Next Article in Journal
Effect of Processing-Induced Oxides on the Fatigue Life Variability of 6082 Al-Mg-Si Alloy Extruded Components
Previous Article in Journal
Investigation of Die Design Influence on Required Forces in Near-Solidus Forging of Complex Components
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Statistical and Machine Learning Classification Approaches to Predicting and Controlling Peak Temperatures During Friction Stir Welding (FSW) of Al-6061-T6 Alloys

1
Department of Automotive and Marine Engineering, NED University of Engineering and Technology, Karachi 75270, Pakistan
2
Department of Mechanical Engineering, NED University of Engineering and Technology, Karachi 75270, Pakistan
3
Department of Materials Engineering, NED University of Engineering and Technology, Karachi 75270, Pakistan
*
Author to whom correspondence should be addressed.
J. Manuf. Mater. Process. 2025, 9(7), 246; https://doi.org/10.3390/jmmp9070246
Submission received: 18 June 2025 / Revised: 14 July 2025 / Accepted: 17 July 2025 / Published: 21 July 2025

Abstract

This paper presents optimization of peak temperatures achieved during friction stir welding (FSW) of Al-6061-T6 alloys. This research work employed a novel approach by investigating the effect of FSW welding process parameters on peak temperatures through the implementation of finite element analysis (FEA), the Taguchi method, analysis of variance (ANOVA), and machine learning (ML) algorithms. COMSOL 6.0 Multiphysics was used to perform FEA to predict peak temperatures, incorporating seven distinctive welding parameters: tool material, pin diameter, shoulder diameter, tool rotational speed, welding speed, axial force, and coefficient of friction. The influence of these parameters was investigated using an L32 Taguchi array and analysis of variance (ANOVA), revealing that axial force and tool rotational speed were the most significant parameters affecting peak temperatures. Some simulations showed temperatures exceeding the material’s melting point, indicating the need for improved thermal control. This was achieved by using three machine learning (ML) algorithms, i.e., Logistic Regression, k-Nearest Neighbors (k-NN), and Naive Bayes. A dataset of 324 data points was prepared using a factorial design to implement these algorithms. These algorithms predicted the welding conditions where the temperature exceeded the melting temperature of Al-6061-T6. It was found that the Logistic Regression classifier demonstrated the highest performance, achieving an accuracy of 98.14% as compared to Naive Bayes and k-NN classifiers. These findings contribute to sustainable welding practices by minimizing excessive heat generation, preserving material properties, and enhancing weld quality.

1. Introduction

FSW is a solid-state joining process gaining popularity in industrial applications due to its ability to produce high-strength and defect-free welds in various materials, including aluminum alloys [1]. Al 6061-T6 is amongst the family of aluminum alloys of the 6xxx series and is widely used in aerospace, automotive, and structural applications due to its excellent mechanical properties and corrosion resistance. This makes it an ideal choice in this study to investigate thermal behavior during the FSW process. Due to the involvement of friction and heat, all physical and chemical changes develop during the thermal process in FSW operation [2,3,4]. Since the material cannot reach melting point during the welding operation, the significance of peak temperature cannot be ignored. The temperature distribution during the welding process not only influences the mechanical strength, grain size, and residual stress in the weld zone but also determines whether the welding process will be successful at all [5]. FSW typically operates within a temperature range of 200 °C to 550 °C in the welding of aluminum alloys [6]; however, the maximum temperature in the welding zone of aluminum alloys can reach up to 80–90% of the alloy’s melting temperature [7]. For welding Al-6061-T6 alloys, the critical temperature is 430 °C, beyond which precipitate coarsening occurs, leading to reduced hardness in the heat-affected zone (HAZ) [8,9]. This makes temperature a critical response parameter significantly influenced by FSW process parameters. The process parameters that affect temperature distribution and the weld zone in the FSW process include tool rotational speed, transverse speed, axial force, and tool geometry, such as the dimensions and profile of the pin and shoulder [7]. Given the significant impact of temperature on weld quality and the potential for temperatures to approach the material’s melting point, optimization of welding parameters to study their effect on welding temperature is crucial and must be investigated.
Researchers have explored the optimization of FSW parameters from various perspectives in their published studies. Sabry et al. [10] investigated the underwater friction stir welding (UWFSW) of Al-6061/SiC metal matrix composites, focusing on optimizing process parameters such as tool rotating speed, welding speed, and silicon carbide content to maximize ultimate tensile strength and microhardness. Their study utilized response surface methodology (RSM) and ANOVA to identify optimal welding parameters. Many other researchers have also used RSM to optimize the FSW welding parameters [11,12,13,14]. Additionally, Sabry [15] examined the influence of clamping parameters, precisely clamping torque and clamping pitch, on FSW quality, highlighting factors beyond the commonly studied tool and process parameters. El-Wazery et al. [16] performed ultrasonic-assisted friction stir welding (UAFSW) to study the effect of vibration amplitudes, rotational speed, and welding speed on tensile strength using Taguchi and ANOVA. Their model predicted tensile strength with an error of 3.5%. J.T. Chinna Rao et al. [17] fabricated joints of AA 6061 T6 and AA 5052 using a Taguchi L9 array. The maximum tensile strength was achieved at 20 mm/min and 1400 rpm welding parameter settings, while a maximum temperature of 349 °C was reported in their work. Nourani et al. [18] employed Taguchi’s design and FEM simulation to optimize the FSW welding parameters in the welding of aluminum alloys. Their work investigated the effect of rotational speed and axial force on peak temperatures and HAZ. In another similar study that employed a Taguchi L9 array and ANOVA, it was revealed that axial force has a greater effect on peak temperature than tool rotational speed [19]. Anandan et al. [6] employed machine learning techniques to predict the peak temperatures during FSW of Al 7075 and Al 2014 A dissimilar joints. They used several regression models to predict the peak temperature and reported a peak temperature of 413 °C at a welding speed of 30 mm/min and a rotational tool speed of 1600 rpm. These studies emphasize the importance of considering a wide range of factors and methodologies for optimizing FSW processes for different materials and conditions.
Previous studies have also demonstrated that the welding process parameters significantly influence the tensile strength, microstructure, temperature distribution, and hardness levels in different weld zones in FSW [20,21,22]. A high rotational speed and a low welding speed elevate peak temperature and enhance tensile strength, while high rotational speed in combination with high axial force increases hardness levels as reported by Sabry et al. [20]. Choudhary and Jain [23] explored the influence of welding process parameters on peak temperatures developed in the stir zone (SZ), tensile strength, and defect formation in AA1100 joints. Their research revealed that high rotational speeds of 1200 and 1500 rpm produced peak temperatures of 593 °C and 612 °C, respectively, causing softening and over-stirring of the material. This led to lowered hardness levels in the SZ and a reduction in tensile strength and weld efficiency. In another study, the effect of the shoulder diameter of the FSW tool—made up of HSS steel of 62 HRC—on temperature distribution in Al 7075-T651 plates was investigated [24]. It was revealed in their study that a larger diameter elevates peak temperature, affecting the microstructure of the welded joints.
In recent years, ML techniques have shown great promise in addressing forecasting challenges and improving the FSW process. Many researchers have used various ML algorithms to predict mechanical properties and temperature in the FSW process. Thapliyal and Mishra [25] presented an ML classification-based model for the FSW of copper in their research. They employed four ML classifiers, i.e., Artificial Neural Network (ANN), k-Nearest Neighbor (kNN), Decision Tree (DT), and Decision Tree with Gini Index (DT-GI). They concluded that ANN predicted tensile strength from the dataset with an accuracy of 94%. Kishan et. al. [26] used a comprehensive dataset of 213 sources from peer-reviewed papers to predict the tensile strength of FSW joints. Six ML classifiers were employed in their research: Decision Tree (DT), Random Forest (RF), Adaptive Boosting Classifier (ABC), k-Nearest Neighbor (kNN), Gaussian Naive Bayes (GNB), and Support Vector Machine (SVM). Adaptive boosting provided the highest accuracy of 83.7% among the classifiers. Verma et. al. [27] predicted the tensile strength of FSW joints of AA7039 using Gaussian process regression (GPR), linear regression, and an Artificial Neural Network (ANN). ANN predicted the most accurate outcome among these algorithms. Renangi and Natarajan [28] predicted the temperature in friction lap welding of AA 7475 with PPS polymer using SVM and GPR models with a PUK kernel. Through this research, we intend to contribute to FSW technology by offering valuable insight into the interaction between welding process parameters and peak temperatures using data-driven ML algorithms. Hence, we employed various ML algorithms including Logistic Regression, k-NN, and Naive Bayes to predict the peak temperatures during FSW of an Al 6061 T6 alloy.
Unlike previous studies that relied on limited datasets with a selection of welding parameters resulting in peak temperatures well below the melting temperature, this study incorporates a comprehensive set of welding parameters, including tool material, pin and shoulder diameter, rotational speed, welding speed, axial force, and coefficient of friction.
A temperature of 430 °C has been reported in the literature [8,9] as the onset of precipitate coarsening and the subsequent reduction in the mechanical properties of Al 6061-T6; this primarily underscores the significance of peak temperature control in FSW. However, the present study is focused on identifying and avoiding parameter combinations that drive the weld zone toward the melting temperature (600 °C), which represents a more critical threshold associated with material failure and joint degradation. Therefore, a threshold of 600 °C was selected for the binary classification model to highlight and predict extreme thermal conditions that are a critical risk to weld quality.
Hence, this study presents a very novel approach by integrating FEA simulations, Taguchi design, ANOVA-based regression, and ML classification to prevent excessive temperature increases and melting through optimization and predictive modeling. This research mainly focuses on thermal control to avoid melting, making the FSW process robust by keeping the temperature below the melting point of the Al 6061 T6 alloy. The novelty in the study is enhanced by validation and prediction through three independent techniques, making the findings applicable for industrial applications.

