeared between TMAZ and SZ due to high deformation and frictional heat between the weldment and the tool pin.
Studying FSWed joints of components made of 2219-T87 and 2195-T8 dissimilar alloys No et al. [
276] found that the microstructure of the weld joint underwent dynamic recrystallization because of high deformation and frictional heat. During studies on FSWed joints of components made of dissimilar Al alloys: wrought 2017A and cast AlSi9Mg Kopyscianski [
277] reported that the weld microstructure comprised alternating bands of the welded alloys. The AlSi9Mg alloy on the advancing side dominated the weld center. The grain size within the bands was close in both alloys. The nugget side comprised a high density of the bands of the 2017A alloy.
Studying Fessed joints of the 6061-7075 alloy Bijanrostami et al. [
279] found that under high heat input conditions including high rotation and low WSs, large grains and smaller dislocation densities appeared in the SZ. Contrarily, under low heat input conditions, various defects developed. Studying the effect of the TRS to the WS ratio (υ ratio) on the strength of the FSWed joints of dissimilar alloys 6082-T6 and 5083-H111 Kasman et al. [
280] found that the small cavity- and tunnel-type defects occurred at the nugget zone, profile of which contained various onion rings. According to Palanivel et al. [
281] the FSW joints can comprise defects such as pinhole, tunnel defect, piping defects, kissing bonds, cracks, etc. caused by the improper flow of metal and insufficient consolidation of the metal in the weld zone. Studying FSWed joints of components made of 5052- H32 to 6061-T6 blanks Doley and Kore [
282] found that the dynamic recrystallization and finer grain size with uniform mixing appeared at the center of SZ. Intermetallic compounds were also formed during welding. Studying FSWed joints of components made of the dissimilar 2024-T6 and 7075-T6 alloys Saravanan et al. [
283] found that the joint fabricated under D/d ratio equal to 3 provided the fine recrystallized structure of SZ, and the grain size smaller than the base material grain size due to smaller shoulder diameter. Studying FSWed joints of sheets made of dissimilar Al-Mg-Si/Al-Zn-Mg alloys Yan et al [
284] found that different joint cross-sections were obtained for different sheet configurations. Coarser β' phases occurred at the heat affected zone HAZ of the AlMgSi alloy side.
Yan et al.[
284] reported that for the Al–Zn–Mg AS joints, the precipitated phases were the AlFeMnSi or AlMnCrSi phases and the β′ phase owned sizes of about 0.5–0.8 μm. For the Al–Mg–Si AS joints, similar precipitated phases occurred with smaller β′ phase owning sizes of about 0.3–0.5 μm. The quantity of the β′ phase on the Al–Mg–Si AS joint was more than that on the Al–Zn–Mg AS joints. The secondary phase particle at the joint fracture region was characterized by submicron β′ phase. For the Al–Zn–Mg AS joint the β′ phases were more dispersed, which was beneficial to the bridging effect. The FSWed joints of dissimilar alloys 2024-T3 and 6061-T6 studied by Zapata et al. [
285] exhibited various regions including the TMAZ and HAZ with similar shapes, locations, and sizes in all the samples. All the sample cross-sections presented a ring flux pattern in the nugget region indicating the vertical movement of the material.
Studying butt FSWed joints of 2 mm thick plates including the one rolled from ultrafine-grained UFGed 1050 Al alloy and the one made of the 6061-T6 alloy Sun et al. [
286] found that in the stir zone, the initial nano-sized lamellar structure of the UFGed 1050 Al alloy plate took the form of an equiaxial-grain one with a larger average grain size caused by the dynamic recrystallization and subsequent grain growth. Simultaneously, an equiaxial-grain structure with a significantly smaller grain size occurred in the 6061 alloy plates, together with coarsening of the precipitates.
Studying FSWed joints of 2024-T6 alloy Sun et al. [
329] found that the metal in the weld nugget zone WNZ came from the base metal of the advancing side, when the thread was the driving force of the downward movement of the FSW plastic metal. All joints formed a particularly good union, with an onion ring pattern appearing in cross-section. The minimum grain size of the WNZ obtained with the conical cam thread stirring head was 7~12 μm. FSW usually allows elimination of porosity, small distortion, and so on [
372,
373]. During studies on butt FSWed joints of components made of dissimilar 6061 to 7050 alloys Rodriguez et al. [
289] found in joints microstructure occurred distinct lamellar bands and various degrees of intermixing affected by TRS .
Studying lap FSWed joints of plates made of dissimilar alloys 6111-T4 and 5023-T4 Yoon et al. [
47] found that the threaded probe well correlated to the onion ring structure formed as soon as it touched the probe. The remnant of original interface between the top and bottom plates after the welding process and asymmetrical flow around rotating tool well correlated to the formation of void defects under low heat input conditions.
During studies on FSWed components of the heat treatable 6061 and non-heat treatable 5086 alloys Ilagovan et al. [
290] found that the use of threaded tool pin profile provided better flow of materials between two alloys and the generation of defect-free stir zone. Such tool provided formation of finer and uniformly distributed precipitates, circular onion rings and smaller grain compared to the tapered pin profiled tool. During studies on butt FSWed joints of components made of dissimilar alloys 2050 and 6061 Reza-E-Rabby et al. [
291] found that the stir zone comprised bands of mixed and unmixed material with the degree of material intermixing increasing with the enhancement of TRS. Under monotonic tensile loading, welds failed via the heat-affected zone on the 6061-alloy side of the weld. For the low TRS, failure appeared in the stir zone due to poor material intermixing.
Studying the anodizing behavior of FSWed joints of dissimilar 5083-O and 6082-T6 alloys in 4M H
2SO
4 solution Donatus et al. [
292] found that the 5083-O rich zones were more oxidized during anodizing compared with the 6082-T6 rich zones. The nugget and the thermo-mechanically affected regions of the individual basic alloys exhibited a decrease in porous anodic oxide thicknesses. Sputtering deposition of pure Al on the weld, prior to anodizing, lowered the variations in the oxide thicknesses across the weld. Such a method prevented the boundary dissolution related to the activity of the Mg
2Si phase often occurred after anodizing the dissimilar weld of the alloys. Studying FSWed joints of plates made of dissimilar cast Al–Si alloys A319 and A413 Karam et al. [
293] found that the joints comprised the Si particles and of α-Al grains. At the center of stirred zone, the Si particles could be more uniformly distributed than in the other zones.
Studying the butt FSWed joints of dissimilar 7075-O/6061-O and 7075-T6/6061-T6 alloys studied by Ipekoglu and Cam [
294] found that the as-welded 1500/400 O joint exhibited no welding defect except its root region, while some weld defects appeared in such a joint after post-weld heat treatment PWHT. The joint areas of the dissimilar T6 joints obtained in both the as-welded and post-weld heat-treated conditions exhibited no weld defects. The dynamically recrystallized zones DXZs of all the as-welded joints both in the O and T6 temper conditions comprised a layered (banded) structure formed due to the mixing of two BMs, known as the intercalation of BMs. Microstructure of the DXZs was dominated by the alloy located on the AS. The typical orientation of the grains in the TMAZ was strongly affected by the material flow resulting from the action of the stirring tool. PWHT had only a slight effect on the shape of the grains in the TMAZ. PWHT also caused the formation of abnormal grain growth AGG, both in O and T6 joints. The AGG formation occurred all over the cross-section, in the O joints, whereas in the shoulder regions in the T6 joints.
Studying the lap FSWed joints of dissimilar 2024-T3 and 7075-T6 alloys Song et al. [
296] found that welded joints comprised four typical zones including base metal BM, HAZ, TMAZ and SZ. The grain structure in the HAZ of the upper 2024 sheet was like that of BM. In TMAZ microstructure comprised severely deformed and elongated grains resulting from drastic plastic deformation of SZ during FSW. In SZ, the microstructure included dynamically recrystallized fine equiaxed grains resulting from the drastic deformation induced by sufficient stirring during welding. Grains in the upper SZ were coarser than those in the bottom SZ due to the former had more time to grow up provided by the higher temperature compared to those in bottom SZ. During studies on FSWed joints of components made of dissimilar 5083-H111 and 6351-T6 alloys Palanivel et al. [
298,
306] reported that the weld zone comprised three types of microstructures, namely unmixed region, mechanically mixed region, and mixed flow region. Studying FSWed joints of components made of dissimilar A356 and 6061 alloys Ghosh et al. [
302] reported that microstructure of WN had uniform dispersion of Si-rich particles, fine grain size of 6061 alloy, and disappearance of the second phase within 6061 alloy.
Studying the FSWed lap joints of 7075-T6 and 2198-T351 studied by Velotti et al. [
330] noticed that the hook defect comprising an S-shaped separation line between the two materials joined, typical for such joining technique in this specific configuration was not fully avoidable, as the stirring action caused by tool motion cannot completely mix two materials initially stacked. Such a defect caused a preferential path for the crack growth and propagation and for the localized corrosion phenomenon, affecting the joint behavior. The kissing bonds resulted from inadequate material mixing and stirring occurred in the core of the nugget zone and in the radii between skin and stringer. Both the alloys exhibited microstructure with round shaped equiaxed grains. The average grain size for the 2198 was about the 30% of the one of the 7075.
Investigating FSWed joints of plates made of dissimilar Al–Cu alloy 2219-T87 and Al–Mg alloy 5083-H321 Koilray et al. [
304] found the material placed on the advancing side dominated the nugget region. Welds comprised base material zone BM, stir zone or weld nugget SZ, thermo-mechanically affected zone TMAZ, and heat-affected zone HAZ. BM contained many undissolved second-phase intermetallic particles. The second-phase particles in alloy 2219 comprised Al
2Cu (θ) eutectic particles, while alloy 5083 included iron/manganese aluminides. Compared to alloy 2219, alloy 5083 comprised fewer and finer second-phase particles. The TMAZ on the advancing side exhibited highly deformed grains, with discernible SZ/TMAZ and TMAZ/HAZ boundaries. However, on the retreating side, these interfaces were diffused, especially the latter. In the HAZ, on either side of the weld nugget, the grain structure exhibited no noticeable changes compared to the respective base materials.
