Effect of rotational speed on the microstructure and mechanical properties of rotary friction welded AISI 1018/AISI 1020 asymmetrical joints

0
135

Abstract: Asymmetrical joints (joining of the plate with rod) were joined using traditional fusion welding processes. However, the usage of unsuitable filler wire tends to lower weld penetration over the material surface, which also results in the attainment of hot or solidification cracks over the weld surface. To overcome these issues, solid-state welding processes are preferable. This study investigates the rotary friction welding (RFW) of AISI 1018 low carbon steel plate with AISI 1020 low carbon steel rod of asymmetrical joints. The friction welding process parameters such as rotational speed were taken as variable, and other parameters like friction pressure, forging pressure, friction time, and forging time were kept constant in this investigation. The impact of rotational speed on macrostructure, microstructure, and mechanical characteristics of joints such as microhardness, tensile strength, and fractography studies was analyzed. The fractured surface of the tensile specimen was examined through a scanning electron microscope (SEM). The maximum tensile strength of the joint about 452 MPa was observed. Maximum hardness at the weld interface was perceived at about 252Hv. Increasing rotational speed tends to increase the strength of the asymmetrical steel joints in rotary friction welding. 

1 Introduction

Asymmetrical components such as the joining of the plate with rod in carbon steel using rotary friction welding have several applications such as gear shafts, axle shafts, drive shafts in automobiles, and flanges in chemical, oil, and gas refineries. Rotary friction welding is the solid-state welding process in having one end as a rotating chuck at high-speed with the holding of a steel plate and in another end having a horizontally movable stationary part with a holding of a steel rod. At both ends of the materials, they get a higher range of heat by friction. Also, applying additional pressure leads to the forging together. The important process parameters in the rotary friction welding process are forging pressure, friction pressure, forging time, friction time, and rotational speed [1].

Subhash Chander et al. [2] have studied the effect of rotational speed on the dissimilar joining of AISI 304 stainless steel with AISI 4140 low alloy steel by analyzing the mechanical and metallurgical characteristics of the joints. Ananthapadmanaban et al. [3] investigated the different welding conditions for joining mild steel with stainless steel and found variations in the mechanical properties of the joints. Ho Thi My Nu et al. [4] evaluated the influences of process parameters on mechanical strength and found that axial pressure has more influence. There-fore, applying higher forging pressure in the weld region produces fine grains, equiaxed grains, and recrystallized grain structures in the welded zones, which leads to superior tensile strength and microhardness. Alex Anandaraj et al. [5] focused on the joining of dissimilar metals such as martensitic stainless steel and Inconel in rotary friction welding. Observed process parameter influence and high-temperature properties of the dissimilar joints were reported as the rotational speed had a greater influence on the join strength by correlated results of tensile, micro-hardness, and microstructures. Radoslaw Winiczenko [6] reported that the tensile strength decreases by increasing the upset force; hence, the upset force has a negative effect on tensile strength. Wenya Li et al. [7] detailed about effects of welding parameters on axial shortening and temperature using a 2D finite element model in ABAQUS. They reported that axial shortening was having a greater influence due to the attainment of a quasi-stable temperature in the weld interface by increasing the axial pressure. Paventhan et al. [8] investigated the joining of medium carbon steel and austenitic stainless steel in the RFW process using response surface methodology (RSM) to optimize suitable welding parameters to attain maximum tensile strength. Sare Celik et al. [9, 10] developed an empirical relationship using RSM to predict maximum tensile strength and temperature for the joining of AISI 316 stainless steel and Ck 45 unalloyed steel. They observed 5.8% higher tensile strength than the parent metal. Damodaram et al. [11] observed lower ductility and tensile strength than the base metal in the joining of tungsten heavy alloys (WHA) and also higher hardness reported at the weld interface region. Sivaraj et al. [1214] analyzed subsequent coarsening and grain refining in friction stir welded armor grade aluminum alloy and obtained 70% greater join strength than the parent metal strength. Authors also studied influence of welding parameter on mechanical properties and microstructural features of resistance spot welded dual phase steel sheets joint. Sankar et al. [15] evaluated the higher tensile strength and hardness at the weld interface region than the parent metal in the joining of high thickness carbon steel in the GMAW technique. Balakrishnan et al. [16] obtained a tensile strength of about 83% when joining AISI 4340 steel in friction welding. Kirik et al. [17] conducted the friction welding experiment by varying the friction time with constant rotational speed and friction pressure. They observed an increase in weld interface temperature with micro-structural changes by increasing friction time.