2. Materials and Methods

2.1. Modeling of FSW Plates in COMSOL

The dimensions of the Al 6061 T6 are a width of 120 mm, a length of 102 mm, and a thickness of 3 mm. Although two plates are needed to join the material using the FSW process, only one plate is sufficient to model due to the symmetry shown in Figure 1a. The material properties of Al 6061 T6 are listed in Table 1. The FSW tool’s properties are imported from the built-in library of COMSOL 6.0; however, the tool geometry is modeled in the COMSOL modeler. The problem is defined as a stationary convection–conduction problem after making the coordinate transformation in which the tool remains plunged. At the same time, the aluminum 6061T6 plates are assumed to be infinitely long. During the FSW process, the joint is fabricated by the movement of the tool along the joint; however, this technique is complex when modeling the tool as a moving heat source. The modeling approach used in this work, in which the moving coordinate system is fixed to the tool axis, is also used by P. Chansoria et al., Song et al., and Zhu and Chao [29,30,31]. This makes the model much simpler to compute the maximum temperature.
The geometry is meshed using optimum mesh characteristics. The shoulder and pin area on top of the workpiece is meshed with a free triangular mesh with a fine size in COMSOL, while the remaining region of the workpiece is meshed with a free quad mesh (Figure 1b). The maximum and minimum sizes of the elements are set to 32 mm and 4.2 mm, respectively, with a curvature factor of 0.6 and a maximum growth rate of 1.45, ensuring accurate temperature prediction with maintained computational efficiency. The mesh is then extruded through the thickness of the workpiece.

2.2. Mathematical Models

The main sources of heat generation in FSW are shoulder–workpiece interaction and tool pin–workpiece interaction. The tool shoulder–workpiece interaction is due to friction and is modeled by Song et al. [30] using the following equation.
Q s = 2 π μ F n A s h l R s h l n
Many other investigations have also reported heat generation variations at the pin–workpiece interface [30,33]. However, few have considered it negligible, as reported by Zhu and Chao [31]. The heat generation and transfer processes are included in this work to model the FSW process accurately. Colegrove [34] provides the expression for the amount of heat from the pin, as shown in Equation (2).
Q p i n = 2 π μ n 3 ( 1 + μ 2 ) Y ¯ R p i n η
The coefficient of friction is computed from the following equation reported by Arora et al. [35]. This equation is used in investigating the effect of the coefficient of friction on temperature, along with other parameters.
μ = μ 0 e δ n r n 0 R S
where
δ = 0.2 + 0.6 e δ 0 n r n 0 R S    
δ0 is a constant value ranging between 0.2 and 0.8, depending upon the sticky or sliding condition. It is 0.3 in this analysis. n0 is the reference rpm, taken as 300 rpm.

2.3. Governing Equations

P. Chansoria et al. [29] presented the following governing equations for modeling the FSW process as a transient heat transfer problem.

2.3.1. Heat Transfer Equation

The governing equation for the heat transfer problem is given as
ρ C p u · T = · k T + Q
This equation depicts the variation of temperature within Al 6061 T6 plates. The left-hand side of the equation represents the rate of change of energy in the material, while the right-hand side contains heat conduction and internal heat generation terms from friction and plastic deformation caused by the rotating tool.

2.3.2. Heat Loss from Upper Plate

The heat lost due to natural convection and surface-to-ambient radiation from the upper plate corresponds to the following heat flux equation.
e · k T = Q u p = h u p T 0 T + ε σ ( T 0 2 T 4 )
The term hup (T0T) represents the convective heat loss in ambient air, while the term εσ(T02T4) accounts for radiative heat loss based on the Stefan–Boltzmann law.

2.3.3. Effect of Backing Plate

The backing plate is used in the FSW process and is placed below the plates to be joined together to maintain and contain the weld pool. The heat transfer through conduction at the backing plate is modeled as a convective heat transfer equation. Zhu and Chao [31] proposed heat flux equation:
e · k T = Q d o w n = h d o w n T 0 T
The convective heat transfer coefficient at the bottom is higher than at the upper surface. Hence, the conduction heat transfer at the bottom is modeled as a convective heat transfer equation, assuming a fictitious convection coefficient approximately five times that of hup. Zhu and Chao [31] use a similar assumption in their research.

2.3.4. Clamping of Plate

To restrict the movement of the workpiece and to ensure the proper weld formation, the plate is clamped at specified locations, as shown in Figure 1a. Due to thermal insulation conditions at clamped locations, Equation (7) can be rewritten as
e · k T = 0    
All governing equations (Equations (1)–(8)) were implemented using the Model Builder section interface in COMSOL Multiphysics. Equations (1) and (2), representing the frictional heat generated from shoulder–workpiece and pin–workpiece interaction, were implemented in the COMSOL Components section as Variable 1 and Variable 2, respectively. The coefficient of friction, dynamically calculated through Equations (3) and (4), was defined as a material-dependent variable under FSW parameters in COMSOL (Table 2). The governing heat transfer equation (Equation (5)) was applied through the Solids with Translational Motion feature to account for conduction and internal heat generation due to plastic deformation and friction. The convective and radiative heat losses from the top surface (Equation (6)) were incorporated using the Surface-to-Ambient Radiation and Heat Flux 1 boundary conditions. The heat transfer through the backing plate (Equation (7)) was modeled via Heat Flux 2, applying a fictitious convection coefficient. The thermal insulation at clamped boundaries (Equation (8)) was set by using a boundary heat source and insulation boundary conditions. These settings ensured accurate thermal modeling across all simulation domains and boundary conditions.

2.4. Process Flow

Figure 2 illustrates the process flow designed to investigate the effect of welding process parameters on peak temperature, followed by optimization and prediction through Taguchi, ANOVA, and ML classification. The design of simulations was the most crucial phase in this study and was used to select appropriate welding process parameters for the analyses. Several FEA simulations were conducted before adopting the L32 Taguchi array to explore thermal response across a wide range of welding parameters. Hence, these parameters were chosen based on preliminary simulation trials, the operational constraints of the milling machine, and a comprehensive previously published work [6,18,22,23], which identified them as having the most significant impact on the temperature distribution during FSW. The welding parameters analyzed in COMSOL were tool material, pin and shoulder diameter, tool rotational speed, welding speed, axial force, and coefficient of friction. The L32 Taguchi array was selected to efficiently optimize these parameters with a significantly reduced number of simulation runs compared to a full factorial design, which would require 21 × 46 = 8192 runs. The levels of each process parameter are presented in Table 3. The Minitab 17 tool was used in this research to perform the Taguchi optimization and ANOVA.