Studying the FSWed dissimilar cast and wrought 6061 alloy Dinaharan et al. [
305] found that the microstructure of the dissimilar joints comprised four zones including base metal, HAZ, TMAZ and weld zone. The weld zone covered an unmixed region and a mechanically mixed one. The unmixed region consisted of the microstructure of cast and wrought 6061 alloy. The mechanically mixed region occurred near the zigzag line containing the microstructure of both Al alloys. Some degree of penetration of one alloy into the other occurred. The plasticized dissimilar alloys were mechanically coupled in the mechanically mixed region. Both the materials after dynamic recrystallization during FSW exhibited finer grain structure than before FSW.
Studying FSWed joints of components made of dissimilar alloys 5052-H34 and 5023-T4 Song et al. [
127] noticed that in the same manner as constitutional liquation, at high heating rate, the main liquation-inducing precipitates were not dissolved in the matrix and reacted with Al to form the partially melted zone PMZ, after which liquation cracking occurred where strain was applied to the PMZ. Solid solution treated 5023 alloy comprised in the matrix many precipitates including Mg
2Si, Al
6CuMg
4, and Al
6(CuFe). Al
6CuMg
4 formed a stable phase at room temperature and reacted with the Al matrix at around 470°C. The main liquation-inducing precipitate was Al
6CuMg
4 forming the PMZ (constitutional liquation) at around 480 °C during the FSW process.
Studying FSWed joints of components made of dissimilar A356 and 6061 alloys, Ghosh et al. [
303] reported that the structure of joints exhibited recovery-recrystallization in the stirring zone and breaking of a coarse eutectic network of Al–Si. Dispersion of fine Si rich particles, refinement of 6061 grain size, low residual stress level and high defect density within weld nugget allowed increasing bond strength. Lowering the tool rotational and traversing speed enhanced the domination of such phenomena.
Studying the FSWed joints of dissimilar 5052/A5J32 alloys, Kim et al. [
307] reported that the weld nugget was formed according to the arrangement of the materials. The softened material moving from the AS toward the RS caused the formation of an empty region with the shape of the tool pin. The material on the RS filled most of the upper half of the empty region. When the A5J32 alloy was fixed on the RS, a high amount of the base material on the retreating side A5J32 was stirred toward the advancing side 5052, as the rigid material A5J32 was easily pushed out from the soft material 5052 and the two base materials were stirred in a zigzag shape. When the 5052 alloy was fixed on the RS, the flow of 5052 (retreating side) was limited by the more rigid material on the advancing side A5J32. The softened A5J32 accumulated unnaturally. Under conditions with a lower heat input, such as at 1000 rpm and 400 mm/min, some welding defects occurred. Under opposite conditions, defect-free welds were obtained.
Studying the FSWed butt joints of 7050-T7451/ 2024-T351 alloys, Prime et al. [
308] noticed that the stirred zone, i.e., the weld nugget or dynamically recrystallized zone comprised fine equiaxed grains. The nugget exhibited the onion ring structure. On both sides of the stir zone, there are TMAZs comprising highly deformed grains from the stirring action. The TMAZ was more uniform on the advancing side and more diffuse on the retreating side. The heat-affected zones extended out of the TMAZs on both sides.
Miles et al. [
332] reported the occurrence of failures in the heat-affected zone HAZ of the 6022 or in the weld nugget itself for the welded joints of components made of the dissimilar 5182/6022 and 5754/6022 alloys. The 5182/5754 alloy pair studied was softening-free, as such alloys were in the annealed condition, while the softening occurred in the 6022 side of the 5182/6022 and 5754/6022 alloy pairs.
Ouyang and Kovacevic [
333] studied material flow and microstructural evolution for welded joints of components the one made of 6061alloy as itself and the second of dissimilar alloys 6061-T6 and 2024-T3 both of 12.7 mm in thickness obtained via FSW under different welding conditions. They found that plastic deformation, flow, and mechanical mixing of the material were characterized by asymmetry characteristics at both sides of the same and dissimilar welds. The microstructure in dissimilar 6061 /2024 alloys welds highly differed from that in case of a 6061 alloy to itself. Vortex-like structures with the concentric flow lines characteristic for a weld of 6061 alloy to itself, and alternative lamellae with various alloy constituents for a weld of 6061 and 2024 alloy, resulted from the stirring by the threaded tool, in situ extrusion, and traverse motion along the welding direction. The nugget zone of dissimilar 6061-Al/2024-Al welds comprised the mechanically mixed region with the dispersed particles of different alloy constituents, the stirring-induced plastic flow region with alternative vortex-like lamellae of the two Al-alloys, and the unmixed region with fine equiaxed grains of the 6061 alloy. Within these regions, the material withstood an extremely high degree of plastic deformation due to the occurrence of dynamic recovery or recrystallization of the microstructure. The degree of material mixing, the thickness of the deformed Al-alloy lamellae, and the material flow patterns were affected by the related positions in the nugget zone and the processing parameters.
Studying the butt FSWed joints of dissimilar 2024-T351/ 6056-T4 alloys, Amancio-Filho et al. [
260] reported that the welds comprised four various regions including base material BM, heat affected zone HAZ, thermo-mechanically affected zone TMAZ and stir zone. The BM of 2024-T351 alloy microstructure exhibited elongated grains in the rolling direction. This BM contained copper-rich particles being second phase θ-CuAl
2. The BM of 6056-T4 alloy also revealed the microstructure with grains oriented in the rolling direction. This BM comprised two kinds of particles: the ones rich in Mg and Si, identified as the intermetallic β-Mg
2Si and the others rich in Mn and Fe. The stirred zone of joint exhibited a lamellar material flow pattern due to material mechanical mixing. The SZ revealed a dynamically recrystallized microstructure with refined grains. The thermo-mechanically heat-affected zone of the alloy 6056-T4 possessed annealed structure. Change in grain orientation started in the transition between TMAZ/stir zones. The grains were rotated by tool action and by reaching approximately 90° tilting, some degree of recrystallisation occurred, represented by a lowered grain size. Lee et al. [
261] found that the microstructures of dissimilar formed A356/6061 joint exhibited the mixed structures of two materials. The stir zone exhibited the onion ring pattern look-like a lamellar structure. The microstructure of the SZ comprised the material fixed at the retreating side.
Studying the FSWed joints of components made of 7003/7046 dissimilar alloys Yang et al. [
312] observed an obvious "S"-shaped dividing line in the weld nugget area of the 7003/7046 dissimilar alloy FSW joint. Both sides of the S line comprised fine equiaxed grains, with a size of about 5 μm; the grain size in the heat-affected zone and the heat-affected zone was higher than that in the nugget zone, and the structure in the heat-affected zone comprised recrystallized grains and recovered grains. The coarsening of subgrains lowered the grain size of the heat-affected zone on the 7003 side, and the grain size of the heat-affected zone on the 7046 side was coarsened.
Investigating FSWed joints of dissimilar 2219-5083 alloys Mastenaiah et al. [
321] found that the nugget zone revealed the mixing pattern highly affected by the tool offset, the tool rotation speed and the tool traverse speed. Intimate mixing of dissimilar alloys occurred at higher tool rotation speeds and lower tool traverse speeds.
Studying the FSWed joints of dissimilar 2618-T87/ 5086-H321 alloys, Sasikala et al. [
336] reported that the nugget region was dominated by material on the forward-moving side. The weld contained four microstructural zones: BM, SZ, TMAZ, and HAZ. Both BMs contained particles of second-phase intermetallic. Iron/manganese aluminides appeared in alloy 5086's second-phase particles, while eutectic Al
2Cu particles were in alloy 2618. Particles in the alloy's second phase 5086 were smaller and finer than those in alloy 2618. The grain structure of the weld nugget was like that in the HAZ. The SZ/TMAZ and TMAZ/HAZ boundaries were clearly distinguishable on the forward-moving side of the TMAZ. Dispersion of them was higher on the RS.
Studying the single-sided butt FSWed joints of 3003-O alloys Aydin et al. [
339] reported that the welds comprised four zones: BM, HAZ, TMAZ and SZ. The BM microstructure revealed the elongated grains resulting from the rolling operations. The stirred zone revealed a fine-grained equiaxial dynamic recrystallized microstructure. The grain size in the SZ was smaller than that of BM. An enhancement in rotation speed and lowering WS lowered the grain size in the SZ due to the higher heat input for dynamically recrystallized microstructure. The TMAZ microstructure exhibited a highly deformed structure near the SZ zone due to less heat and deformation appearing in TMAZ compared to those in the SZ. After the TMAZ appeared the HAZ exposed only to a thermal cycle, but the plastic deformation therein was insufficient to modify the initial grain structure. The transition zones from the stirred zone to TMAZ of the joints also occurred. On the AS a distinct boundary between the SZ and the TMAZ appeared while the boundary between the SZ and the TMAZ on the RS was unclear . The region with Al
2O
3 particles, stretching from the top to bottom across the whole section of all 3003-O weld zones at the RS contained the ‘kissing bond’ defects.
Aydin et al. [
337] found that BM microstructure comprised the elongated grains resulting from the rolling operations. In the weld centre the NZ occurred, which revealed dynamically recrystallized grains. On the AS, microstructure varied rapidly due to the higher speed of plastic material than on the RS, and a distinct boundary between the NZ and TMAZ occurred. On the RS, microstructures from the NZ to TMAZ varied more smoothly. In the TMAZ the grain structure was deformed but no recrystallization occurred. The microstructure in the HAZ affected by the heat but not by deformation, was like that of BM; the grains were slightly overgrown due to the exposure to welding heat.
