Next Article in Journal
Oxidation Behavior of Lightweight Al0.2CrNbTiV High Entropy Alloy Coating Deposited by High-Speed Laser Cladding
Previous Article in Journal
Study on the Refining Behavior of Al-Si Alloys Doped with Al-Ti-C Intermediate Alloys at Different B Doping Temperatures
Previous Article in Special Issue
Microstructures and Properties of Laser-Cladded FeCoCrNiAlTi High-Entropy Alloy with Intensive Repair Potential
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimization of Laser Cladding Parameters for High-Entropy Alloy-Reinforced 316L Stainless-Steel via Grey Relational Analysis

1
School of Power and Mechanical Engineering, Wuhan University, Wuhan 430072, China
2
School of Physics and Technology, Wuhan University, Wuhan 430072, China
3
College of Engineering, Huazhong Agricultural University, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Coatings 2024, 14(9), 1103; https://doi.org/10.3390/coatings14091103
Submission received: 16 July 2024 / Revised: 19 August 2024 / Accepted: 22 August 2024 / Published: 1 September 2024
(This article belongs to the Special Issue Research and Application of High Entropy Alloys)

Abstract

:
Laser cladding technology serves as a pivotal technique in industrial production, especially in the realms of additive manufacturing, surface enhancement, coating preparation, and the repair of part surfaces. This study investigates the influence of metal powder composition and processing parameters on laser cladding coatings utilizing the Taguchi orthogonal experimental design method. To optimize the laser cladding parameters, multi-response grey relational analysis (GRA) was employed, aiming to improve both the microhardness and the overall quality of the coatings. The optimal parameter combinations identified through GRA were subsequently validated through experimental tests. The results reveal that the microhardness and quality of the coatings are substantially influenced by several critical factors, including the powder feed rate, laser power, high-entropy alloy (HEA) addition rate, scanning speed, and substrate tilt angle. Specifically, the powder feed rate exerts the most significant effect on the microhardness, dilution rate, and average contact angle. In contrast, laser power primarily impacts the mean contact angle difference. The HEA addition rate notably affects the mean contact angle difference, while the scanning speed affects the microhardness and the substrate tilt angle influences the average contact angle. The results of the validation experiment showed a deviation of only 0.95% from the predicted values, underscoring the efficacy of the grey relational analysis (GRA) in optimizing the laser cladding process parameters. The methodology presented in this paper can be applied to determine the ideal processing parameters for multi-response laser cladding processes, encompassing applications such as surface peening and surface repair.