Based on an extensive literature review, it is found that many researchers studied the joining of the rod with the rod using the rotary friction welding process. There is no study found on the joining of asymmetrical components using rotary friction welding. Therefore, the main objective of this study is to investigate the effect of rotational speed on the microstructure and mechanical properties of rotary friction welded AISI 1018/AISI 1020 asymmetrical joints.

Table : Chemical composition of base metal.

Material Cr Mn C Mo Si S P Ni Fe
AISI(plate) . . . . . . . . Bal
AISI(rod) . . . . . . . . Bal

Table : Mechanical properties of base metal.

Material Tensile Yield Elongation in Hardness
strength strength gauge length (Hv)
(MPa) (MPa) of  mm (%)
AISI(plate)
AISI(rod)

2 Experimental procedure

2.1 Materials and methods

The chemical composition (wt%) and mechanical properties of the base material (BM) were given in Table 1 and Table 2, respectively. Figure 1a, b illustrate the optical microstructure of the base metal, which is highly composed of equiaxed ferrite grains with an average grain size of 12 μm along with pearlite. The welding trials were conducted to find a feasible working range for the carbon steel base material in the rotary friction welding machine.

The feasible range was analyzed by viewing the amount of flash formation with test results. From those trial runs and previous literature [18], the constant and varying parameters were taken for the welding, which is shown in Table 3.

AISI 1018 carbon steel plate thickness of about 15 mm with length and breadth about 32 mm and AISI 1020 carbon steel rod about 12 mm diameter with a length of 120 mm used for experimentation. Specimens before welding are viewed in Figure 2 and were welded in a numerically controlled rotary friction welding machine (Make: RV ma-chine tools). To study the effect of rotational speed, five different speeds 800, 900, 1000, 1100, and 1200 rpm are chosen, which were named S1, S2, S3, S4, and S5, respectively. The asymmetrical joints of AISI 1020 and AISI 1018 low carbon steels fabricated by rotary friction welding are shown in Figure 3.

Figure 1: Light microscopy of microstructure, a) AISI 1020 (rod), b) AISI 1018 (plate).

Table : Welding parameters.

  1. No Parameters Range

Rotational speed (rpm) S , S , S ,

S,S

Forging pressure (MPa)
Friction pressure (MPa)
Forging time (s)
Friction time (s)

Figure 2: AISI 1018 plate and AISI 1020 rod specimens used for experimentation.

2.2 Mechanical testing

Microhardness surveys were conducted using Vicker’s hardness testing machine (Make: SHIMADZU, Japan; Model: HMV-T1) to evaluate the hardness of weld zones and base material. The diamond indenter was used with a 0.5 kg load at 15 s dwell time to get the appropriate result.

Figure 3: AISI 1018 plate and AISI 1020 rod welded specimens.

The tensile test of the welded specimens was tested by extracting miniature tensile specimens from the welded joints as per standard ASTM E8 M. The dimensions of the tensile specimens (in mm) were shown in Figure 4, which were tested in a servo-controlled Tinius Olsen (H50KL)

Figure 4: Miniature tensile specimen dimension.

Figure 5: Tensile specimen extraction from welded specimen.

Universal Testing Machine. Tensile specimen extraction from the welded joint is shown in Figure 5. In each parameter condition, three welded specimens are tested, and the average value is tabulated. The fractured surface of tensile specimens was examined with a Scanning Electron Microscope (SEM).

2.3 Microstructural analysis

Welded specimens were prepared as metallographic samples and mirror polished by using different grades of emery sheets.

The etchant was used for obtaining clear microstructure from samples that were 2 vol% of nital reagent [19]. The macro-structure and microstructure of metallographic samples were analyzed by using a stereo zoom and an optical micro-scope at high and low magnifications to evaluate the different weld zones and their structures.