Studying the FSWed of 4 mm thick plates made of alloy 2024 Weglowski et al. [
340] reported that the joints exhibited a characteristic shape of a nugget zone, heat-affected zone and thermo-mechanically affected zone.
Studying the FSWed joints of 2024-T4 alloy Nejad et al. [
341] reported that the best welds features including the finest grain in stir zone, the best visual quality and smoothness, were obtained with rotation speed of 500 rpm, traverse speed of 55 mm/min, plunge depth of 2.7 mm and by threaded tool, and with rotation speed of 1300 rpm, traverse speed of 115 mm/min, plunge depth of 2.9 mm and by unthreaded tool. Unthreaded tool provided more uniform structure from the point of view of smoothness. The WE widely varied in range from 35.6% to 95.7%.
Investigating FSWed joints of 8 mm thick plates made of 2014-T6 Al alloy Lin et al. [
343] found that the different regions of the joint exhibited different microstructure variation affected by different thermomechanical actions therein.
Liu et al. [
374] reported that defects including void, unbonded interface and incomplete refilling appeared when using 7075 alloy as the upper sheet. No defects occurred when using 6061 alloy as the upper sheet. With enhancement of the sleeve plunge depth, better material mixing appeared between the upper and lower sheets.
Studying FSWed joints of components made of 2017A alloy Mroczka et al. [
311] reported that the weld nugget exhibited an average grain size of 5 μm, moderate density of dislocations as well as the presence of nanometric precipitates located mostly in grains interiors. The weld nugget presented ductile fracture with brittle precipitates in the lower part.
Investigating FSWed joints of components made of a cast AlSi9Mg and 2017A alloy Mroczka [
310] found that welds comprised defects, despite greater plasticity of the material occurring due to the raising of temperature. The constituent stable phases within the cast alloy exhibited considerable fragmentation to various degrees. The material above the weld nugget was unmixed and comprised non-welding micro-defects. A metastable state of the 2017A alloy occurred within the weld nugget zone due to the natural ageing.
Studying butt FSWed joints of 6013 Al plates obtained via pin offset technique Kasman and Ozan [
313] found kissing bonds in welds, originating from the broken oxide layers, and formed particularly in the stir zone. The microstructure of joints comprised phases belonging to Mg
2Si, Al
4Cu
2Mg
8Si
7 and Al(MnFe)Si.
Kasman and Yenier [
322] reported that the microstructure of the two BMs comprised elongated grains in the rolling direction. The nugget zone included the fine equiaxed grains formed by the occurrence of dynamic recrystallization. The structure of grains in the TMAZ was deformed, elongated, and oriented to the rotation of the pin. The size and orientation of grains of the TMAZ differed from that on the NZ. Compared to the TMAZ structure, that of HAZ comprised grains overgrown and was like that of the BM.
Investigating butt FSWed joints of components made of dissimilar 7020-T651 and 5083-H111 alloys Torzewski et al. [
352] found various shapes of the stir zone and defects caused by excess and insufficient heat input.
Studying spot FSWed joints of sheets made of 5454 alloys Choi et al. [
353] found that the enhancement of tool rotation speed changed the macrostructure of the friction-stir-spot-welded zone, especially the geometry of the welding interface. However, the change in the dwell time at the plunge depth of the tool only slightly affected the microstructure of the welds.
Dong et al. [
355] reported that SZ comprised fine and equiaxed grains due to dynamic recrystallization. With the enhanced cooling rate, microstructure of the UFSW joint is finer than that of air-cooling FSW, and the area of HAZ and TMAZ in the UFSW joint becomes smaller. The precipitation evolution was strongly affected by the processing parameters of UFSW. In the 1000-120 sample, fine precipitates exist in HAZ, TMAZ and SZ
Zhou et al. [
357] reported that the pattern of the weld cross section was a “flat T” and no obvious “S curve” occurred in nugget zone NZ. Heat affected zone HAZ and thermo-mechanically affected zone TMAZ were also narrow. The nugget zone NZ comprised the grains finer than those in other zones. A clear band line appeared between the nugget zone NZ and thermo-mechanically affected zone TMAZ located between NZ and the heat affected zone HAZ. There was no clear dividing line between HAZ and the base-metal BM. No obvious “S curve” appeared in all joints.
Tra et al. [
375] reported that in case of FSW process the fatigue crack propagation FCP rates depended to the propagating location, the test temperature, and the PWHT conditions.
Studying FSWed joints of sheets made of 6013-T6 alloy Kafali and Ay [
389] reported that the microstructure of the welding zone comprised four subzones including a base material, heat affected zone HAZ, thermo-mechanical affected zone TMAZ and weld nugget. The parent material and the weld region contained homogenous distributions of the fine and coarse Mg
2Si particles. A dynamically recrystallized grain structure appearing in the weld nugget exhibited a smaller grain size compared to the BM. Such dynamically recrystallized grains were equiaxed contrary to the elongated grains in the rolled BM. Fine equiaxed grains in the FSW region occurred due to dynamic recrystallization due to plastic deformation during the welding process.
Investigating the butt FSWed joints of sheets made of the dissimilar 2014-T3/5059-H11 Al alloys Saleh [
377] found that a structure of fine grain occurred in the nugget zone due to recrystallization.
Studying the FSWed joints of components made of 2219-T87/2219-T62 alloys Venkateswarlu et al. [
378] found that the microstructure of the 2219-T62 welds exhibited coarse grains formation in the thermo-mechanically affected zone and heat-affected zone.
For FSWed joints of Al alloy 2014-T651 Kollapuri [
379] reported that HAZ comprised large grain size compared to the nugget zone.
Studying FSWed joints of components made of 3003 alloy with different initial microstructures Tan et al. [
380] reported that the size of recrystallized grains and the number of second-phase particles in the weld nugget zone WNZ lowered with a decrease in welding ambient temperature. At the same welding condition, both the size of recrystallized grains and the volume fraction of (Fe,Mn)Al
6 particles in the hot bands were below those in the annealed hot bands.
Studying the FSWed joints of plates made of 7204-T4 alloys Deng et al. [
381] reported that the average grain size AGS and recrystallization fraction of nugget zone NZ reached 4.7 μm and 81.9 % in as-welded AW treatment, 4.8 μm and 82.4% under the post-weld artificial aging AA treatment, 5.9 μm and 86.5% under the heat treatment of solid solution followed by artificial aging SAA, respectively. The grain structure of NZ was slightly influenced by AA treatment, and the AGS and recrystallization fraction of NZ enhanced by 25.5% and 5.6% under SAA treatment.
Zhao et al. [
314] studied the influence of exchanging advancing AS and retreating RS side material on microstructure, mechanical properties, and electrochemical corrosion resistance FSWed joints of components made of dissimilar 6013-T4 and 7003 alloys. The joint with the 6013-T4 placed at the advancing side AS was called the A6R7 joint. Accordingly, the A7R6 referred to the joint with the 7003-alloy placed at the AS. The authors reported that various joint cross-sections appeared when exchanging AS and RS materials. The material on the AS was more deformed during the welding process. When the 6013 alloy was positioned on the AS, the plastic flow of the weld was more sufficient.
Studying FSWed joints of components made of dissimilar 2024-T3 and 2198-T3 alloys Texier et al. [
287] found banded macrostructures with heterogeneous mechanical properties in the shoulder-affected region. They were accompanied by pronounced textures regions. The banded macrostructures appeared in the nugget region.
For FSWed joints of 6061-T6 alloy Juarez et al. [
316] noticed four characteristic zones of friction welding including agitation zone (stir zone), TMAZ, HAZ and unaffected material UFM. The fracture surfaces of tension specimens revealed the micro-voids present in the fracture zones for the three cases BMW, HTBW and HTAW. The fracture surface of the base material presented micro-holes of 6-8 μm in diameter. For the case of BMW and HTBW, the diameter of the micro-holes was greater compared to the base material at 8-10 and 10-12 μm, respectively. It was due to the reduction and separation of nucleation sites, allowing them to grow at a larger size. For HTAW case, the micro-holes had small and shallow sizes, due to the occurrence of numerous nucleation sites causing the merge of micro-holes, limiting growth at a larger size. Solubilized and partial ageing heat treatments of 6061-T6 alloy initiated the formation and distribution of the precipitates in the material. The predominant precipitates for the BMW and HTBW cases were Al-Mg and Al-Si, respectively; while for the HTAW case, it was Fe-Mg
2Si.
Unfried-Silgado et al. [
382] studied the influence of shoulder geometry of tool (flat and featured by concentric circles and by spirals) on microstructure and mechanical properties of FSWed joints of AA1100 alloy obtained using a milling machine revolutionary under pitch value (R) constant of 0.1 mm/rev. They reported that the featured shoulder tools strongly affected the thermal cycles, generating a plasticized wide region and the biggest grain size in the stir zone when compared with the flat shoulder tool. The featured shoulder tools induced thermal cycles in the regions out of the stir zone less severe than flat shoulders. The surface area of tested tools was 27% and 11% lower in flat shoulders than in featured shoulders, respectively.
Studying butt FSWed joints of dissimilar alloys 6061-T6/7075-T6 Godhani et al. [
317] noticed that the size of the grains varies in the different joint zones including nugget zone NZ, thermo-mechanically affected zone TMAZ, heat-affected zone HAZ, and base metal BM. Size of the grain deciding the strength was affected by amount of heat input, mixing of the materials, and the rate of cooling. The high grain density of 7075 induced the strength of 7075 was higher than of 6061. The grain density in the nugget zone is the highest, and hence the welded specimen has higher strength compared to the base metals. The failure can occur from the advancing side of joint as the density of grain was less compared to other locations.