1. Introduction

Laser cladding is an advanced technology utilized for surface strengthening and repair. In this process, a laser creates a metallurgically bonded, dense coating on the surface of a substrate, enhancing properties such as hardness, abrasion resistance, corrosion resistance, and oxidation resistance [1,2,3,4,5,6,7]. As a consequence, laser cladding technology has seen widespread application in specialized sectors, such as aerospace, biomedical fields, and power transmission [8,9,10,11]. High-entropy alloys are composed of various metallic elements with small differences in mixing enthalpies and atomic sizes [12], requiring rapid cooling rates during the preparation process. Among existing methods for coating preparation, metal matrix composite coatings produced via electrochemical deposition are significantly affected by the deposition rates of constituent elements. This dependency leads to reduced processing efficiency, instability of the coatings, and increased susceptibility to secondary pollution [13]. Thermal spraying methods for composite coatings are susceptible to microcracks and other defects, which can be mitigated by increasing the temperature; however, this approach leads to higher energy consumption [14]. Vapor-phase deposition techniques, particularly magnetron sputtering, demand precise control over coating thickness and operational parameters, making them economically costly [15]. In contrast, laser cladding technology provides rapid cooling and shaping, high bonding strength with the substrate [16], and stable properties of the resulting composite coatings. Consequently, employing laser cladding technology for preparing metal matrix composite coatings holds considerable promise.
Laser cladding is a complex procedure, with the quality of the resulting coating primarily controlled by process parameters, including the HEA addition rate, laser power, scanning speed, powder feed rate, and substrate tilt angle [2,3,4]. Numerous studies have analyzed the interrelationships between process parameters and the quality of coatings produced by laser cladding. For instance, Tan et al. examined the interactions between laser power, powder composition, and substrate tilt using analytical models and deposition profile measurements. Their results indicated that the substrate tilt angle significantly affects the laser energy density and powder flow distribution on the tilted substrate [2]. Additionally, Zhang et al. systematically investigated the impact of scanning speed on the microstructure and mechanical properties of samples produced by laser melting FeCoNiAlTi HEA powder mixed with 316L stainless-steel powder [3]. Kahya et al. employed laser cladding technology to prepare 316L stainless-steel and investigated the effects of laser power, scanning speed, and powder feed rate on the material’s microstructure and mechanical properties, determining that the deposition rate was directly proportional to the energy density but inversely proportional to the powder feed rate [4]. Wang et al. analyzed the effects of process parameters on the width and height of the cladding layer using orthogonal tests. The results showed that the width of the cladding layer was primarily influenced by laser power, followed by powder feed rate and scanning speed [5]. Yang et al. explored the relationship between laser cladding process parameters and the geometrical characteristics of a single-track cladding layer using a regression analysis model, establishing a simplified relationship between the process parameters, molten pool characteristics, and the quality of laser-clad Ti-1023 titanium alloy [6]. To address the challenge of achieving both high strength and high ductility in the additive manufacturing of metal matrix composites, Zhang et al. innovatively selected FeCoNiCr HEA as the reinforcing phase to improve the properties of 316L stainless-steel processed by laser powder bed melting. This approach successfully enhanced the strength, ductility, and corrosion resistance of the stainless-steel composites, demonstrating that high-performance HEA particles with lattice structures and elemental types similar to those of the stainless-steel matrix have significant potential as a reinforcing phase [7].
Most of the aforementioned studies primarily focus on analyzing the influence of specific process parameters on the quality of coatings prepared by laser cladding. However, in practical industrial production, the application of laser cladding technology for additive manufacturing or surface coatings on non-horizontal substrates presents an unavoidable challenge. Currently, laser cladding frequently presents various defects during the fabrication of high-entropy alloys, which poses challenges for industrial-scale production [17]. Nevertheless, numerous studies have demonstrated that the morphology, microstructure, and properties of high-entropy alloys are significantly affected by the processing parameters employed during laser cladding [18]. Therefore, optimizing these parameters has the potential to substantially enhance the characteristics of the resulting high-entropy alloys. When laser cladding processes involve tilted substrates, it is crucial to consider the synergistic effects of metal powder composition, laser power, scanning speed, and powder feed rate on the quality of the coatings. Addressing these industrial challenges requires studying and predicting the optimal process parameters for laser cladding to achieve multi-objective optimization of coating quality. In this study, grey relational analysis (GRA) is employed to effectively optimize process parameters across multiple objectives, in contrast to previous methods like ANOVA, which are confined to single-objective optimization [19]. Additionally, single-track laser cladding layers are beneficial for analyzing the cladding process and investigating coating morphology.
This study aims to investigate the effects of the compositional and processing parameters of HEA-added metal powders on coatings prepared by single-track laser cladding. Utilizing multi-response grey relational analysis (GRA), this study seeks to identify the laser cladding processing parameters that optimize microhardness and ensure high coating quality.