3 Results and discussion

3.1 Tensile properties

The tensile specimens before and after the test and the tensile and yield strengths of welded joints for different rotational speeds compared with base metal strength are given in Figures 6 and 7, respectively. The tensile test results of welded specimens are shown in Table 4. From the table, it is found that tensile strength varies from 284 to 452 MPa. This is mainly due to different rotational speeds. An increase in the rotational speed caused an increase in the tensile strength of the joint. Specimen S5 shows a higher tensile strength of about 452 MPa and the other S1, S2, S3, and S4 are 284 MPa, 326 MPa, 368 MPa, and 410 Mpa, 

   

Figure 6: Tensile specimens, a) specimens before the tensile test, b) specimens after the tensile test.

Figure 7: Comparison of tensile strength of joints with base materials.

Table : Tensile test result of welded specimens.

Specimens Tensile Yield Elongation in Joint
strength strength gauge length of efficiency
(MPa) (MPa) mm (%) (%)
S
S
S
S
S

those zones, so failure occurred in that area. From the fractured surface, a ductile mode of failure was observed. Figure 7

3.2 Fractography

The fractured surface of the S5 (1200 rpm) specimen was analyzed through SEM and is presented in Figure 8. Finer and deeper dimples with tear ridges were observed over the fractured surface area. From Figure 8, it is observed that the fracture occurred in the grain boundaries in an elongated manner, which shows the ductile mode of failure. The crack initiation zone had prominent striation marks with a river-like appearance (Figure 8a). The fractured surface consists of equiaxed dimples in the middle area. The elongated dimples have some parabolic shape at the edges, which shows that the edges are being sheared [20]. More elongation was observed due to the presence of larger and compares the tensile strength of base material with welded specimens, in which the AISI 1018 plate has 440 MPa and the AISI 1020 rod shows 478 MPa, whereas the S5 specimen obtains 452 MPa of tensile strength, which projects that the joint efficiency is 94% in the rotary friction welded asymmetrical joints.

Figure 8: SEM images of fractured surfaces, a) and b) rod side fracture, c) and d) plate side fracture.

more elongated dimples [21]. The presence of more fine elongated dimples over the fractured surface area was the cause of getting high strength and more elongation [22]. Voids present in the fractured surface reveal that the grains were completely separated from the surface. When comparing SEM images of fractured components, the voids were higher in the plate specimen (Figure 8c), which indicates the rod was pulled out from the plate specimen.

3.3 Microhardness analysis

The microhardness profile of welded specimens for different rotational speeds in various regions is illustrated in Figure 9. The figure shows that S5 welded joints have higher micro-hardness in all the regions compared to other welded specimens in that the S5 specimen projects maximum hardness at the weld interface, which is fully deformed zone (FDZ) of about 252 Hv. Maximum microhardness decreased gradually from the weld interface to other weld zones, until finally at base metal, a lower hardness value was observed. In the partially deformed zone (PDZ) region, the hardness range varies from 251Hv to 219Hv and from 251Hv to 203Hv for the rod and plate side, respectively. Similarly, HAZ region 218Hv to 180Hv on the rod side and 202Hv to 170Hv on the plate side were observed. In both PDZ and HAZ, higher hardness range was observed on the rod side compared to the plate side of the specimen. This is mainly due to higher deformation of the rod side leading to the higher flash formation in rotary friction welded asymmetrical joints. When compared with base metal hardness, the weld interface has 31% higher hardness.

Compared to the microhardness profile, the fractured weld specimen shows that the fracture initiates from the HAZ of the plate due to its lower hardness and then prop-agates along with the base metal of the plate because the plate side hardness was lower when compared to the rod.

3.4 Macrostructural analysis

Macrostructures of welded specimens with various rotational speeds, such as 800 rpm (S1), 900 rpm (S2), 1000 rpm (S3), 1100 rpm (S4), and 1200 rpm (S5), were shown in Figure 10. Formation of higher flash due to increasing the rotational speed. Higher flash is observed on the rod side of the specimen. This is mainly due to the smaller surface area of rod contact with the wider space of the plate under forging pressure, which leads to higher flash on the rod. It shows a higher variation in weld interface width, geometry, and also in flash formation. The S1 specimen was not welded sufficiently, and there was no proper flash formation. The S2 joint has lower flash formation and observed minimum tensile strength. To that, the S3 specimen shows average flash content and average weld strength, while the S4 specimen has proper flash formation but tensile strength is less. The S5 welded specimen, having desirable flash content and superior mechanical strength, also welded properly when compared to the other four welded specimens. From macrostructure, it is also observed that the clear weld region is free from weld defects such as cracks and voids due to the deformation of the soft ferrite phase during welding; hence, more material is left out in the form of flash from the interface region [23].