Aval et al. [
358] studied the effect of tool on the mechanical properties and microstructural behaviour of FSWed 5 mm-thick plates made of 5086-O and 6061-T6 alloys. They found that the tool with a concave shoulder and a conical probe with three grooves provided higher heat input and temperatures promoting more homogeneous stir zones than the tools with flat shoulders and threadless or threaded cylindrical probes. The grain sizes of the SZ on the 6061 side were finer than those on the 5086 side, and lowering the weld pitch, i.e., a ratio of WS/TRS induced coarser grain structures in the SZ. The material in the weld nugget zone was a mixture of the two alloys, with closer to Mg content in 6061 region.
Studying the FSWed joints of components made of 5052 and 6061 alloys obtained with various pin-eccentric stir tools, Wang et al. [
383] found that the welding heat input caused both the coarsening of strengthening precipitates and dynamic recrystallization and softening of the nugget zone (NZ). The use of pin eccentricity promoted the material flow in the NZ and the higher area of the “onion ring”. The average grain size and fraction of recrystallized grain in the NZ lowered with the enhancement of the pin eccentricity.
Guo et al. [
384] studied FSWed joint of components made of dissimilar 6061/7075 alloys. They found that the material mixing was much better for 6061 alloy on the AS and multiple vortexes centers formed vertically in the nugget center. The onion ring comprised three distinct sub-layers: 6061 alloy sub-layer, 7075 alloy one, and mixed sub-layer of the two alloys. The thicknesses of these onion ring sub-layers were in range 30–100 μm. Both AA6061 and AA7075 alloys were dynamically recrystallized. The grain size highly lowered with the enhancement of WS. The grain size of 7075 alloy sub-layer was much lower than that of AA6061 sub-layer in the same weld. The fractured surfaces of tensile tested specimens exhibited many equiaxed dimples with various size. Shallower dimples occurred in the fractured surface of the joints obtained under lower heat input. Second phase particles comprising incoherent β-Mg2Si and various Al–Fe–Si intermetallics effectively provided nucleation sites for microvoids during fracture process.
Sato et al. [
385] investigated the mechanical and microstructural behavior of FS welded joints of 2024/7075 alloys. They found that the amount of heat generation highly influenced the material movement around the pin. The high heat input caused onion ring patterns in the SZ, while low heat input clearly divided the SZ into 2024 and 7075 regions.
Da Silva et al. [
386] investigated FSWed joints between components made of dissimilar 2024/7075 alloys. The maximum WE was of about 96% and the welded specimens were fractured in HAZ of RS.
The 3 mm thick plates made of dissimilar 7075/5083 alloys, studied by Dewangan et al. [
387], were FSWed using tool made of H13 steel, under TRS of 1400 rpm and WS of 20 mm/min and 45 mm/min, obtaining grain size in range 6-10 µm.
The 2.5 mm thick plates made of 2219/7475 alloys, studied by Khan et al [
388], were FSWed using tool made of High Carbon Steel and possessing cylindrical threaded pin, under rotating speed of 900 rpm and WS of 100 mm/min, obtaining grain size below 6 µm.
Abidi et al. [
389] studied T-FSW joint of 2 mm thick 7075/2024 alloys with 7075 placed as stringer and 2024 as skin obtained using tool made of High chromium high carbon steel with shoulder diameter in range of 12-16 mm, tapered cylindrical pin with length of 1.9 mm and diameter of 1.8 mm on the insertion side and of 5.6 mm on the shoulder side, under TRS in the range of 560-900 rpm, WS range of 40-63 mm/min and tilt angle of 2 degrees. They found that the TMAZ zone underwent induced plastic deformation caused by lesser heat input. This led to partial recrystallization forming coarse grains.
4.8.1.5. Mechanical Properties Hardness
The hardness of the FSW joint was strongly associated with the joint strength and its deformation behavior. The hardness distributions exhibited a high asymmetry along the cross-section of dissimilar material joints due to the various microstructural zones SZ, TMAZ, HAZ resulting from the thermo-mechanical history during welding. As the maximum temperature occurred at the SZ, precipitates or strengthening particles dissolved partially or completely lowering hardness in SZ. The lowest hardness values appeared in the HAZ due to the coarsening of precipitates or over-ageing. Thus, failures occurred most in the HAZ zone or site. The hardness values in SZ were higher than those in the BM (exhibiting sometimes low strength values) due to the combined influence of grain refinement and the effect of both BMs in the SZ. However, various initial conditions of heat-treatable alloy combinations could make such hardness distribution completely different [
256].
Studying FSWed joints of components made: the first of 2024-T351/5083-H112 alloys and the second of 7075-T651/2024-T351 alloys, Niu et al. [
262] characterized joints hardening by the ratio of HVf/HVw, where HVf and HVw were the microhardness of the fractured and the as-welded joints, respectively. This ratio was higher than one in the SZ, TMAZ and HAZ, which related to the strain-hardening behavior of the joints. The hardness distribution in the dissimilar material joints was strongly affected by strain hardening and the fracture origin.
For the FSWed joints of 7075-T651/2024-T351 alloys, Hasan et al. [
130] found that the distribution of weld hardness determines the tensile strength of welding joint. The weld hardness drops at the HAZ of softer material. Maximum reduction in weld hardness at the heat-affected zone was achieved using the pin tool with a flute of radius equal to that of the pin.
Ge et al. [
61] reported that 7075 BM exhibited higher hardness compared to 2024 BM value. The SZ of 7075 upper sheet possessed much higher microhardness than HAZ or TMAZ of 2024 lower sheet.
Kalemba-Rec et al. [
265] reported that for the AS 7075 - RS 5083 welds obtained using the Triflute pin, the profiles for all TRSs were close, however, under TRS of 560 rpm, an abrupt hardness drop in the SZ occurred due to the presence of voids in the weld area. For the AS 5083- RS 7075 configuration, the hardness profiles were different. Under a TRS of 280 rpm, the hardness profile was like that for the AS 7075-RS 5083 configuration. For other speeds, the hardness in the stir zone lowered to 80 HV characteristic of the base 5083 alloy bands appearing in SZ. For both configurations, the maximum hardness in the stir zone was of 150 HV remaining constant from the weld center up to approaching the 7075 alloy. Regardless of the alloy configuration, on the 5083-alloy side, the hardness was about 80 HV remaining constant transverse the weld. On the 7075-alloy side, the hardness lowered from 150 HV to 120 HV and then enhanced to 160 HV (characteristic half of a W-shape). Hardness profiles for the case of the threaded taper tool were like those for the Triflute pin, but values of hardness were higher. The maximum hardness in the weld center can reach a value of 180 HV for the AS 7075-RS 5083 joints, while for the AS 5083–RS 7075 configuration, it was about 160 HV.
Investigating FSWed joints of components made of Al-Mg
2Si/5052 alloys, Huang et al. [
265] reported that the hardness gradually enhanced from the BMZ of the 5052 to the welded joint to the Al- Mg
2Si BMZ.
Studying FSWed joints of components made of dissimilar 2024/6061 alloys, Moradi et al. [
270] found that the microhardness profile on the advancing side was almost identical, while it comprised three distinguishable regions on the retreating side.
Investigating double-sided FSWed joints of components made of 6082-T6/7075-T6 alloys, Azeez and Akinlabi [
272] reported irregular profiling in the Vickers hardness distribution, contrary to the conventional 'W'-shape trend, due to the difference in the chemical composition of alloys and the rate of precipitation. The microhardness evolution deviated from the conventional trend.
Studying FSWed joints of dissimilar 6061-T651/5A06-H112 alloys, Peng et al. [
273] reported that the nano-hardness for each zone varied according to relation BM > NZ > HAZ for the 6061 side, indicating that the mechanical properties of 6061 were weakened after FSW. On the 5A06 side, the change of mechanical properties for each zone was small after FSW. The nano-hardness in NZ and TMAZ was slightly higher compared to that in BM. The mechanical properties of 6061 were more vulnerable to heat input than those of 5A06.
Investigating FSWed joints of components made of 2219-T87 and 2195-T8 alloys, No et al. [
276] found that the microhardness in the upper part of stirring part exhibited even distribution. In the middle- and lower-part hardness on AS with 2195 alloy was clearly higher compared to that on RS of joint. Hardness increased with enhancing TRS and WS.
During studies on the effect of shoulder diameter to pin diameter ratio on microstructure and mechanical properties of FSWed joints of dissimilar alloys 2024-T6 and 7075-T6 Saravanan et al. [
282] placed 2024-T6 alloy on the AS, and 7075-T6 on the RS. They reported that the joints fabricated with the ratios of 2 and 2.5 fractured in heat-affected zone HAZ region of the advancing side, and joints fabricated with the ratios of 3, 3.5, and 4 fractured at stir zone SZ. For all the D/d ratios, minimum hardness was seen at HAZ region in the advancing side and was maximum in the SZ and again decreases at HAZ in the retreating side.
Studying butt FSWed joints of ultrafine-grained UFGed 1050/6061-T6 alloys, Sun et al. [
286] reported that the base metal of both the UFGed 1050 and 6061-T6 alloys exhibited the highest microhardness value. For both materials from the BM to HAZ microhardness lowered gradually. For the 6061-T6 alloy, such a decrease was due to the intensive solid solution of precipitates and the simultaneous occurrence of coarsening of particles resulting from the weld thermal cycles. The stir zone also comprised some regions with a high hardness value like that of the base metal, due to the significantly refined grain size. For the UFG 1050 alloy, the hardness was lowered due to the grain growth and the dislocation density.