2.5. Experimental Details

A conventional universal milling machine was customized with a specially designed fixture and data acquisition system to conduct FSW experiments, as shown in Figure 3. A fixture was fabricated from a mild steel backing plate with a thickness of 23 mm to ensure rigidity during the FSW process. The fixture was designed to weld plates in butt joint configuration with a plate length not exceeding 200 mm. The plates were clamped through locally manufactured clamps and bolts to maintain a consistent joint during the welding operation. The axial force was measured using two compression-type load cells of 8 kN capacity. The load cells were attached just beneath the mild steel backing plate at both ends of the plate to measure the normal force during the FSW process. Arduino Uno-based software was developed to acquire force data during the welding process. The temperature was measured through K-type thermocouples, which were placed at eight different locations on the backing plate (Figure 4a,b). To ensure accurate peak temperature measurement without disturbing the weld surface, thermocouples were inserted through precisely drilled holes from the underside of the backing plate (Figure 4c). The sensor tips were tightly fitted to keep contact with the bottom surface of the welded plates, not the backing plate itself. This configuration minimized thermal lag and gradient effects, allowing reliable comparison with FEA-simulated temperature values at corresponding locations on the plate.
Thermocouples T8 and T4 were placed near the SZ in TMAZ, T3 and T7 near TMAZ in HAZ, T1 and T6 near HAZ, and T2 and T5 outside HAZ. The temperature data was captured through an ET3916-08 multichannel data logger at a sampling rate of 10 Hz. The Al 6061 T6 plates were cut to dimensions of 120 mm × 60 mm × 3 mm using a CNC machine and were cleaned with acetone before placement on the fixture. The welding operation was performed by four FSW tools made up of H13 steel with a hardness of 52 HRC, as shown in Figure 5. The experiments were conducted at a room temperature of 30 °C. Due to the operational limitations of the milling machine—which could operate at axial loads of 1–2 kN and rotational speeds of 480, 600, 670, 900, 1200, 1320, and 1800 rpm—the experiments were conducted using one axial load and four rotational speeds. The experimental parameters were selected from the Taguchi L32 array for the H13 tool material for an axial load of 1 kN. Due to equipment limitations, the experimental setup allowed welding only at an axial force of 1 kN. Consequently, the experimental trials were conducted solely at this level and used to validate the simulation model by comparing predicted and measured peak temperatures. Despite this constraint, the full factorial simulation design incorporated a broader range of axial forces (1, 5, 8, 10 kN), which enabled robust statistical and machine learning analysis. Therefore, while axial force was experimentally constant, its significance was evaluated comprehensively through a simulation-driven ANOVA. This constraint ensured that the experimental conditions were directly applicable to the available equipment while still allowing for a comprehensive investigation within the feasible parameter space.

3. Results and Discussion

3.1. Temperature Distribution in Plates

The temperature distribution on the plate with a pin diameter of 3 mm, a shoulder diameter of 12 mm, axial force of 5 kN, a tool rotational speed of 900 rpm, transverse speed of 38 mm/min, and a coefficient of friction of 0.33 is shown in Figure 6a. The temperature distribution demonstrated a localized heat zone around the pin, extending outward along the shoulder contact area. The highest temperature attained in the weld region was 347.765 °C due to intense plastic deformation and frictional heat generation at the contact surface between the tool and the plate. Asmare et al. [36] in their experimental work reported this temperature to be 342 °C under similar assumed parameters; hence, this simulation analysis is in general agreement with previous published studies [23,36]. The top view of the temperature contour is shown in Figure 6b. It is evident from Figure 6b that the isotherms were closely packed in front of the tool axis than on the trailing side. This shows a rapid addition of heat at the weld zone and a slower heat loss rate on the trailing side.
The peak temperatures obtained from FEA simulations are listed in Table 4. A large variation in peak temperature can be observed in the variation of selected process parameters. It is important to note that amongst other welding parameters, the tool rotational speed and axial force seem to be affecting the peak temperature in the FSW process. The higher rotational speed in combination with high axial force exhibited higher peak temperatures. This pattern showed a direct relationship between peak temperature and tool rotational speed and axial force. The highest temperature of 679 °C was achieved at a pin diameter of 3 mm, a shoulder diameter of 9 mm, a tool rotational speed of 1800 rpm, and axial force of 10 kN. The lowest temperature of 174 °C was obtained at the lowest levels of tool rotational speed and axial force, i.e., 600 rpm and 1 kN. Out of 32 simulations, the peak temperatures exceeded the threshold value of 600 °C on 11 occasions, as shown in Table 4. Though it was reported in several previously published research studies that peak critical temperature leads to precipitate coarsening, when the temperature exceeds the melting temperature, precipitate coarsening becomes a secondary issue, because the precipitate distribution is disrupted in the melted zone. The development of the precipitate-coarsening process cannot be ignored in the surrounding regions of the melting zone where the temperature reaches a critical value but remains below the melting point. It is crucial to note that this study focuses exclusively on predicting peak temperatures and identifying critical welding conditions during the FSW process of Al 6061 T6; thus, the effects of remelting or process resumptions are beyond the scope of this research. It is suggested that future research should address remelting strategies, including investigation of re-engagement speed and dwell time, thereby ensuring consistent weld quality.
The comparison of peak temperatures acquired from FEA simulations and experiments is presented in Table 5. An error of less than 7% demonstrated close agreement between FEA and experiments, thus validating the computational model’s accuracy. Eight thermocouples (T1–T8) were placed on the Al 6061 T6 plate to measure temperatures on advancing and retreating sides; however, only four temperatures (T5–T8) were computed from FEA by placing a coordinate system in COMSOL. A comparison of COMSOL-analyzed and experimentally measured temperatures is presented in Figure 7a–d.
A comparative evaluation of peak temperature profiles at eight thermocouple locations for four distinctive experimental conditions is presented in Figure 8a–d. These figures reflect the relationship between tool geometry, tool rotational speed, and welding speed governing the heat generation, heat distribution, and heat dissipation dynamics during the FSW process. Importantly, the peak temperature is raised from 195 °C to 314 °C when the rotational speed is increased from 600 rpm (Exp # 1) to 1800 rpm (Exp # 4). This is due to high friction between the tool and the material caused by increased tool rotational speed, resulting in greater heat generation. Additionally, the time taken to reach the peak temperature is also reduced from 85 s (Exp # 1) to 55 s (Exp # 4), showing the generation of heat at higher speeds. This may improve material plasticization; however, it could also increase the risk of overheating and the grain growth process. Welding speed is another important parameter in understanding thermal behavior in FSW. Although the welding speed is nearly doubled from 20 to 38 mm/min in Exp # 3 as compared to Exp # 2, the increase in peak temperature is still observed from 239 °C to 279 °C. This reveals that while higher welding speeds tend to reduce heat input due to a shorter interaction time, the effect of increasing rotational speed on the temperature increase is more dominant than that of welding speed, likely due to a nonlinear relationship with frictional energy. Exp # 4 presents the combined effect of the highest levels of both speeds, resulting in the highest peak temperature of 314 °C. However, the reduced dwell time at the peak indicates a shorter thermal exposure at higher heating, leading to a condition that could favor a fine-grained microstructure with rapid cooling. The pin and shoulder diameters also influence the thermal behavior during welding by affecting the contact area and frictional heat generation. A larger pin diameter contributes to internal heat generation and appropriate material mixing by increasing the stirring volume beneath the surface; similarly, a larger shoulder diameter enhances frictional heat generation by increasing the contact area between the tool and the workpiece. This can be observed in Exp # 4, where the largest pin diameter of 6 mm and shoulder diameter of 15 mm in combination with high rotational and welding speed contribute to the highest heat input and rapid thermal response.
It is worth emphasizing that the temperature difference between the advancing and retreating sides (T8 and T4) in TMAZ revealed asymmetry in heat generation and material flow. At other locations, the temperature remained higher on the advancing side as compared with the retreating side, with minimal temperature difference. These patterns could be observed in all four temperature distribution plots. Keeping the asymmetric temperature distribution across advancing and retreating sides in mind, future studies should incorporate microstructural characterization to investigate material flow behavior and defects formation in weld zones.