Figure 9: Microhardness profile.

Figure 10: Macrostructures of welded specimens, a) S1 – 800 rpm,

b) S2 – 900 rpm, c) S3 – 1000 rpm, d) S4 – 1100 rpm, e) S5 – 1200 rpm.

From welded joints macrostructure analysis, it was found that heat generation in the interface reaches a higher temperature at a higher rotational speed, which leads to higher plastic deformation in the base metals. Increasing the rotational speed produces more heat in a short period on the weld surface of both materials. Whereas this increase in the rotational speed gets to a lower cooling rate in the weld region, hence heat affected zone becomes wider [24].

At the same time, dynamic recrystallization occurs due to higher plastic deformation by applying higher rotational speed and forging pressure, increasing the strain rate up to 10³ s1 [25], which leads to grain refinement in the weld interface. The width of the weld zone becomes wider by increasing rotational speed due to higher heat input, which was observed from five macrographs. The weld zone increased at a rate of 0.5 mm, which is from 1.5 to 3 mm. This also proves that higher deformation occurs between base metals by increasing rotational speed.

3.5 Microstructural analysis

To study the effect of rotational speed, microstructures were taken from four different zones like FDZ, PDZ, HAZ, and BM for different rotational speeds of S1 (800 rpm), S3 (1000 rpm), and S5 (1200 rpm), respectively. These are illustrated in Figure 11. Figures 11a, b shows the weld interface microstructure of the S5 welded specimen. Figure 11a reveals the presence of a fully deformed zone and high deformation of grains toward the weld interface from the base metal of the rod. Bainite microstructure of plate-like structure is observed as shown in Figure 11b, which consists of rich ferrite and cementite. It is mainly due to microstructural heterogeneity and pseudo-grains formation according to the cooling rate in weld interference. This heterogeneity of grains is due to chemical gradients and temperature gradients that occur in cooling [26].

Figure 11: Microstructures of welded specimens, a) S5 – weld region, b) S5 – fully deformed zone, c) S5 – heat affected zone – AISI 1020 (rod), d) S5 – heat affected zone – AISI 1018 (plate), e) S3 – heat affected zone – AISI 1020 (rod), f) S3 – heat affected zone – AISI 1018 (plate), g) S1 – weld defect.

The HAZ of the rod specimens of S5 and S3 welded joints is shown in Figure 11c, e. From the figures, an elongated ferrite structure is observed in the heat flow direction. The grains near the fusion zone are oriented along with large heat flow directions [27]. Hence, the HAZ of the rod specimen of S5 (Figure 11c) has elongated finer grains compared to the HAZ of the S3 rod specimen (Figure 11e). This is mainly because of higher heat flow due to an increase in rotational speed and also the distribution of pressure over grains. Similarly, HAZ of the plate specimens of S5 and S3 welded joints are illustrated in Figure 11d, f. Figure 11 d shows a fully refined finer grain structure in the S5 joint, whereas in S3 partially refined coarser grain structure is observed as shown in Figure 11f. This is due to the increase of temperature at HAZ by increasing rotational speed, leading to a fully refined finer grain structure.

When comparing the S5 HAZ of both rod and plate (Figure 11c, d), it shows that deformation occurs highly on the HAZ of the rod, having finer elongated grains when compared to the plate. Therefore, the rod gets higher deformation than the plate in the asymmetrical joints. Whereas the S1 specimen shows the defect in the macrograph (Figure 10a) and also in the microstructure as shown in Figure 11g. From the figure, it is observed that grain structure varies compared to other weld specimens due to lower friction and insufficient heat flow at the weld interface leads to lower material deformation during the welding process.

4 Conclusions

This study investigated the joint efficiency of the asymmetrical components of AISI 1018 steel plate and AISI 1020 steel rod during rotary friction welding. Further effects of rotational speed on macrostructure, microstructure, tensile properties, and microhardness are investigated and the following conclusions are made:

The tensile strength of friction-welded asymmetrical steel joints increases by increasing the rotational speed, whereas 1200 rpm rotational speed gives higher tensile strength of about 452 MPa and its joint efficiency of about 94%.