Investigating FSWed joints of components made of dissimilar 2024-T3/ 2198-T3 alloys Texier et al. [
287] found significant differences appearing between hardness and local tensile properties.
During studies on butt FSWed joints of components made of dissimilar 6061/7050 alloys, Rodriguez et al. [
289] found that due to the distinct mechanical properties of the two alloys, a consistent asymmetric microhardness distribution profile appeared across the weld nugget, independently of TRS.
Ghosh et al. [
302] found that the microhardness profile was related to welded joint microstructure. Low hardness of A356 alloy appeared at RS. Enhancement in hardness near the weld line occurred due to the composite microstructure affected by both alloys. Further increment at AS appeared due to the higher strength of 6061alloy with respect to A356 alloy.
Studying FSWed joints of plates made of dissimilar 2219-T87/ 5083-H321 alloys, Koilray et al. [
304] found the lowest hardness in the weldment occurred in the heat-affected zone on alloy of 5083 side.
Kim et al. [
307] reported that the 5052 and A5J32 base materials had hardness values of 72 HV and 78 HV, respectively. The hardness in the welded zone of A5052 were lower compared to that of the base metal 5052 due to the dissolution of the second phase particles and annealing during the welding progress. When 5052 was fixed on the RS, the hardness in the vicinity of the shoulder exceeded that of the base metal 5052. The flow of the softened 5052 on RS was restricted by the material on the advancing side A5J32, causing the concentration of work hardening. The hardness values in the welded zone of A5J52 exceeded that of the base metal A5J52_78HV, due to the interaction of the recrystallized fine-grain microstructure and agglomeration of the precipitates. When 5052 was fixed on the RS, excessive agglomeration occurred in a narrow region, due to the restricted flow, and higher hardness occurred compared with the other region. Therefore, the hardness of A5J32 in the welded zone significantly exceeded that for the A5J32 Al alloy fixed on RS.
Studying FSWed joints of sheets made of 6013-T6 alloy, Kafali and Ay [
376] reported that the average hardness of the base material reached 130 HV while for the weld nugget it was 100 HV. The average hardness in the TMAZ was lower than in the weld nugget .
During studies on butt FSWed joints of plates made of dissimilar 2014-T6/ 6061-T6 alloys Jonkheere et al. [
300] found that the welds’ hardness profile were affected by the proportion of each alloy included in the stirred zone, due to the difference between the softening temperatures of both alloys. The 6061 alloy's HAZ was the weak link in all dissimilar welds. The alloys placement or tool lateral shift affected the welds hardness as they influence the precipitate radius and volume fraction.
Investigating butt Fessed joints of sheets made of the dissimilar 2014-T3/5059-H11 alloys, Saleh [
377] found that the TMAZs and HAZs of 2014-alloy possessed the lowest hardness values. The hardness lowered through the weld zone compared to both base metals .
Studying FSWed joints of components made of 2014-T6 alloy Aydin et al. [
337] found that the hardness in the softened weld region lowered with a decrease in the WS.
Investigating FSWed of 4 mm-thick plates made of 2024 alloy, Weglowski et al. [
340] reported that the hardness profile of welds had a characteristic run, typical for FSWed joints.
Studying FSWed joints of 6 mm thick plates made of alloys 2024-T351 Milčić et al. [
342] reported that the distribution and allocation of microhardness were affected by the level of temperature and plastic deformation being highest under the tool shoulder and around the pin.
Investigating FSWed joints of plates made of 2024-T4 alloy Nejad et al. [
341] reported that joints obtained with the threaded tool exhibited elevated average hardness over various welding zones in comparison to the joints prepared by unfeatured tool.
Investigating FSWed joints of components made of 2219-T87/2219-T62 alloys, Venkateswarlu et al. [
378] found that the hardness distribution in the stir zone differed significantly for two different heat-treatment material conditions 2219-T62 against 2219-T82.
Studying stress corrosion cracking SCC of FSWed joints of 2014-T651 alloy, Kollapuri [
379] reported that at 70 % yield, stress induced was lower and so the material failure was determined by its hardness.
Investigating FSWed joints of 8 mm-thick plates made of 2014-T6 alloy Lin et al. [
343] found that the different regions of the joint exhibited different microhardness distribution affected by different thermomechanical actions therein.
During studies on FSWed joints of components made of dissimilar alloys 2017A-T451 and 7075-T651 Hamilton et al. [
270] found that the positron lifetime profiles across the weld comprised many local maxima and minima on the advancing and retreating sides, corresponding to the hardness behavior. Such variations in positron lifetime and hardness away from the weld center were due to the temperature distribution in these areas relative to the critical temperatures for secondary phase nucleation and/or dissolution in the two alloys.
Studying FSWed sheets made of dissimilar alloys 2017A-T451 and 7075-T651 Hamilton et al. [
326] reported that during the flow of surface material into the welded sheet thickness, mid-plane and bottom-plane material flew toward the sheet surface and embraced the surface material core. Within such region, the weld temperatures exceeded the equilibrium θ phase in 2017A, lowering the hardness, and simultaneously dissolved the equilibrium η/T phase in the 7075, causing re-precipitation of GP zones upon cooling and a hardness recovery.
Investigating FSWed sheets of dissimilar 2017A-T451/ 7075-T651 alloys, Hamilton et al. [
46] found that near the weld center, process temperatures allowed the fully dissolving of the equilibrium η phase in 7075 and the partially dissolving of the equilibrium S phase in 2017A. Upon cooling hardness recovered for both alloys. Due to the more complete dissolution of the equilibrium phase in 7075, the hardness recovery skewed toward the AS or the RS of the weld of the 7075 workpiece.
During studies on FSWed joints of components made of dissimilar wrought 2017A and cast AlSi9Mg alloys Kopyscianski [
277] reported that the hardness of the base materials was 80 HV1 and 136HV1 for the AlSi9Mg and 2017A alloys, respectively. The local maximum on the advancing side was on the nugget side with a high density of the bands of the 2017A alloy.
Studying FSWed joints of components made of 2017A alloy Mroczka et al. [
311] reported that microhardness in the cross-section of the joints only slightly varied; however, after the artificial ageing process hardness enhanced. The variation in hardness of the joint after the ageing process pointed out post-process partial supersaturation in the material and higher precipitation hardening of the joint. For the FSWed joints of components made of a cast AlSi9Mg and 2017A alloys Mroczka [
310] reported that the hardness distribution within the weld nugget zone revealed a low strengthening of both cast and wrought alloys. A metastable state of the 2017A alloy occurred, although, the alloy hardness enhanced within the weld nugget zone due to the natural ageing. The hardness of the heat-affected zone in such an alloy slightly changed also due to the natural ageing.
Investigating the stud joints of 2017 alloy obtained by the friction welding process Morikawa et al. [
390] reported that at the weld interface it was formed a stair zone with a hardness close to that of base metals, while the heat-affected zone of the bar and the plate was softened.
Studying FSWed joints of 2024-T6 alloy Sun et al. [
329] reported that with the conical cam thread stirring head the obtained hardness was lowest at the junction of the heat-affected zone HAZ and the thermo-mechanically affected zone TMAZ. The hardness obtained with the conical cam thread at that point exceeded that of other stirring heads.
Investigating the FSWed joints of components made of 7003-7046 dissimilar alloys Yang et al. [
312] found that the hardness was much higher on the retreating side 7046 alloy side than that on the advancing side 7003 Al-alloy side, and the average hardness difference between the two sides was about 30HV. After artificial ageing, the hardness enhanced significantly, while the hardness difference rose to about 50HV for the two sides.
During studies butt FSWed joints of components made of dissimilar 7075-O/6061-O and 7075-T6/6061-T6 alloys Ipekoglu and Cam [
294] found that a hardness enhanced in the joint area for the joints produced in the O-temper condition, whereas a hardness loss occurred in the joint area of the joints formed in the T6-temper condition.
Studying FSWed joints of components made of dissimilar cast Al–Si alloys A319 and A413, Karam et al. [
293] reported that the average hardness of the welded regions enhanced with the rise of the WS and/or lowering the TRS.
Investigating the FSWed joints of components made of dissimilar 6013-T4 and 7003 alloys Zhao et al. [
314] found that independently on the AS or RS, the 6013-T4 side was the weak region for the hardness. The fracture position coincided with the minimum hardness position.
During studies on FSWed components of the heat treatable 6061 and non-heat treatable 5086 alloys Ilagovan et al. [
290] found that the use of threaded tool pin profile allowed obtaining higher hardness values of 83 HV in the stir zone compared to those of the other two pin profiles. The enhanced hardness resulted from the formation of fine grains and intermetallics in the stir zone.
Yoon et al. [
315] found that when the soft material was on the top, the softening material and the deformed surface height resulted from friction heat generation by the rotating shoulder. The more influencing deformed surface height was lowered with the enhancement of revolutionary pitch. When the soft material was at the bottom, the movement of the un-bonded line and hooking appeared resulting from the vertical flow of the rotating tool pin. The more influencing un-bonded line appeared along the interface between two materials deformed toward the hard material.
Studying spot FSWed joints of sheets made of 5454 alloys, Choi et al. [
353] reported that in all cases, the average hardness in the friction-stir-spot-welded zone exceeded that of the base metal zone.
Investigating FSWed joints of components made of 5052-H32 to 6061 T6 blanks, Doley and Kore [
281] reported that microhardness values of the dissimilar welds were lower at heat-affected zones HAZ on both the sides of weld line, whereas the lowest one occurred at HAZ of 5052.
Studying the FSWed joints of components made of AA1100 alloy Selvarajan and Balasubramanian [
354] reported that hardness value of 67 HV in the stir zone region was showed by the FSW joints obtained under the optimized welding parameters and tool material hardness of 45.4 HRC.