3.2. Taguchi Optimization

Taguchi’s optimization and analysis of variance (ANOVA) were used to optimize the process parameters to achieve the nominal temperature in the FSW welding process at a confidence level of 95%. This study aimed to optimize the FSW parameters to control the temperature, preventing it from exceeding the melting temperature, i.e., 600 °C, which is the critical threshold at which melting could occur in Al-6061-T6 alloys. The material’s structural integrity can be maintained further and poor-quality joints prevented by avoiding temperatures near the threshold. Hence, the “smaller is better” criterion was selected in the Taguchi method to minimize the maximum temperature. This approach was preferred over the “nominal is better” approach because it provides parameter combinations that yield lower maximum temperatures, ensuring control of overheating of the workpiece during FSW operation. This technique identifies the optimal settings that reduce peak temperatures, optimizing the process to remain within a safe range. The “smaller is better” criterion was utilized as it identifies settings that achieve low average values and maintain those values consistently. This is crucial for FSW processes to achieve a target value and minimize variability.
The rank and delta are two important indicators in Taguchi analysis. Rank represents the impact of each welding parameter on the response variable, while the delta value shows the difference between the highest and lowest S/N ratios calculated by the Taguchi method. The rank 1 and highest delta values signify each parameter’s sensitivity in the Taguchi method.
The response table was calculated from Equation (9) for signal-to-noise (S/N ratio), considering the “smaller is better” criterion, and it is tabulated in Table 6 (while means are listed in Table 7). The higher negative values in the S/N table represent more stable and lower heat generation during FSW. Based on S/N ratios, mean values, rank, and delta values, it is evident from Table 6 and Table 7 that the axial force and tool rotational speeds are the most influential process parameters affecting peak temperature in FSW operations.
S N = 10 log 10 Y 2 n
where Y = the responses for the given factor level combination and n = the number of responses in the factor level combination.

3.3. Implementation of the ANOVA

An ANOVA was employed along with the Taguchi approach to validate and compare the ranking of FSW process parameters. The F-values and p-values are tabulated in Table 8. The p-values (<0.005) indicated that the axial force and the tool rotational speed significantly affected the FSW process. The higher F-values further revealed that axial force had the greatest effect on the process, followed by tool rotational speed, which is consistent with Taguchi’s results. The regression of the model is shown in Equation (10), and observation, normal probability, interaction, and main effect plots are shown in Figure 9a–c. Through residual analysis, the adequacy of the suggested regression models was tested. The observation order plots did not suggest any significant inadequacies in the model. The normal probability plot of residuals indicated that the residuals were approximately normally distributed. This suggested that the model adequately captured the variability in the maximum temperature data and supported the validity of the statistical analysis. A main effects plot revealed the influence of welding parameters on the maximum temperature. It was observed that the H13 resulted in a slightly lower maximum temperature than the M2 tool; however, the difference was minimal, showing the negligible impact of tool materials on maximum temperature. Pin diameter had a more moderate influence, with a gradual increase in maximum temperature as the diameter increased from 3 mm to 6 mm, likely due to a larger contact area and increased material deformation. Notably, tool rotational speed and welding force exhibited a clear and substantial impact on peak temperatures. The plot showed a strong positive correlation between rotational speed and maximum temperature, with temperatures rising dramatically as rotational speed increased from 600 rpm to 1800 rpm. Similarly, welding force demonstrated a pronounced effect, with maximum temperature increasing sharply as force increased from 1 kN to 10 kN. This relationship was attributed to the greater frictional resistance and plastic deformation occurring at higher forces, both of which contributed to heat generation.
The regression equation is as follows:
Max. Temperature (°C) = 483.63 + 2.96 Tool Material_H13 − 2.96 Tool Material_M2
−25.3 Pin Dia (mm)_3 − 20.4 Pin Dia (mm)_4 + 6.2 Pin Dia (mm)_5
+39.6 Pin Dia (mm)_6 − 0.6 Shoulder Dia (mm)_9
−33.5 Shoulder Dia (mm)_12 + 12.5 Shoulder Dia (mm)_15
+21.6 Shoulder Dia (mm)_18 − 118.1 Tool Rotational Speed (rpm)_600
−20.4 Tool Rotational Speed (rpm)_900
+51.6 Tool Rotational Speed (rpm)_1200
+86.9 Tool Rotational Speed (rpm)_1800
+55.3 Welding Speed (mm/min)_20 − 30.2 Welding Speed (mm/min)_38
−15.2 Welding Speed (mm/min)_50 − 9.9 Welding Speed (mm/min)_80
−222.7 Axial Force (kN)_1 − 15.6 Axial Force (kN)_5
+99.1 Axial Force (kN)_8 + 139.3 Axial Force (kN)_10
−11.0 Coefficient of Friction_0.30
−21.2 Coefficient of Friction_0.33
+22.5 Coefficient of Friction_0.36
+9.6 Coefficient of Friction_0.40

3.4. ML Classification Approaches

The selection of input variables to implement ML classifiers was carried out through ANOVA analysis. Table 8 reveals that all parameters investigated in the study were significant except for tool material, shoulder diameter, and coefficient of friction, with a p-value greater than 0.05. Hence, for the implementation of a machine learning-based classification model, these parameters were not considered. Although shoulder diameter was found to be statistically insignificant in the ANOVA analysis, it was included in the COMSOL simulations for geometric consistency. Specifically, the tools were designed and fabricated based on a pin-to-shoulder diameter ratio of 1:3, which necessitated defining the shoulder diameter for each corresponding pin diameter in the simulations. Therefore, shoulder diameter was indirectly represented in the thermal simulations, but it was not treated as a separate variable in the machine learning classification due to its statistical insignificance.
Considering the large dataset requirement of the ML models, a new dataset of 324 datapoints was prepared through factorial design. A factorial design (34 × 41) containing 4 parameters of 3 levels and 1 parameter of 4 levels resulted in a simulation set of size 324 for performing FEA in COMSOL Multiphysics; a total of 324 independent simulations were carried out to achieve transient temperatures. The levels of the parameters are presented in Table 9, and the peak temperatures from the COMSOL simulations are provided in Appendix A (Table A1).
The dataset was projected through three ML classification models, i.e., Logistic Regression, k-Nearest Neighbor, and Naive Bayes. In solid-state welding, the material’s temperature remains below its melting point. For Al 6061-T6, a melting temperature of 600 °C was used to categorize the dataset. Kishan et. al. [26], Mishra and Morisetty [37], and Thapliyal and Mishra [25] adopted a similar approach in their research, classifying tensile strength based on a joint efficiency criterion of more than 70%. However, this study employed the peak temperature criterion, ensuring that it did not exceed the material’s melting point, as the basis for categorization. The response variable was binary, indicating whether the temperature exceeded 600 °C. The classification assigned a value of 0 to outputs where the temperature fell below 600 °C and 1 to those where the temperature exceeded 600 °C. Logistic Regression is an ML classification statistical method that uses a sigmoid function to map input features to a probability between 0 and 1, making it ideal for determining whether the temperature is above or below a threshold value.
In this study, the Logistic Regression was employed to classify the temperature as either class 1, where temperature > 600 °C, or class 0, where temperature ≤ 600 °C. k-NN is a distance-based, instance-based learning algorithm. The “k” in k-NN indicates the number of neighbors used to classify a new data point. In the k-NN algorithm, the k closest points from the training set are considered for each data point in the testing set. The algorithm assigns the majority class among these neighbors to the testing point. In this analysis, k = 1 yielded optimal results. Naive Bayes is a probabilistic classifier based on Bayes’ theorem. It assumes conditional independence among features, enabling simplified computation by evaluating feature likelihoods independently. It classifies data as 0 or 1 by calculating the posterior probabilities for each class using Bayes’ theorem and assigning the class with higher probability.

3.5. Implementation of ML Classification Models

After the dataset preparation, ML models were employed to train and test the dataset. These models were trained using the K-fold cross-validation technique to assess their robustness across different subsets of data. Cross-validation is a technique extensively used in ML algorithms to avoid overfitting, especially in scenarios with small datasets. This method divides the fixed portion of a dataset for validation and the remaining portion for training a model. The process is reiterated after testing, with a different portion of data allocated as the validating set. The “K” in cross-validation is the split count for the validation and training datasets. A 10-fold cross-validation was implemented in this work, where the model was trained on 90% of the data in each fold and validated on the remaining 10%; this process was repeated ten times. The ML classification models were compared through five model evaluation metrics: accuracy, F1 score, precision, recall, and error. These metrics were calculated using Equations (11)–(15) and are presented in Table 10. TP indicates true positive, TN is true negative, FP is false positive, and FN represents false negative. The highest accuracy was achieved with the Logistic Regression ML classifier, indicating that 98.14% of all predictions were correct, suggesting higher overall performance in binary classification. The F1 score balances precision and recall. A higher F1 score suggests that the ML algorithm is reasonably good at predicting temperatures > 600 °C without too many false negatives or positives. The accuracy of positive predictions is reflected by precision, whereas recall measures how well the model identifies all true positives. Table 10 depicts that the Logistic Regression ML classifier showed the best performance among all evaluated ML classifiers, achieving the highest accuracy, F1 score, recall, and precision. Further, its accuracy of 98.14% is also the highest reported in the published literature. In comparison, Thapliyal and Mishra [25] reported an accuracy of 94% using an Artificial Neural Network (ANN), while Kishan et. al. [26] achieved an accuracy of 81.6% utilizing the Adaptive Boosting classifier. The confusion matrices for all three ML classifiers are shown in Figure 10.
Accuracy = (TP + TN)/N
F1 Score = [(Precision × Recall)/(Precision + Recall)] × 2
Precision = TP/(TP + FP)
Recall = TP/(TP + FN)
Error = (FP + FN)/N
To further evaluate the performance of the classification models, ROC curves (Figure 11) were generated, and the corresponding AUC values were calculated. The AUC values obtained were 0.99757 for Logistic Regression, 0.98677 for k-NN, and 0.98319 for Naive Bayes. These values indicate a high degree of separability and excellent classification capability for all models. Logistic Regression demonstrated the best overall performance, not only achieving the highest classification accuracy of 98.14% but also the highest AUC, confirming its superior ability to distinguish between classes. Although k-NN and Naive Bayes yielded slightly lower AUCs, they still maintained strong classification potential, with k-NN exhibiting good precision and Naive Bayes offering effective probabilistic separation. The ROC curves complement traditional performance metrics and validate the robustness of the classifiers in predicting thermal thresholds in FSW simulations.