  1. S1 specimen (welded at 800 rpm) shows incomplete penetration of material as a weld defect due to insufficient heat flow at the weld interface because of lower rotational speed. Hence, lower rotational speed gives poor joint efficiency.
  2. Fractography analysis from fractured tensile specimens reveals that the ductile mode of failure occurs, which was confirmed by the presence of elongated dimples and higher voids over the fractured surface area.
  3. The microhardness of the welded joint S5 (welded at 1200 rpm) shows peak hardness in the weld interface at about 252 HV. The PDZ of a rod has a higher hardness value when compared to the other zones on the plate because higher deformation occurs on the rod side of the joint.
  4. From the macrostructure analysis of the joints, it was found that the rotational speed is the most influential parameter on the flash formation of the asymmetrical steel joints. Hence, the S5 specimen (welded at 1200 rpm) gets higher flash and superior joint efficiency.
  5. At high rotational speed, higher deformation is observed in the rod side of the welded specimen, which is proved by the microstructure of the HAZ of the rod getting finer elongated ferrite structure when compared to the HAZ of steel plate.

Acknowledgment: The first author express his gratitude to Centre for Materials Joining and Research (CEMAJOR), Department of Manufacturing Engineering, Annamalai University Annamalai Nagar, India, for their technical assistance.

Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

Research funding: None declared.

Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

  1. M. Sahin,Joining with friction welding of high-speed steel and medium-carbon steel, J. Mater. Proc. Technol., vol. 168, no. 2, pp. 202210, 2005, https://doi.org/10.1016/j.jmatprotec.2004.11.015.
  2. G. Subhash Chander, G. Madhusudhan Reddy, and A. Venugopal Rao,Influence of rotational speed on microstructure and mechanical properties of dissimilar metal AISI 304-AISI 4140 continuous drive friction welds, J. Iron Steel Res. Int., vol. 9, no. 10, pp. 6473, 2012, https://doi.org/10. 1016/S1006-706X(12)60154-X.
  3. D. Ananthapadmanaban, V. Seshagiri Rao, N. Abraham, and Prasad Rao,A study of mechanical properties of friction welded mild steel to stainless steel joints, Mater. Des., vol. 30, no. 7, pp. 26422646, 2009, https://doi.org/10.1016/j.matdes. 2008.10.030.
  4. Ho T. My Nu, N. Huu Loc, and L. P. Minh,Influence of the rotary friction welding parameters on the microhardness and joint strength of Ti6Al4V alloys, J. Eng. Manufact., vol. 235, no. 5, 795805, 2020, https://doi.org/10.1177/0954405420972549.
  5. J. Alex Anandharaj, S. Rajakumar, V. Balasubramanian, and S. Kavitha,Influence of process parameters on hot tensile behavior of rotary friction welded Inconel 718/AISI 410 dissimilar joints, CIRP J. Manufact. Sci. Technol., vol. 35, pp. 830838, 2021, https://doi.org/10.1016/j.cirpj.2021.09.010.
  6. R. Winiczenko,Effect of friction welding parameters on the tensile strength and microstructural properties of dissimilar AISI 1020-ASTM A536 joints, Int. J. Adv. Manufact. Technol., vol. 84, pp. 941955, 2015, https://doi.org/10.1007/s00170-015-7751-5.
  7. W. Li and F. Wanga,Modeling of continuous drive friction welding of mild steel, Mater. Sci. Eng., A, vol. 528, no. 18, 59215926, 2011, https://doi.org/10.1016/j.msea.2011.04.001.
  8. R. Paventhan, P. R. Lakshminarayanan, and V. Balasubramanian, Optimization of friction welding process parameters for joining carbon steel and stainless steel, J. Iron Steel Res. Int., vol. 19, no. 1, pp. 6671, 2012, https://doi.org/10.1016/S1006-706X(12) 60049-1.
  9. S. Celik, A. Deniz Karaoglan, and I. Ersozlu,An effective approach based on response surface methodology for predicting friction welding parameters, High Temp. Mater. Process., vol. 35, no. 3, pp. 235241, 2015, https://doi.org/10.1515/htmp-2014-0201.
  10. S. Çelik and I. Ersozlu,Investigation of microstructure and mechanical properties of friction welded AISI 316 and Ck 45 Steels, High Temp. Mater. Process., vol. 33, no. 2, pp. 161170, 2014, https://doi.org/10.1515/htmp-2013-0042.
  11. R. Damodaram, G. Manogna Karthik, and S. Vardhan Lalam, Microstructure and mechanical properties of a rotary friction welded tungsten heavy alloy, Mater. Test., vol. 61, no. 3, 209212, 2019, https://doi.org/10.3139/120.111307.
  12. P. Sivaraj, P. Hariprasath, C. Rajarajan, and V. Balasubramanian, Analysis of grain refining and subsequent coarsening along on adjacent zone of friction stir welded armour grade aluminium alloy joints, Mater. Res. Exp., vol. 6, no. 6, pp. 112, https://doi. org/10.1088/2053-1591/ab0e37.
  13. P. Sivaraj, M. Seeman, D. Kanagarajan, and R. Seetharaman, Influence of welding parameter on mechanical properties and microstructural features of resistance spot welded dual phase steel sheets joint, Mater. Today Proc., vol. 22, pp. 558562, 2020, https://doi.org/10.1016/j.matpr.2019.08.201.
  14. P. Sivaraj, C. Rajarajan, M. Seeman, and V. Balasubramanian, Influence of electrode force on metallurgical studies and mechanical properties of resistance spot welded dual phase (DP800) steel joints, Mater. Today Proc., vol. 22, pp. 614618, https://doi.org/10.1016/j.matpr.2019.09.009.
  15. N. Sankar, S. Malarvizhi, and V. Balasubramanian,Mechanical properties and microstructural characteristics of rotating arc-gas metal arc welded carbon steel joints, J. Mech. Behav. Mater.,vol. 30, no. 1, pp. 49–58, 2021, https://doi.org/10.1515/jmbm-2021-0006
  16. M. Balakrishnan and V. Balasubramanian,Tensile properties of friction welded AISI 4340 joints, Mater. Test., vol. 59, no. 1,22–28, 2017, https://doi.org/10.3139/120.110958.
  17. I. Kirik, N. Özdemir, and F. Sarsilmaz, “Microstructure and mechanical behaviour of friction welded AISI 2205/AISI 1040 steel joints,” Mater. Test., vol. 54, no. 10, pp. 683–687, 2012,  https://doi.org/10.3139/120.110379 .
  18. K. Karthick, S. Malarvizhi, and V. Balasubramanian, “Mechanical properties and microstructural characteristics of rotary friction welded dissimilar joints of rolled homogeneous armor steel and medium carbon steel,” J. Mech. Behav. Mater., vol. 30, no. 1, pp. 171–178, 2021, https://doi.org/10.1515/ jmbm-2021-0017.
  19. B. Prasanna Nagasai, S. Malarvizhi, and V. Balasubramanian, “Mechanical properties of wire arc additive manufactured carbon steel cylindrical component made by cold metal transferred arc welding process,” Mater. Test., vol. 64, no. 2, 260–271, 2022, https://doi.org/10.1515/mt-2021-2051.
  20. B. Prasanna Nagasai, S. Malarvizhi, and V. Balasubramanian, “Effect of welding processes on mechanical and metallurgical characteristics of carbon steel cylindrical components made by wire arc additive manufacturing (WAAM) technique,” CIRP J. Manufact. Sci. Technol., vol. 36, pp. 100–116, 2021, https://doi.org/10.1016/j.cirpj.2021.11.005.
  21. A. Ramaswamy, S. Malarvizhi, and V. Balasubramanian, “Post-weld heat treatment effects on the tensile properties of cold metal arc welded AA 6061-T6 aluminum joints,” Mater. Test., vol. 62, no. 1, pp. 69–76, 2020, https://doi.org/10.3139/120. 111454.
  22. S. Tushar, V. Balasubramanian, and S. Malarvizhi, “Microstructural characteristics and tensile properties of gas tungsten constricted arc (GTCA) welded Inconel 718 super alloy sheets for aeroengine components,” Mater. Test., vol. 62, no. 11, pp 1099–1108, 2020, https://doi.org/10.3139/120.111576.
  23. G. Madhusudhan Reddy and P. Venkata Ramana, “Role of nickel as an interlayer in dissimilar metal friction welding of maraging steel to low alloy steel,” J. Mater. Proc. Technol., vol. 212, no. 1, pp 66–77, 2012, https://doi.org/10.1016/j.jmatprotec.2011.08.005.
  24. N. Ozdemir, “Investigation of the mechanical properties of friction-welded joints between AISI 304L and AISI 4340 steel as a function rotational speed,” Mater. Lett., vol. 59, no. 19, pp 2504–2509, 2005, https://doi.org/10.1016/j.matlet.2005.03.034.
  25. M. B. Uday, M. N. Ahmad Fauzi, H. Zuhailawati, and A. B. Ismail, “Advances in friction welding process,” Sci. Technol. Weld. Join., vol. 15, no. 7, pp. 534–558, 2010, https://doi.org/10.1179/ 136217110X12785889550064.
  26. R. Paventhan, P. R. Lakshkminarayanan, and V. Balasubramanian, “Fatigue behaviour of friction welded medium carbon steel and austenitic stainless steel dissimilar joints,” Mater. Des., vol. 32, no. 4, pp. 1888–1894, 2010, https://doi.org/10.1016/j.matdes.2010.12.011.
  27. E. Bayraktar, D. Kaplan, L. Devillers, J. P. Chevalie, “Grain growth mechanism during the welding of interstitial free (IF) steels,” J. Mater. Proc. Technol., vol. 189, nos. 1–3, pp. 114–125, 2007, https://doi.org/10.1016/j.jmatprotec.2007.01.012.