For butt FSWed joints of components made of 6063/ 5083 alloys, Kumar and Kumar [
335] reported that the joints with maximum hardness were fabricated at the TRS of 1000 rpm with a cylindrical profile. The hardness enhanced with the rise of the TRS.
Studying FSWed joints of 6061-T6 alloy Juarez et al. [
316] reported that the micro-hardness at the stir zone was of 85 HV for the base material welded BMW, 109 HV for the material with heat treatment before welding HTBW, and 134 HV for the material with heat treatment after welding HTAW. For the case of HTAW, the micro-hardness exhibited the lowest dispersion of values between 124 HV and 148 HV along the four characteristic zones. The hardness of the BMW case was much lower compared to the base material due to the ageing of the material and the thickening of the precipitates resulting from the mechanical work and heat generation during welding. The hardness for the HTAW case exceeded that of the base material due to a uniform distribution of precipitates in the zone of agitation inside the welded zone, combined with a smaller size of precipitates.
Dixit [
391] studied the effect of different pin profiles including straight cylindrical, triangular, and square on microhardness of butt FSWed joints of 4 mm thick strips made of AA1200 alloy. The joints were obtained with the help of high carbon high chromium alloy tool of various pin profiles under two different TRSs. The hardness of the stir zone varied with position and ranged from 30 HV to 40 HV and was higher than that of the parent metal equal to 32 HV. It was due to grain refinement affecting material strengthening and since the grain size in the friction stir zone was much finer than that of parent metal thus enhancing the hardness of FSZ. Also, the small particles of intermetallic compounds also increased the hardness.
Attah et al. [
392] studied the influence of AISI H13 steel-tapered tool on the FSWed joints of components made of dissimilar 7075-T651 and 1200-H19 alloys. They found that the hardness values were of 50 and 175 HV for AA1200- H19 and 7075-T651, respectively, under three WSs including 30, 60 and 90 mm/min, at a constant TRS of 1500 rpm, a tool tilt angle of 2°. The hardness enhanced with the WS from 81.99 to 98.5 HV as the speed raised from 30 to 60 mm/min and lowered to 77 HV at 90 mm/min. The hardness at 1500 rpm and 60 mm/min enhanced from 70.22 to 98.58 HV with increasing the tilt angle from 1- 2°, a further increase from 2–3° lowered the hardness to 66 HV.
Studying FSWed 5 mm thick plates made of dissimilar 5086-O/6061-T6 alloys Aval et al [
358] found that the hardness profile on 6061 side quickly lowered. Such hardness variation was smoother for samples FSWed by the tool with a concave shoulder and a conical probe.
Investigating FSWed joints of components made of dissimilar 7075/ 5083 alloys, Dewangan et al. [
387] reported that the hardness of weld region reached value 116 HV.
Studying FSWed joints of components made of dissimilar 6061/ 7075 alloys Guo et al. [
384] found that all joints failed in HAZ on the 6061 side where hardness minima was located regardless of the relative materials position or the welding process parameters.
Studying the FSWed joints of 5 mm thick 7075-T651/ 2024-T351 similar and dissimilar alloys Zhang et al. [
319] found that the hardness enhanced and then lowered from the top to the bottom along the welding center thickness direction. The tensile fracture locations coincide with that of minimum hardness values at various TRSs.
Tensile strength and residual stresses
The joint strength raised with the rotation speed due to the increased material mixing effect [
267,
297,
299,
306] tool rotation speed increased plastic deformation and WS governed the thermal cycle, residual stresses, and rate of production. The selection of the appropriate combination of such speeds strongly affected weld quality or joint strength.
Studying the butt FSWed plates of 2219-T62 alloy, Xu et al. [
359] found that the residual stresses on the top surface reached about 171 MPa, while only 243 MPa for the weld with tunnel defect and had the conventional “M” profile with tensile stress peaks in the HAZ zone. Those on the bottom surface exhibited the inverted “V” profile with tensile stress peaks of 99.4 MPa in the weld centre.
Bijanrostami et al. [
279] studied the 6061-7075 joint and found that maximum joint strength was reached under a combination of moderate rotation and low WS. However, the maximum joint strength of an A356-6061 joint was reached under low rotation and WS by Ghosh et al. [
302,
303]. The fine grain size, fine distribution of Si particles and lower residual stresses in the SZ appeared under low rotation and WSs. Together with rotation and WSs, the effect of tool geometry like the pin profile or features [
264,
267,
290,
291,
306] pin shapes [
280,
299] and shoulder diameter to pin diameter ratio [
282,
304] influenced joint strength. The pin profile or feature governed material flow and mixing at the joint interface, the pin shape influenced SZ size and material movement, while the shoulder-to-pin diameter ratio affected frictional heat generation between the tool and the BM. The conical threaded pin was the best configuration for the 6061–5086 joint, as it provided uniformly distributed precipitates and the generation of the onion rings resulted from appropriate material mixing in the SZ, as reported by Ilangovan et al. [
290] The tensile strength of the dissimilar FSWed Al joints depended on the microstructure evolution during FSW, which in turn was influenced by the heat input controlled by the welding parameters.
Studying FSWed joints of components made: the first of 2024-T351/ 5083-H112 alloys and the second of 7075-T651/ 2024-T351 alloys, Niu et al. [
262] reported that the tensile strength and elongation of FSWed joints were much deteriorated in comparison to the weaker BM, especially for 2024-T351/ 5083-H112 joint.
Investigating the FSWed joints of 7075-T651 and 2024-T351 alloys Hasan et al. [
130] found that placing the softer 2024 alloy on the AS slightly enhanced the tensile properties of welding joints. The introduction of pin tool flute/flat improved the ultimate tensile strength and elongation of welds regardless of the fixed location of base materials. Using a truncated threaded pin tool with a flute of radius equal to that of the pin the tensile strength of weld reached the maximum value of 424 MPa, which represents an efficiency of about 94.3% with respect to the softer material.
Ge et al. [
61] studied how EST affects the shear failure load of lap joints. Shear and tensile fracture modes can occur. Mode I was the shear fracture mode, in which the failure occurred along the original lap interface of two sheets. This mode occurred in lap joints when the mixing of materials between the upper and lower sheets hardly occurred and the nature of the joining mechanism at the interface was close to diffusion bonding. It was combined with the alclad layer owning low strength.
The tensile fracture mode exhibited two different fracture paths [
54]:
Mode II - wherein the crack initiated from the tip of the cold lap defect CLD, propagated upward along the SZ/TMAZ interface, and finally fractured at the top surface of upper sheet,
Mode III - wherein the crack initiated from the tip of the hook defect HD, propagated downward along the HAZ/TMAZ interface, and fractured at the bottom surface of lower sheet. Such two different fracture modes are strongly affected by the size and orientation of the HD or CLD. The cracks occurring in the HD and the CLD of the lap joint continued their propagation upwards or downwards when the lap joint underwent tensile stress during the tensile test.
Studying FSWed joints of components made of 7075-T651/ 5083-H111 alloys, Kalemba-Rec et al. [
265] reported that defect-free joint of the highest tensile strength 371 MPa was obtained with 5083 alloy on the AS, 7075 alloy on the RS, a TRS of 280 rpm, and the Triflute pin.
Investigating FSWed joints of components made of Al-Mg
2Si alloy and 5052 alloy Huang et al. [
269] reported that the UTS and elongation of the welded joint were greater than those of the base material of the Al-Mg
2Si, whereas they were lower than those of the 5052-base alloy. The WE exceeded 100% relative to the softer material and reached 61% relative to the 5052 alloy. The particle-matrix interfacial debonding fracture mechanism occurred.
Studying FSWed joints of dissimilar 6061-T651/ 5A06-H112 alloys, Peng et al. [
273] noticed that after welding, the yield strength of 6061 lowered by 50% to about 115 MPa and the ultimate tensile strength decreased from 277 MPa to about 190 MPa mainly due to the unstable work hardened state of rolled 6061 was destroyed by elevated temperature generated in FSW. The WE reached 68.5% relative to 6061 alloy, and decreased from 70% to 68% with increased the ratio of rotation speed and WS from 4 to 12 r/mm.
Studying FSWed joints of dissimilar 2024-T3/ 6063-T6 alloys, Sarsilmaz [
275] found that under TRS of 900 rpm and WS of 200 mm/min the highest tensile strength of 348 MPa was obtained which was 74% of the tensile strength of the 2024 base metal. An enhancement in tensile strength was 45% higher than the base tensile strength of 6063. All tensile failures occurred at HAZ location always at the 6063 side.
Investigating FSWed joints of components made of 2219-T87/ 2195-T8 dissimilar alloys No et al. [
276] found lack of correlation between tensile strength and WS, however, tensile strength increased with enhancing TRS, up to 800 rpm.
During studies on FSWed joints of components made of dissimilar wrought 2017A/cast AlSi9Mg alloys Kopyscianski et al. [
277] reported that the tensile strength was of 132 MPa, while the elongation was below 1 %.
Studying FSWed joints of dissimilar 6082-T6/ 5083-H111 alloys, Kasman et al. [
280] obtained the highest tensile strength for the weld joint obtained using a triangular-shaped pin and the UTS was of 198.48 MPa. At a lower TRS and WS for each tool pin shape, lower UTS values appeared. The UTS enhanced with increasing the TRS and the WS, while keeping their υ ratio constant for the triangular-shaped pin. The WE varied from 55% to 68% depending on both presence of defects in the weld joint and the strength of the base material.
Saravanan et al. [
282] reported that the tensile strength enhanced with increasing in shoulder diameter to pin diameter ratio up to 3 and then lowered by further enhancement in the ratio. The maximum tensile strength was of 356 MPa at the ratio of 3, while the lowest one of 316 MPa at the ratio 4. The WE varied from 76 % to 86 %, depending on the mentioned ratio.