4. Industrial Applications

The findings of this study have significant potential for industries that depend on advanced manufacturing processes, particularly additive manufacturing and FSW operations. In the aerospace industry, where highly efficient materials and defect-free joints are critical, the ML modeling approach can facilitate reliable and efficient processes. By accurately forecasting temperature profiles and optimizing welding parameters, manufacturers can reduce defects, improve mechanical properties, and improve the quality of aerospace components.
Furthermore, this research can be applied to other high-precision manufacturing industries, such as automotive and defense, in which strong lightweight materials are in high demand. The incorporation of machine learning models into manufacturing workflows results in real-time process control, thus minimizing experimental costs and material wastage. As industries shift towards Industry 4.0 and smart manufacturing, the adoption of artificially intelligent predictive models like the one proposed in this research has tremendous potential for application in process automation, quality assurance, and performance optimization of various engineering applications.

5. Conclusions

This paper presents the application of the Taguchi method, ANOVA, and ML classification approach to control and predict peak temperatures in FSW of Al-6061-T6 aluminum alloys. The “smaller is better” criterion was employed in Taguchi analysis to keep the temperatures below the melting point, preventing overheating during FSW operation. An ANOVA was used to rank the process parameters, and it was found that axial force and tool rotational speed are the most significant parameters affecting peak temperatures. However, the coefficient was found to have an insignificant effect on temperature, as revealed by Taguchi and the ANOVA. Taguchi and ANOVA ranks were in good agreement, providing valuable insights into the parameter hierarchy, enabling targeted adjustments to improve weld quality effectively. The dataset of 324 datapoints was subsequently classified using three ML algorithms: Logistic Regression, k-NN, and Naive Bayes. Based on ML performance metrics (i.e., accuracy, F1 score, precision, recall, and error), it was found that Logistic Regression demonstrated the best performance, achieving an accuracy of 98.14%, which is among the highest in the FSW domain.
The findings herein are significant for sustainable manufacturing, as precise predictions of peak temperatures in FSW can minimize energy consumption and material waste while optimizing process parameters. This approach contributes to reducing experimental costs and carbon footprints associated with traditional trial-and-error methods in manufacturing. By leveraging advanced computational techniques, this research aligns with global efforts to promote sustainable manufacturing practices through intelligent process optimization.
This research work has primarily focused on optimizing and predicting peak temperature during the FSW process using FEA simulations, Taguchi, an ANOVA, and an ML classification approach. Experimental validations were performed under milling machine constraints, where the axial force was kept constant. The effect of pin profiles and shoulder shapes, high dynamic axial force, and remelting behavior was beyond the scope of this study. Future studies should incorporate tool geometrical features, varied axial force, microstructural analysis, defect characterization, and remelting procedure to establish a strong correlation between thermal behavior and joint quality. Additionally, future studies could expand this comparative approach to other alloys, adding experimental validations and exploring additional machine learning algorithms to enhance the model’s robustness for broader industrial applications, particularly in sustainable manufacturing.

Author Contributions

A.A. (the corresponding author) conceived the research idea; developed the methodology; performed simulations, data analysis, and machine learning implementation; interpreted the results; and wrote the original draft of the manuscript. M.S. provided supervision, guided the research direction, reviewed the manuscript critically for technical and intellectual content, and contributed to the final editing and refinement of the paper. M.S.H. contributed to the literature review, assisted in structuring the discussion section, and helped with the initial data preprocessing and referencing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology (MoST), Pakistan, grant number “Acad/50(48)/4398”. The APC was self-funded by the authors.

Data Availability Statement

All data generated or analyzed during this study are included in this published article.

Acknowledgments

The authors gratefully acknowledge the support of the Advanced Studies and Research Board (ASRB) of NED University of Engineering and Technology and Ministry of Science and Technology (MoST) for facilitating this research work. The authors also extend sincere thanks to the technical staff of the Industrial and Manufacturing Engineering Department Laboratories and the Materials Engineering Laboratories of NED University for their valuable assistance during experimental setup, sample preparation, and testing procedures. Their expertise and support were instrumental in the successful execution of this study.

Conflicts of Interest

The authors declare no conflicts of interest regarding the publication of this paper.

Abbreviations

The following abbreviations are used in this manuscript:
AUCArea under the curve
FEAFinite element analysis
FSWFriction stir welding
HAZHeat-affected zone
MLMachine learning
ROCReceiver operating characteristic
SZStir zone
TMAZThermomechanically affected zone