The authors of this contribution

Dhamothara Kannan Thirumalaikkannan

Dhamothara Kannan Thirumalaikkannan was born in Tamil Nadu, India on 15.11.1996, currently pursuing his Ph. D (Manufacturing Engineering) in the Department of Manufacturing Engineering, Annamalai University, Annamalai Nagar, India. He completed his B.E. (Mechanical Engineering) at E.G.S Pillay Engineering College, Anna University, Chennai, in 2017 and received his M.E. (Manufacturing Engineering) from E.G.S Pillay Engineering College, Anna University, Chennai, India, in 2019. He has presented his research articles in various national and international conferences. He has 3 years of research experience in the area of solid-state welding. His research interests include asymmetrical components manufacturing, mechanical testing, and the characterization of materials.

Sivaraj Paramasivam

Dr. Sivaraj Paramasivam born in 1976 is currently an Associate Professor at the Centre for Materials Joining & Research (CEMAJOR), Department of Manufacturing Engineering, Annamalai University, Annamalai Nagar, India. He received his Ph.D. at Annamalai University, Chidambaram, in 2015. He has 20 years of teaching experience and 15 years of research experience. He has contributed about 40 research papers in both national and international journals and conferences at the national and international levels. His areas of interest are welding technology and materials science. He has completed various sponsored R&D projects from various funding agencies such as NRB, AICTE, and DRDO (as principal investigator).

Seeman Murugesan

Dr. Seeman Murugesan was born in Tamil Nadu, India on 30-06-1975, and holds Ph.D. in Manufacturing Engineering from Annamalai University, Annamalai Nagar, India, in 2014. He received his Bachelors degree in Mechanical Engineering from Bharathidhasan University and Masters degree in Production Engineering from Annamalai University in 19972002. He has 20 years of teaching experience. He is currently working as an Associate Professor in the Department of Manufacturing Engineering, Annamalai University. He has contributed about 43 research papers in journals and conferences at the national and international levels. His areas of interest are composites, metal joining, acoustic emission signal monitoring, modeling, and optimization.

Balasubramanian Visvalingam

Dr. Balasubramanian Visvalingam, born in 1968, is currently a Professor, in the Department of Manufacturing Engineering and as Director of CEMAJOR (Centre for Materials Joining and Research) and DRD (Directorate of Research and Development) Annamalai University, Annamalai Nagar, India. He graduated from the Government College of Engineering, Salem, University of Madras in 1989, and obtained his post-graduate degree from the College of Engineering, Guindy, Anna University, Chennai, in 1992. He received his Ph.D. at the Indian Institute of Technology Madras (IITM), Chennai, in 2000. His areas of interest are materials joining, surface engineering, and nano-materials. He has completed 25 R&D projects funded by various funding agencies such as DST, DRDO, UGC, AICTE, DAE, NRB, ARDB, and Ministry of Environment & Forest.