Studying FSWed joints of sheets made of dissimilar Al-Mg-Si/Al-Zn-Mg alloys Yan et al. [
283] found that the tensile strengths of the dissimilar Al-Mg-Si/Al-Zn-Mg joints using both configurations exceeded that of the Al-Mg-Si FSW joint.
According to Sun et al. [
286] FSWed joints produced at the revolutionary pitches of 1.25 and 1 mm/rev, exhibited both the tensile strength and elongation lower than those obtained at the smaller revolutionary pitches. Due to the insufficient mixing of the two materials in the stir zone and a couple of micro-defects at the 6061-T6 alloy zone the strength of the joints was lowered. Decreasing the revolutionary pitch to 0.75 or 0.5 mm/rev enhanced the heat input, intensifying the plastic deformation and the mixing of the two materials in the stir zone. This resulted in a more homogenous microstructure of the stir zone increasing the tensile strength and elongation increased to about 110 MPa and 13 % for the case of 0.75 mm/rev and 110 MPa and 22.5 % for the case of 0.5 mm/rev. The largest joint efficiency was 55 % with respect to the UFG 1050 Al base metal. Remarkably high heat input in the joints produced at 0.5 mm/rev caused grain growth of both materials causing again a small decrease of the tensile strength compensated by a much-enhanced elongation.
During studies on butt FSWed joints of components made of dissimilar 6061/ 7050 alloys, Rodriguez et al. [
289] found that under monotonic tensile loading, the joint strength enhanced with the rise in the TRS. The WE reached up to 62 %.
During studies on FSWed components of the heat treatable 6061 and non-heat treatable 5086 alloys Ilagovan et al. [
290] found that the use of threaded tool pin profile allowed obtaining higher tensile strength of 169 MPa compared to those of the other two pin profiles. The lowered size of weaker regions, such as TMAZ and HAZ ones, caused higher tensile properties. The WE varied from 50.4 % to 67.6 %.
Studying FSWed joints of components made of dissimilar cast Al–Si alloys A319 and A413 Karam et al. [
293] found that the welded joints exhibited better tensile properties than the base alloys. The A413 base alloy exhibited lower ultimate tensile and yield strengths when compared with A319 base alloy, thus under tensile the welded specimens fractured at A413 alloy side.
During studies butt FSWed joints of components made of dissimilar 7075-O/6061-O and 7075-T6/6061-T6 alloys Ipekoglu and Cam [
294] found that the strength values of all the O-joint specimens were close to those of the 6061-O BM, and all the specimens failed within the 6061 BM side away from the joint area. This was due to the shielding effect provided by the strength overmatching resulting from the grain refinement or precipitation of strengthening particles in this zone during FSW process of Al alloys in the O-temper condition, i.e., softened state. The WE in case of initial O state was about 100 %. After PWHT (T6 treatment) the highest WE of about 93 % for the 1000/150-PWHT specimens, and of 87.5 % in case of 1500/400-PWHT specimens. In case of initial T6 state the highest WE was of about 80 % for the 1000/150 specimens, and of 67.8 % for the 1500/400 ones. After PWHT the WE was of 89.1 % for 1000/150-PWHT specimens and of 90.8 % for the 1500/400-PWHT ones.
During studies on butt FSWed joints of 4.76mm thick sheets made of dissimilar 6061-T6 and 7075-T6 alloys Cole et al. [
295] found that weld tool offsets into the retreating side 7075 enhanced the tensile strength of the dissimilar joint. Such an enhancement was facilitated by lower average weld temperatures with the enhanced amount of 7075 stirred into the nugget. The WE enhanced with a lowered amount of power input to the weld, where the subsequent WE was highly affected by the alloy most sensitive to heat input and weld temperature.
Ghosh et al. [
302] found that tensile properties of welding nuggets WN were highly dependent on its microstructure. Kim et al. [
307] reported that the tensile strength exhibited similar values, regardless of the arrangement of the materials. The welding defects occurring under welding conditions with a lower heat input did not affect the tensile properties. When A5J32 was fixed on the retreating side, the highest strength of the welded joints equal to 224.1 MPa appeared under conditions of 1000 rpm and 300 mm/min.
Prime et al. [
308] studied FSWed butt joints of components made of 7050-T7451/ 2024-T351 alloys. They found that the stresses in the test specimen removed from the parent welded plate reached values up to 32 MPa and had the “M” profile with tensile stress peaks in the heat-affected zone outside the weld. The peak residual stress values were below 20% of the material yield strength. Such low stresses were achievable only by solid state welding with less distortions, while for fusion welding, such a low value was hardly possible. The fatigue behaviour is strongly affected by these low values of residual stresses. The peak tensile residual stresses appeared in the HAZ on both sides due to local frictional heating at the tool material interface. Tensile residual stress resulted from that the hotter material was forced by the other material during welding.
Studying the butt FSWed joints of dissimilar 2024-T351/ 6056-T4 alloys, Amancio-Filho [
260] reported that the tensile strength of the weld joint was up to 90% of the 6056-T4 alloy. Fracture occurred in the thermo-mechanically heat-affected zone of the alloy 6056-T4, where annealed structure led to lowered microhardness. The drop in tensile strength and the associated increase of strain appeared in the regions of microhardness drop. The obtained joint efficiency in terms of the UTS was of 55.8% for 2024-T351 alloy and of 71.4% for 6056-T4 alloy. However, the joint efficiency in terms of elongation at the rupture was poor (9-14%).
Ivanov et al. [
334] studied FSWed joints of rolled sheets made of 2024/ 5056 alloys obtained for various thicknesses. For weld joints with tensile strength not less than 0.9, the welding process parameters were complexly affected by the tensile strength of the base metal.
Investigating butt FSWed joints of components made of 6063/ 5083 alloys, Kumar and Kumar [
335] reported that the joints of higher tensile strength were fabricated at the TRS of 1000 rpm with a cylindrical profile. The tensile strength enhanced with the rise of the TRS. The WE varied from 32.3% to 43%, when TRS enhanced from 600 to 1000 rpm.
Studying of FSWed joints of components made of dissimilar 5083-H111/6351-T6 alloys, Palanivel et al. [
306] reported that the TRS and pin profile influenced the joint tensile strength because of varying material flow, loss of cold work in the HAZ of 5083 side, dissolution and over aging of precipitates of 6351 side and formation of macroscopic defects in the weld zone.
Investigating the FSWed dissimilar cast and wrought 6061 alloy Dinaharan et al. [
305] found that the joint exhibited maximum tensile strength when cast Al alloy was positioned on the advancing side at all TRSs. Studying FSWed joints of components made of Al- Mg
2Si alloy and 5052 alloy Huang et al. [
289] found that the UTS end elongation of the welded joint were greater than those of the base material of the Al- Mg
2Si, whereas they were lower than those of the 5052-base alloy.
Saleh [
380] studied the microstructure and mechanical properties of butt FSWed joints of sheets made of the dissimilar 2014-T3/ 5059-H11 alloys obtained by bonding the two materials perpendicular to their rolling directions. They found that the ultimate tensile strength values of the dissimilar joint varied in range from 54% to 66% of those of the base metal.
Sasikala et al. [
336] studied the effect of rotational and cross-sectional speeds, geometry, and the tool-to-pin-diameter ratio on the tensile strength of the FSWed joints of plates made of dissimilar 2618-T87/ 5086-H321 alloys. Heat affected zones with tensile failures appeared on the alloy 5086 side of the weldment.
Lin et al. [
343] studied FSWed joints of 8 mm thick plates made of 2014-T6 alloy. They found that the weld tensile strength was affected by welding parameters. The maximum ultimate tensile strength of 360 MPa equal to 78% that of base metal appeared at a rotating speed of 400 rev min
−1 and WS of 100 mm min
−1. The different regions of the joint exhibited different tensile strengths dependent on the microstructure variation and microhardness distribution affected by different thermomechanical actions therein.
Morikawa et al. [
390] studied the strength of the stud joints of 2017 alloy obtained by the friction welding process. They reported that the tensile strength of joints enhanced with pressure and friction time, and the highest tensile strength reached 275 MPa (63.1 % joint efficiency for the bar base metal). The fatigue strength of joints enhanced under their high tensile strength.
Studying the FSWed joints of 2024-T6 alloy Sun et al. [
329] found that the tensile strength for all joints was more than 80% of the BM, and the maximum tensile strength of the joint welded with the conical cam thread tool pin reached 364.27 MPa, which was 86.73% of the base metal BM. The elongation after break reached14.95 %. All joints were tensile fractured by plastic fracture.
Investigating the FSWed joints of 3003 alloy Chekalil et al. [
347] found that the tensile properties of joints remained good. The tensile strength of the weld joint was up to 75% of that of the base metal.
Kasman and Ozan [
348] studied the influence of welding process on the structural properties of the butt FSWed joints of plates made of 3003-H24 alloy. They found that the highest ultimate tensile strength among all the welded joints equal to 128 MPa was obtained under 50 mm/min WS and 800 rpm TRS. The WE was remarkably close to 100%. The size of the defects was affected by the tensile properties of welded joints.
Aydin et al. [
338] studied the effect of welding parameters (rotation speed and WS) on the mechanical properties of FSWed joints of components made of 3003-H12 alloy joints. The tensile weld strength enhanced with an increase in the WS or a decrease in the rotation speed. The tensile fractures of the joints were in base metal under welding parameter combinations of 1070 rpm and 40 mm/min or 2140 rpm and 224 mm/min. The ultimate tensile strengths of the joints lowered linearly with an enhancement of rotation speed at a constant WS, while such strength of the joints enhanced almost linearly with a rise of WS at a constant rotation speed.