Appendix A

Table A1. Dataset Obtained from COMSOL 6.0 FEM Simulations.
Table A1. Dataset Obtained from COMSOL 6.0 FEM Simulations.
S. NoPin Dia Shoulder Dia Tool Rotational Speed Welding Speed Axial ForceTransient Temperature
15151320451292.818
2515950451264.1
3412950855386.242
4415670205394.013
54121320451272.048
6418950203378.807
74121320453392.547
85181320855539.81
9418950458659.231
10518670855325.289
11415950203365.481
125121320205659.375
13615950458659.585
14618670855342.688
15415670858413.094
165121320203479.977
17515950858585.875
186151320455658.771
196181320205660.333
20618950453369.446
21412670203296.422
225151320858659.744
235151320201316.519
24515670451243.544
255121320851281.011
266181320853401.116
274181320201293.268
284181320203492.424
29612670858466.992
30412950203356.779
31518950203399.155
32418670451223.778
336121320855574.604
34512670851233.754
35615670203341.087
36612950201303.703
37618670208658.742
38618950853334.315
39612670201270.615
40515670453291.801
41618670858458.704
424151320201289.975
435151320205659.488
44418950451247.73
455121320453424.1
46618950851266.384
47518950858600.957
484181320451269.7
49612670455386.693
50615670451257.594
51415670208576.149
52615670855346.576
534121320853357.716
54415950451247.879
55612950453322.545
564181320858659.736
576151320203463.688
58515950458659.199
59612670458522.931
60618670203344.028
61415670853253.939
62412670453273.461
63518950855411.647
644121320851260.299
65412950201261.205
66618950855432.147
674151320455592.014
686151320853405.928
694121320208660.214
705181320201317.835
71612950853347.379
725151320455627.364
73615950205593.034
74615950208659.976
75612950458659.632
765151320208661.123
77418950851235.239
78515950851254.247
79512950453347.59
804151320208660.931
81512670855333.912
82615670205438.032
83418670203304.853
84512950858583.25
85515950453346.186
865121320455607.639
874151320205659.257
886121320455655.283
896121320851301.315
904181320455624.77
916121320858660.282
92515950203388.187
93512950201280.751
94515670858434.039
95612670203341.785
965121320208660.509
97618670458545.751
98612670853298.136
99615670458527.343
100412670851212.511
1015181320208661.346
1026121320201341.735
103512950855413.368
1045181320458666.28
105512950208659.68
106612950851272.391
107412670853256.74
1084121320858659.498
109612670855357.356
110515670853272.636
1115181320858659.813
112615950858618.321
113412670458460.784
1146181320858660.373
115612950451284.199
116512950205529.689
117612670451259.458
118618670455386.008
119415670855307.758
120418670201239.002
121618950205620.185
122415670201238.341
1235121320458660.81
124418950858578.271
125615670455383.145
126518670205432.07
127512950458658.901
128518670853270.116
129415950201260.672
130515950855418.712
1316181320208661.485
1326181320201340.176
133512950853324.037
1346121320208673.359
135512670451244.762
136412950851237.408
137512670208572.431
138615950201302.608
1395181320851273.705
140415670455341.266
1415121320853386.198
1425151320855530.211
1436151320851297.056
144612950203412.583
145415950855386.272
146415670458478.643
147512670201256.415
148515670205414.613
149515670201255.507
150512670453295.295
151615950203414.508
152612950858625.06
153612670453315.11
154612670205431.979
155615670851248.203
1565181320453432.244
157518950205588.731
1584181320853357.402
159618950203423.268
160418670855309.079
1616181320451315.105
162618950451280.207
163615670208635.289
164618950201303.04
165512670458489.421
166412950455429.714
167515670203319.667
168518950853314.025
1696151320201339.517
1705151320853379.556
1716151320208661.913
172518670208639.006
173412950458630.716
174512670205404.608
175515670208603.143
1764151320853354.456
177515950455465.658
1785181320205659.634
179412950853300.504
180518950201280.419
181418670853252.997
1825121320855531.863
1834121320458660.301
184412670855310.599
1856181320203562.547
1865181320853378.661
187612950855445.463
1884151320451270.199
189418950855393.108
190518670453291.351
1916151320451316.14
192415950458658.301
1936181320458661.268
194618670851247.05
1954181320205659.513
196512670455361.008
1974121320203444.476
198518670201255.399
199512950851256.482
200615670453310.019
2015121320451294.821
202615950455493.584
2034121320455565.355
204615670858457.317
205415950858558.211
206418950201261.028
207518950458659.412
2084151320855501.443
209615670853290.748
210512950203383.803
2116181320855569.763
212515670855327.086
213612950208659.952
214612670208612.025
215415950455441.186
216618670205452.63
217418670858423.258
218415950205526.294
219618950858628.289
2205151320203497.149
221518670455368.077
2224181320458660.429
2234181320208668.412
224512950455459.279
225415950208659.671
226412670201238.335
227418950853294.206
228412670858409.463
2296151320453454.659
230615950853338.244
231415950851235.9
232515950208660.217
233418670458503.201
2346181320455659.168
235418950205562.5
236415670203298.397
237518950451263.197
2386151320858660.004
239618670201268.63
240412950208659.44
241418670455351.118
242418670208616.466
243412950858546.594
2446151320855564.397
2455121320858659.762
246618670451256.686
2474181320851256.972
2486151320458661.012
249618670453309.203
250518950455480.662
251515950201279.666
252615950453369.521
253518950208663.023
2544121320205658.853
2556151320205659.78
2564151320851258.077
2575181320203523.272
2584121320855495.184
259612950455493.49
2606181320851294.85
261518670851231.443
262512950451266.37
263512670858436.211
264418950208659.801
265615670201269.212
266415950453321.854
267518950453348.319
268518670858439.735
269615950855433.973
270418670851210.794
271512670853277.828
272515950205556.827
2734151320458659.966
274412670205379.14
2754181320453409.336
276418670205414.603
277518670451243.11
278412950453322.545
2794151320453397.964
2805151320458661.692
281412670451224.087
282618950455506.02
283412950205495.285
284415670453272.24
285415670851211.323
2866121320853418.386
2875181320451292.745
2885181320455654.997
2894151320858659.62
290618950458659.562
291515670458500.942
292518950851253.043
2936121320451319.199
294418950453327.388
295418950455460.245
2965151320851275.743
2974151320203463.688
2984181320855515.518
299418670453273.386
3006121320203523.121
3016121320205660.75
302415670451223.613
3035151320453423.877
3046121320458660.261
3055121320201316.857
3066181320453459.276
307412670455337.992
308515950853316.986
309515670851232.214
3104121320201289.45
311518670203324.975
312412670208538.335
313612950205573.128
314512670203317.596
3156121320453458.121
316618950208659.961
317615950451281.306
318412950451248.896
319612670851250.301
320618670853286.851
321415950853295.739
322615950851268.773
323518670458522.429
324515670455361.759

References

  1. Thomas, W.M.; Nicholas, E.D.; Needham, J.C.; Murch, M.G.; Temple-Smith, P.; Dawes, C.J. Friction Stir Butt Welding, International Patent Application PCT/GB92/02203, GB Patent Application 9125978.8, US Patent 5.460.317. Available online: https://www.mendeley.com/catalogue/d39ca923-225d-3e7c-8bc9-7ccca2e23b99/?utm_source=desktop&utm_medium=1.19.8&utm_campaign=open_catalog&userDocumentId=%7B01c2fd14-2fc6-391f-8c00-6ed88f2dca45%7D (accessed on 17 March 2025).
  2. Kumar, S.S.; Murugan, N.; Ramachandran, K.K. Effect of tool tilt angle on weld joint properties of friction stir welded AISI 316L stainless steel sheets. Measurement 2020, 150, 107083. [Google Scholar] [CrossRef]
  3. Zhang, S.; Shi, Q.; Liu, Q.; Xie, R.; Zhang, G.; Chen, G. Effects of tool tilt angle on the in-process heat transfer and mass transfer during friction stir welding. Int. J. Heat Mass Transf. 2018, 125, 32–42. [Google Scholar] [CrossRef]
  4. Shankar, S.; Vilaça, P.; Dash, P.; Chattopadhyaya, S.; Hloch, S. Joint strength evaluation of friction stir welded Al-Cu dissimilar alloys. Measurement 2019, 146, 892–902. [Google Scholar] [CrossRef]
  5. Yau, Y.H.; Hussain, A.; Lalwani, R.K.; Chan, H.K.; Hakimi, N. Temperature distribution study during the friction stir welding process of Al2024-T3 aluminum alloy. Int. J. Miner. Metall. Mater. 2013, 20, 779–787. [Google Scholar] [CrossRef]
  6. Anandan, B.; Manikandan, M. Machine learning approach for predicting the peak temperature of dissimilar AA7050-AA2014A friction stir welding butt joint using various regression models. Mater. Lett. 2022, 325, 132879. [Google Scholar] [CrossRef]
  7. Chalurkar, C.; Shukla, D.K. Temperature Analysis of Friction Stir Welding (AA6061-T6) with Coupled Eulerian-Lagrangian Approach. IOP Conf. Ser. Mater. Sci. Eng. 2022, 1248, 012035. [Google Scholar] [CrossRef]
  8. Woo, W.; Choo, H.; Prime, M.B.; Feng, Z.; Clausen, B. Microstructure, texture and residual stress in a friction-stir-processed AZ31B magnesium alloy. Acta Mater. 2008, 56, 1701–1711. [Google Scholar] [CrossRef]
  9. Woo, W.; Choo, H.; Withers, P.J.; Feng, Z. Prediction of hardness minimum locations during natural aging in an aluminum alloy 6061-T6 friction stir weld. J. Mater. Sci. 2009, 44, 6302–6309. [Google Scholar] [CrossRef]
  10. Sabry, I.; Hewidy, A.M. Underwater friction-stir welding of a stir-cast AA6061-SiC metal matrix composite: Optimization of the process parameters, microstructural characterization, and mechanical properties. Mater. Sci. Pol. 2022, 40, 101–115. [Google Scholar] [CrossRef]
  11. Rajakumar, S.; Muralidharan, C.; Balasubramanian, V. Establishing empirical relationships to predict grain size and tensile strength of friction stir welded AA 6061-T6 aluminium alloy joints. Trans. Nonferrous Met. Soc. China 2010, 20, 1863–1872. [Google Scholar] [CrossRef]
  12. Heidarzadeh, A.; Khodaverdizadeh, H.; Mahmoudi, A.; Nazari, E. Tensile behavior of friction stir welded AA 6061-T4 aluminum alloy joints. Mater. Des. 2012, 37, 166–173. [Google Scholar] [CrossRef]
  13. Heidarzadeh, A.; Saeid, T. Prediction of mechanical properties in friction stir welds of pure copper. Mater. Des. 2013, 52, 1077–1087. [Google Scholar] [CrossRef]
  14. Elangovan, K.; Balasubramanian, V.; Babu, S. Predicting tensile strength of friction stir welded AA6061 aluminium alloy joints by a mathematical model. Mater. Des. 2009, 30, 188–193. [Google Scholar] [CrossRef]
  15. Sabry, I. Experimental and statistical analysis of possibility sources—Rotation speed, clamping torque and clamping pith for quality assessment in friction stir welding. Manag. Prod. Eng. Rev. 2021, 12, 84–96. [Google Scholar] [CrossRef]
  16. El-Wazery, M.S.; Mabrouk, O.M.; El-Sissy, A.R. Optimization of Ultrasonic-assisted Friction Stir Welded using Taguchi Approach. Int. J. Eng. 2022, 35, 213–219. [Google Scholar] [CrossRef]
  17. Rao, J.T.C.; Harikiran, V.; Gurudatta, K.S.S.; Raju, M.V.D.K. Temperature and strain distribution during friction stir welding of AA6061 and AA5052 aluminum alloy using deform 3D. Mater. Today Proc. 2022, 59, 576–582. [Google Scholar] [CrossRef]
  18. Nourani, M.; Milani, A.S.; Yannacopoulos, S.; Nourani, M.; Milani, A.S.; Yannacopoulos, S. Taguchi Optimization of Process Parameters in Friction Stir Welding of 6061 Aluminum Alloy: A Review and Case Study. Engineering 2011, 3, 144–155. [Google Scholar] [CrossRef]
  19. Dadi, S.S.O.; Patel, C.; Naidu, B.A. Effect of friction-stir welding parameters on the welding temperature. Mater. Today Proc. 2021, 38, 3358–3364. [Google Scholar] [CrossRef]
  20. Sabry, I.; El-Kassas, A.M.; Mourad, A.H.I.; Thekkuden, D.T.; Qudeiri, J.A. Friction Stir Welding of T-Joints: Experimental and Statistical Analysis. J. Manuf. Mater. Process. 2019, 3, 38. [Google Scholar] [CrossRef]
  21. Li, D.; Yang, X.; Cui, L.; He, F.; Shen, H. Effect of welding parameters on microstructure and mechanical properties of AA6061-T6 butt welded joints by stationary shoulder friction stir welding. Mater. Des. 2014, 64, 251–260. [Google Scholar] [CrossRef]
  22. Choudhary, S.; Choudhary, S.; Vaish, S.; Upadhyay, A.K.; Singla, A.; Singh, Y. Effect of welding parameters on microstructure and mechanical properties of friction stir welded Al 6061 aluminum alloy joints. Mater. Today Proc. 2020, 25, 563–569. [Google Scholar] [CrossRef]
  23. Choudhary, A.K.; Jain, R. Influence of stir zone temperature and axial force on defect formation and their effect on weld efficiency during friction stir welding of AA1100: A simulation and experimental investigation. Mater. Today Commun. 2023, 37, 107413. [Google Scholar] [CrossRef]
  24. Shah, P.H.; Badheka, V. An Experimental Investigation of Temperature Distribution and Joint Properties of Al 7075 T651 Friction Stir Welded Aluminium Alloys. Procedia Technol. 2016, 23, 543–550. [Google Scholar] [CrossRef]
  25. Thapliyal, S.; Mishra, A. Machine learning classification-based approach for mechanical properties of friction stir welding of copper. Manuf. Lett. 2021, 29, 52–55. [Google Scholar] [CrossRef]
  26. Fuse, K.; Venkata, P.; Reddy, R.M.; Bandhu, D. Machine learning classification approach for predicting tensile strength in aluminium alloy during friction stir welding. Int. J. Interact. Des. Manuf. 2024, 19, 639–643. [Google Scholar] [CrossRef]
  27. Verma, S.; Gupta, M.; Misra, J.P. Performance evaluation of friction stir welding using machine learning approaches. MethodsX 2018, 5, 1048–1058. [Google Scholar] [CrossRef] [PubMed]
  28. Sandeep, R.; Natarajan, A. Prediction of peak temperature value in friction lap welding of aluminium alloy 7475 and PPS polymer hybrid joint using machine learning approaches. Mater. Lett. 2022, 308, 131253. [Google Scholar] [CrossRef]
  29. Chansoria, P.; Solanki, P.; Dasgupta, M.S. Parametric study of transient temperature distribution in FSW of 304L stainless steel. Int. J. Adv. Manuf. Technol. 2015, 80, 1223–1239. [Google Scholar] [CrossRef]
  30. Song, M.; Kovacevic, R. Thermal modeling of friction stir welding in a moving coordinate system and its validation. Int. J. Mach. Tools Manuf. 2003, 43, 605–615. [Google Scholar] [CrossRef]
  31. Zhu, X.K.; Chao, Y.J. Numerical simulation of transient temperature and residual stresses in friction stir welding of 304L stainless steel. J. Mater. Process. Technol. 2004, 146, 263–272. [Google Scholar] [CrossRef]
  32. Kumar, R.R.; Kumar, A.; Kumar, A.; Ansu, A.K.; Goyal, A.; Saxena, K.K.; Prakash, C.; Prasad, J.L. Thermal simulation on friction stir welding of AA6061 aluminum alloy by computational fluid dynamics. Int. J. Interact. Des. Manuf. 2024, 18, 3495–3505. [Google Scholar] [CrossRef]
  33. Nandan, R.; Roy, G.G.; Lienert, T.J.; Debroy, T. Numerical modelling of 3D plastic flow and heat transfer during friction stir welding of stainless steel. Sci. Technol. Weld. Join. 2006, 11, 526–537. [Google Scholar] [CrossRef]
  34. Colegrove, P. 3 Dimensional Flow and Thermal Modelling of the Friction Stir Welding Process. Ph.D. Dissertation, University of Adelaide, Adelaide, Australia, 2001. [Google Scholar]
  35. Arora, A.; Nandan, R.; Reynolds, A.P.; DebRoy, T. Torque, power requirement and stir zone geometry in friction stir welding through modeling and experiments. Scr. Mater. 2009, 60, 13–16. [Google Scholar] [CrossRef]
  36. Asmare, A.; Al-Sabur, R.; Messele, E. Experimental Investigation of Friction Stir Welding on 6061-T6 Aluminum Alloy using Taguchi-Based GRA. Metals 2020, 10, 1480. [Google Scholar] [CrossRef]
  37. Mishra, A.; Morisetty, R. Determination of the Ultimate Tensile Strength (UTS) of friction stir welded similar AA6061 joints by using supervised machine learning based algorithms. Manuf. Lett. 2022, 32, 83–86. [Google Scholar] [CrossRef]
Figure 1. (a) FSW Al 6061 T6 plate and pin (meshed model) and (b) mesh patterns.