Goyal et al. [
350] studied the effect of welding parameters on the UTS of FSWed joints made of 3003 alloy. The best UTS equal to 127.2 MPa was obtained for the process parameters including a WS of 74.64 mm/min, a TRS of 971.77 rpm and a tool tilt angle of 1.52, respectively. The WE reached up to 89.4%.
Deng et al. [
381] studied the effect of post-weld heat treatment on the microstructure and mechanical properties of the FSWed joints of plates made of 7204-T4 alloys. They reported that the ultimate tensile strength UTS of the FSW joints were 296.6, 318.2, 357.4 MPa under the heat treatments of AW, AA, and SAA, respectively .
Yang et al. [
312] studied the influence of post-weld artificial ageing on microstructure and mechanical properties of FSWed joints of components made of 7003-7046 dissimilar alloys. They found that after artificial ageing, the tensile strength enhanced slightly, and the elongation was slightly changed for the joint. The WE slightly exceeded 100% in both cases of natural and artificial ageing.
Kasman and Ozan [
313] studied butt FSWed joints of 6013 Al plates obtained via pin offset technique. They found that the highest tensile strength equal to 206 MPa was obtained under 1.5 mm pin offset towards the advancing side and 500 min
−1 TRS, leading to the ratio of tensile strength of the joint to the ultimate tensile strength of the base metal (joint efficiency) equal to 74 %.
Zhao et al. [
314] studied the FSWed joints of components made of dissimilar 6013-T4/7003 alloys. The joint with the 6013-T4 placed at the advancing side AS was called the A6R7 joint. Accordingly, the A7R6 referred to the joint with the 7003 placed at the AS. The authors reported that independently on the AS or RS, the 6013-T4 side was the weak region for tensile strength. The WE for joint A6R7 was of about 93%, while that for the joint A7R6 was of 87 %.
Studying FSWed joints of sheets made of dissimilar Al-Mg-Si/Al-Zn-Mg alloys Yan et al. [
283] found that the tensile strengths of the dissimilar Al-Mg-Si/Al-Zn-Mg joints using both configurations exceeded that of the Al-Mg-Si FSW joint.
Liu et al. [
309] studied mechanical properties of FSWed joints of components made of 1050 - H24 alloy. They reported that the maximum tensile strength of the joints was equivalent to 80% of that of the base material. Under deviation of the welding parameters from the optimum values, the tensile properties of the joints deteriorated, and the fracture locations of the joints varied.
Studying butt FSWed joints of the ultrafine grained UFGed 1050 alloy plates with a thickness of 2 mm with the 2 mm-thick 6061-T6 alloy plates Sun et al. [
286] found that the maximum tensile strength reached about 110 MPa. The WE reached up to about 55%.
Investigating FSWed joints of 1100 alloy Selvarajan and Balasubramanian [
354] reported that a maximum tensile strength of 105 MPa was showed by the FSW joints obtained under the optimized welding parameters. The WE reached up to 95.4%.
Studying the underwater FSWed joints of dissimilar 6061/7075 alloys Bijanrostami et al. [
279] found that the maximum tensile strength of 237.3 MPa and elongation of 41.2% were reached under a TRS 1853 rpm and a traverse speed of 50 mm/min. In comparison with the optimum condition, greater heat inputs induced lowered joint strength and the higher elongations for the joints.
Investigating welded joints of components made of dissimilar 7003-T4/6060-T4 alloys obtained by underwater friction stir weld UFSW Dong et al. [
355] reported that the ultimate tensile strength of the joints reached up to 185 MPa. The strength increased due to the microstructure modification caused by water cooling. The WE was of 90.4 % and was higher compared to the case of the classic FSW process.
Jassim and Al-Subar [
393] studied FSSWed joints of 3mm thick sheets made of 1100 alloy by overlapping the edges of sheet as lap joint. The joint tensile strength enhanced with raising the TRS and the maximum value of tensile strength equal to 233 MPa, twice higher than the one of base metal, occurred at a TRS of 2000 rpm. The WE varied in the range of 74.5-141 %.
Studying the FSWed joints of components made of 1100 alloy Senapati and Bhoi [
394] reported that the UTS of the welded specimen enhanced by 20% compared to that of the parent material due to the uniform dispersion of silicon particles present within the base material. Studying the butt FSWed joints of 3 mm thick strips made of 1200 alloys Joseph et al. [
395] found that the ultimate tensile strength of the welded region lowered due to insufficient mixing of the material or due to heat evolved during friction stir welding. The lower feed rate provided a joint with higher tensile strength due to better mixing of the material.
Investigating the FSWed joints of components made of dissimilar 7075-T651 and 1200-H19 alloys Attah et al. [
392] found that the UTS enhanced from 126.04 to 151.54 MPa with an increase in the WS from 30–60 mm/min and lowered to 128.37 MPa at 90 mm/min. The UTS increased from 123.32 to 151.54 MPa as tilt angle increased from 1–2° and lowered to 122.2 MPa as tilt angle enhanced to 3°.
For FSWed joints of 6061-T6 alloy Juarez et al. [
316] reported that the tensile strength for the BMW case was close to that in joints obtained by fusion welding. For the HTBW case, the tensile strength was enhanced by 10% compared to that obtained in BMW. For the HTAW case, an efficiency of 96% of tensile strength compared to that of the base material was observed.
Studying FSWed joints of dissimilar 7075-T651 and 1200-H19 alloys Attah et al. [
392] found the impact energy enhanced from 12.9 to 21.4 J with an increase in the WS from 30 to 60 mm/min and lowered to 5.4 J at 90 mm/min. The UTS enhanced from 126.04 to 151.54 MPa with an increase in the WS from 30–60 mm/min and lowered to 128.37 MPa. The UTS increased from 123.32 to 151.54 MPa as tilt angle increased from 1–2° and lowered to 122.2 MPa as tilt angle enhanced to 3°. Under the tilt angle of 2°, rotational and traverse speeds of 1500 rpm and 60 mm/min, respectively, the highest impact energy of 21.4 J was obtained.
Investigating butt FSWed joint between dissimilar 6061-T6/7075-T6 alloys, Godhani et al. [
317] found that during the tensile tests of specimen the fracture took place from the HAZ of the 6061 side under all investigated welding conditions. The breakage in the cup-and-cone form pointed to the ductile nature of the failure. The WE reached up to 61.4%.
Sato et al. [
396] applied FSW to a 1.5 mm thick pieces made of accumulative roll-bonded (ARBed) 1100 alloy with ultrafine grained microstructure and high hardness. Transversely to the rolling direction an elongated ultrafine grained microstructure occurred. These pancake-shaped ultrafine grains with some dislocations and sub-boundaries therein, typically resulted from the ARB process, were surrounded by high angle boundaries with misorientations above 15°. The mean thickness and length of the pancake-shaped grains were 260 and 450 nm, respectively. The initial material hardness was of 30 HV, while after ARB it raised to about 85 HV, due to the grain refinement. The FSWed joints were obtained under rotation speed of 500 rpm, WS of 12 mm/s using the tool with a shoulder diameter of 9 mm, a pin diameter of 3 mm, a pin length of 1 mm and pin threaded. The welding direction was identical to the rolling direction (RD) of the ARB process. The tool-to-workpiece angle was 3° from the vertical axis in the weld. The authors found that FSW suppressed high lowering of hardness in the ARBed material, however the SZ and the TMAZ exhibited small lowering of hardness due to dynamic recrystallization and recovery. The FSW prevented softening in an ARBed Al alloy 1100 with an equivalent strain of 4.8 in the as-ARBed condition.
Studying FSWed 5 mm thick plates made of 5086-O and 6061-T6 alloys, Aval et al.],[
358] , in the case of 5086/6061 joint obtained the weld UTS varying in the range of 219-240 MPa and the weld elongation of 17%/23%, while WE varied in the range of 87-95%. In the case of 6061/5086 joint the weld UTS varied in the range of 228-248 MPa, while WE varied in the range of 90- 98%.
For the FSWed joints of components made of dissimilar 5052/6061 alloys obtained with various pin-eccentric stir tools reported [
318] found that all tensioned joints failed in the NZ, and the joint obtained by the 0.8 mm-pin-eccentric stir tool exhibited the highest tensile strength of 196 MPa due to the raised grain-boundary and dislocation strengthening. The WE reached up to 86%.
Investigating FSWed joints of components made of dissimilar 7075/5083 alloys, Dewangan et al. [
387] reported that the UTS of weld region reached value of 237.4 MPa and WE was of 80%.
Studying FSWed joints of sheets made of 2219 and 7475 alloys, Khan et al. [
388] reported that the UTS of weld region reached value of 267.2 MPa, elongation was of 5%, and the WE was of 57%-92%.
For FSWed joints of components made of dissimilar 6061/7075 alloys, Guo et al. [
384] found that the UTS of joints enhanced with the lowering heat input. The highest UTS was of 245 MPa.Summarizing, it can be noticed that most studies carried out on FSW were focused on BM in the as-rolled condition for 2xxx-5xxx, 2xxx-6xxx, 2xxx-7xxx, 5xxx-6xxx, 5xxx-7xxx Al series. Some studies considered the effect of base material placement of welded joints features and it remains inconclusive. Base material placement is an issue for the cases of high differences in mechanical properties of the BMs as in the 6xxx-7xxx and the 5xxx-7xxx combinations. Some studies concerned welding parameters optimization studies, particularly the effect of tool offset on welds quality. The further studies are necessary including these using microstructure characterization to better recognize the material flow in the SZ.
During FSW process the residual stresses are very low, much low compared to those of the fusion processes.
The higher mechanical properties mainly resulted from the fine grains in the stir zone of FSWed joints.
The FSWed weld (joint) efficiency can widely vary in range of 57-98%, and even exceed 100% relative to the softer material.