Figure 1. (a) FSW Al 6061 T6 plate and pin (meshed model) and (b) mesh patterns.
Jmmp 09 00246 g001
Figure 2. Process flow.
Figure 2. Process flow.
Jmmp 09 00246 g002
Figure 3. FSW experimental setup.
Figure 3. FSW experimental setup.
Jmmp 09 00246 g003
Figure 4. (a) FSW fixture; (b) location of thermocouples; (c) attachment of thermocouples to FSW plate.
Figure 4. (a) FSW fixture; (b) location of thermocouples; (c) attachment of thermocouples to FSW plate.
Jmmp 09 00246 g004aJmmp 09 00246 g004b
Figure 5. Tools’ geometry and design.
Figure 5. Tools’ geometry and design.
Jmmp 09 00246 g005
Figure 6. (a) Temperature distribution on the Al 6061 T6 plate and (b) temperature contours on the Al 601 T6 plate.
Figure 6. (a) Temperature distribution on the Al 6061 T6 plate and (b) temperature contours on the Al 601 T6 plate.
Jmmp 09 00246 g006
Figure 7. Comparison of peak temperatures obtained from FEA and experiments: (a) comparison of Exp # 1 with FEA in Table 5; (b) comparison of Exp # 2 with FEA in Table 5; (c) comparison of Exp # 3 with FEA in Table 5; (d) comparison of Exp #4 with FEA in Table 5.
Figure 7. Comparison of peak temperatures obtained from FEA and experiments: (a) comparison of Exp # 1 with FEA in Table 5; (b) comparison of Exp # 2 with FEA in Table 5; (c) comparison of Exp # 3 with FEA in Table 5; (d) comparison of Exp #4 with FEA in Table 5.
Jmmp 09 00246 g007
Figure 8. Temperature distributions in FSW of Al 6061 T6. (a) Temperature distribution according to Exp # 1 welding parameters in Table 5; (b) temperature distribution according to Exp # 2 welding parameters in Table 5; (c) temperature distribution according to Exp # 3 welding parameters in Table 5; and (d) temperature distribution according to Exp # 4 welding parameters in Table 5.
Figure 8. Temperature distributions in FSW of Al 6061 T6. (a) Temperature distribution according to Exp # 1 welding parameters in Table 5; (b) temperature distribution according to Exp # 2 welding parameters in Table 5; (c) temperature distribution according to Exp # 3 welding parameters in Table 5; and (d) temperature distribution according to Exp # 4 welding parameters in Table 5.
Jmmp 09 00246 g008
Figure 9. (a) Observation plot, (b) normal probability plot, and (c) main effect plot for max. temperature.
Figure 9. (a) Observation plot, (b) normal probability plot, and (c) main effect plot for max. temperature.
Jmmp 09 00246 g009
Figure 10. Confusion matrix for (a) Logistic Regression, (b) k-NN, and (c) Naive Bayes.
Figure 10. Confusion matrix for (a) Logistic Regression, (b) k-NN, and (c) Naive Bayes.
Jmmp 09 00246 g010
Figure 11. ROC curves for (a) Logistic Regression, (b) k-NN, and (c) Naive Bayes.
Figure 11. ROC curves for (a) Logistic Regression, (b) k-NN, and (c) Naive Bayes.
Jmmp 09 00246 g011
Table 1. Material properties of Al 6061 T6 [32].
Table 1. Material properties of Al 6061 T6 [32].
Temperature
(K)
Density
(kg/m3)
Specific Heat (cp)
(J/kg K)
Thermal Conductivity, K
(W/m K)
2982700896180
3112685920187
3662685978194
42226671004201
47726571028206
53326571052214
58926301078220
64426201104227
70026021133233
Table 2. FSW analysis parameters in COMSOL.
Table 2. FSW analysis parameters in COMSOL.
ParametersSymbolValue
Ambient temperatureTo300 [K]
Convection heat transfer coefficient on the top portionhup12.25 [W/m2 K]
Convection heat transfer coefficient on the bottom portionhdown61.25 [W/m2 K]
Surface emissivity ε (epsilon)0.3
Welding speeduVariable [mm/s]
Coefficient of friction μ Variable
Rotational speed n Variable [rpm]
Angular velocity ω o 2 π n [rad/s]
Normal force (axial force)FnVariable [kN]
Radius of pin R p i n Variable [mm]
Shoulder radius R s h l Variable [mm]
Shoulder surface area A s π ( R s h l 2 R p i n 2 )
Mechanical efficiency (fraction of deformational work converted to heat) η 0.9
Yield stress Y ¯ [18]
Table 3. Levels of FSW process parameters.
Table 3. Levels of FSW process parameters.
FactorLevel 1Level 2Level 3Level 4
Pin Dia (mm)3456
Shoulder Dia (mm)9121518
Tool Rotational Speed (rpm)60090012001800
Welding Speed (mm/min)20385080
Axial Force (kN)15810
Coefficient of Friction0.30.330.360.40
Tool MaterialH13M2----
Table 4. Peak temperatures obtained from COMSOL FEM simulations.
Table 4. Peak temperatures obtained from COMSOL FEM simulations.
S No. #Tool MaterialPin Dia (mm)Shoulder Dia (mm)Tool Rotational Speed (rpm)Welding Speed (mm/min)Axial Force (kN)Coff. of FrictionPeak Temperature (°C)
1H13396002010.3204.859
2H133129003850.33347.765
3H1331512005080.36659.193
4H13318180080100.4677.67
5H13496003850.36310.307
6H134129002010.4257.821
7H13415120080100.3658.577
8H1341818005080.33668.699
9H135990050100.3585.227
10H135126008080.33358.117
11H1351518002050.36659.917
12H1351812003810.4287.83
13H13699008080.36582.452
14H1361260050100.4555.809
15H1361518003810.3325.941
16H1361812002050.33645.24
17M239180020100.33679.418
18M231212003880.3529.137
19M23159005050.4393.454
20M23186008010.36174.855
21M24918003880.4660.469
22M2412120020100.36660.224
23M24159008010.33220.454
24M24186005050.3269.265
25M25912005010.33273.268
26M251218008050.3549.489
27M25156002080.4544.746
28M251890038100.36659.703
29M26912008050.4568.452
30M261218005010.36342.561
31M261560038100.33506.463
32M26189002080.3658.856
Table 5. Comparison of experimental and FEA results.
Table 5. Comparison of experimental and FEA results.
Exp No.Pin Dia (mm)Shoulder Dia (mm)Tool Rotational Speed (rpm)Transverse
Speed (mm/min)
Axial Force (kN)Peak Temperature, FEA
(°C)
Peak Experimental Temperature (°C)
139600201204.85195.10
2412900201257.82239.40
35181200381287.83279.30
46151800381325.94314.40
Table 6. Response table for signal-to-noise ratio (using the “smaller is better” criterion).
Table 6. Response table for signal-to-noise ratio (using the “smaller is better” criterion).
LevelTool MaterialPin Dia (mm)Shoulder Dia (mm)Tool Rotational Speed (rpm)Welding Speed (mm/min)Axial Force (kN)Coff. of Friction
1−53.1−52.24−52.95−50.54−53.90−48.13−52.81
2−52.94−52.40−52.67−52.65−52.68−52.96−52.62
3 −53.33−53.37−54.10−52.88−55.15−53.28
4 −54.11−53.09−54.79−52.62−55.84−53.38
Delta0.161.870.704.251.277.710.76
Rank7362415
Table 7. Response table for mean values.
Table 7. Response table for mean values.
LevelTool MaterialPin Dia (mm)Shoulder Dia (mm)Tool Rotational Speed (rpm)Welding Speed (mm/min)Axial Force (kN)Coff. of Friction
1486.6458.3483.1365.6538.9260.9472.7
2480.7463.2450.1463.2453.5468.0462.4
3 489.8496.1535.2468.4582.7506.2
4 523.2505.3570.5473.8622.9493.3
Delta5.964.955.1205.085.4361.943.7
Rank7452316
Table 8. ANOVA analysis.
Table 8. ANOVA analysis.
SourceDFAdj SSAdj MSF-Value p-Value
Tool Material12802800.120.731
Pin Dia (mm)321,30971033.160.064
Shoulder Dia (mm)313,97646592.070.158
Tool Rotational Speed (rpm)3196,58065,52729.110.000
Welding Speed (mm/min)334,33811,4465.080.017
Axial Force (kN)3632,325210,77593.640.000
Coeff of Friction3936031201.390.294
Error1227,0122251
Table 9. Factorial design of selected parameters.
Table 9. Factorial design of selected parameters.
FactorLevel 1Level 2Level 3Level 4
Pin Dia (mm)456---
Shoulder Dia (mm)121518---
Tool Rotational Speed (rpm)6709501320---
Welding Speed (mm/min)204585---
Axial Force (kN)1359
Tool MaterialH13 Tool Steel
Table 10. Evaluation metrics of classifier models.
Table 10. Evaluation metrics of classifier models.
Classifier ModelAccuracyF1 ScorePrecisionRecallError
Logistic Regression0.98140.95770.97140.94440.0180
k-NN0.95980.90640.94020.87500.0401
Naive Bayes0.92590.82080.88700.76380.0740
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Anis, A.; Shakaib, M.; Hanif, M.S. Statistical and Machine Learning Classification Approaches to Predicting and Controlling Peak Temperatures During Friction Stir Welding (FSW) of Al-6061-T6 Alloys. J. Manuf. Mater. Process. 2025, 9, 246. https://doi.org/10.3390/jmmp9070246

AMA Style

Anis A, Shakaib M, Hanif MS. Statistical and Machine Learning Classification Approaches to Predicting and Controlling Peak Temperatures During Friction Stir Welding (FSW) of Al-6061-T6 Alloys. Journal of Manufacturing and Materials Processing. 2025; 9(7):246. https://doi.org/10.3390/jmmp9070246

Chicago/Turabian Style

Anis, Assad, Muhammad Shakaib, and Muhammad Sohail Hanif. 2025. "Statistical and Machine Learning Classification Approaches to Predicting and Controlling Peak Temperatures During Friction Stir Welding (FSW) of Al-6061-T6 Alloys" Journal of Manufacturing and Materials Processing 9, no. 7: 246. https://doi.org/10.3390/jmmp9070246

APA Style

Anis, A., Shakaib, M., & Hanif, M. S. (2025). Statistical and Machine Learning Classification Approaches to Predicting and Controlling Peak Temperatures During Friction Stir Welding (FSW) of Al-6061-T6 Alloys. Journal of Manufacturing and Materials Processing, 9(7), 246. https://doi.org/10.3390/jmmp9070246

Article Metrics

Back to TopTop