2. Materials and Methods

The substrate was made of Q235 steel with dimensions of 100 mm × 50 mm × 5 mm. The elemental composition of Q235 steel is detailed in Table 1. The high-entropy alloy (FeCuNiCrAl) powder employed in the laser cladding process was supplied by Willary New Materials Technology Co., Ltd. (Xuzhou, China), featuring a particle size range of 45–105 µm. The elemental composition of the FeCuNiCrAl HEA is presented in Table 2, and Figure 1 illustrates the microstructure of the HEA powder. Additionally, the 316L stainless-steel powder used in the laser cladding process was provided by Lianhong New Material Science and Technology Co., Ltd. (Tengzhou, China), with a particle size range of 50–150 µm. Both the HEA powder and the 316L stainless-steel powder met the feeding specifications required by the laser cladding system. The elemental composition of 316L stainless-steel is provided in Table 3.
The laser cladding system utilized in this study, as depicted in Figure 2, consisted of an industrial robot laser cladding integrated system (HWL-RF2000W, Huawei Laser, Dongguan, China), a control system (PLC, Mitsubishi, Japan), and pneumatic powder feeding equipment (HW-05SF, Huawei Laser, Dongguan, China). Throughout the laser cladding process, the metal powder was protected by argon gas.
Prior to the laser cladding, the Q235 substrate was polished to remove the oxide layer, followed by rinsing and drying with anhydrous ethanol. The metal powder was dried under vacuum at 100 °C for two hours before use. During the laser cladding process, the laser was focused on the molten pool, causing the melted metal powder to form a metallurgical bond with the substrate material upon cooling [20]. Post-cladding, the samples were cut, mounted, ground, and polished for evaluation.
The morphological characteristics of the coatings were examined using an optical microscope (Axio Lab. A1, Carl Zeiss AG, Oberkochen, Baden-Württemberg, Germany). The Vickers hardness of the coatings was measured from the fusion zone towards the coating direction using an HXS-1000A microhardness tester (load: 3 N, dwell time: 10 s). Measurements were taken at 0.1 mm intervals between consecutive points, and the averaged measurements were used to characterize the Vickers hardness of the laser fusion coating. The phase composition and crystal structure of the coating were analyzed using an X-ray single crystal diffractometer (XtaLAB Synergy Custom, Rigaku, Tokyo, Japan).
In this study, the Taguchi orthogonal method was employed to design the single-track laser cladding experiment. Five process parameters were evaluated: HEA addition rate, laser power, scanning speed, powder feed rate, and substrate tilt angle. Given that each process parameter has four levels, a full factorial experiment would necessitate 45 = 1024 different combinations. However, by using the Taguchi orthogonal design, only 16 runs were required, significantly reducing the number of experiments needed and conserving both manpower and resources [21]. The Taguchi orthogonal array is a balanced design that allows each factor to be evaluated independently. Table 4 presents the factors and their corresponding levels used in the Taguchi orthogonal experimental design.
The key responses in this study included microhardness (H), dilution rate (λ), average contact angle ( a ¯ ), and mean difference of contact angles (Δa). These parameters are represented by Equations (1)–(3), respectively [22,23,24].
λ = A 2 / A 1 + A 2
a ¯ = a 1 + a 2 2
Δ a = a 1 a 2 2
where A 2 is the area of the substrate melted during the laser cladding process in the center cross-section, while A 1 is the area of the cladding layer obtained in the center cross-section. The parameters a 1 and a 2 are the contact angles on the left and right sides of the sample in the center cross-section, respectively (Figure 3).
The signal-to-noise ratio (SNR or S/N) is a crucial parameter for predicting optimization results in data analysis, as it evaluates the error between the actual response and the expected estimate [25]. The response is considered the signal, while the irregular fluctuations that occur with responses are regarded as noise. The SNR quantifies the relationship between the noise and the signal, thereby enabling a quantitative evaluation of the response effect [26].
The signal-to-noise ratio for microhardness (H) was calculated by Equation (4). For microhardness (H), a higher value was preferable [27].
S / N = 10 log 1 n Σ y i 2
The signal-to-noise ratios for the dilution rate (λ), the average contact angle ( a ¯ ), and the mean difference of contact angles ( Δ a ) were calculated by Equation (5). For these parameters, lower values were preferable [28,29,30].
S / N = 10 log 1 n Σ y i 2
where n is the number of experiments in this study, which is 2, and yi represents the different values of the response target, which are the results of the experiments [27].
Converting the experimental results into corresponding signal-to-noise ratios was pivotal for optimizing the process parameters. Higher signal-to-noise ratios indicated superior outcomes for microhardness (H), dilution rate (λ), average contact angle ( a ¯ ), and mean difference of contact angles ( Δ a ). Therefore, the optimal process parameter level was determined based on achieving the highest signal-to-noise ratio [25,26]. However, multi-objective process parameter optimization posed greater complexity compared to single-objective optimization.
A higher signal-to-noise ratio for one performance characteristic might correlate with a lower signal-to-noise ratio for another. Consequently, an overall assessment of the signal-to-noise ratio was essential to optimize multiple performance characteristics [31]. To address this challenge, this study employed the grey relational analysis (GRA) method to facilitate multi-objective optimization using the Taguchi orthogonal method, aiming to optimize multiple quality characteristics of laser cladding coatings. In GRA, the S/N ratio was initially normalized to a range from zero to unity. Subsequently, the grey relational coefficient was derived from the normalized S/N ratio to illustrate the relationship between the desired and actual S/N ratios. The grey relational grade was then calculated by aggregating the weighted grey relational coefficients associated with each performance characteristic. The comprehensive assessment of the multiple characteristics relied on the grey relational grade, thereby simplifying the optimization of complex characteristics into the optimization of a single grey relational grade [32].

3. Results

Table 5 displays the 16 runs conducted in the Taguchi orthogonal experiment, detailing their respective process parameter configurations. Additionally, it presents the experimental data for microhardness (H), dilution rate (λ), average contact angle ( a ¯ ), and mean difference of contact angles (Δa).
The optimal process parameter set optimized by GRA (A4B1C2D1E4) was selected, with only the HEA addition rate being varied, to investigate its effect on the phase composition and crystal structure of the coating. The experimental design is shown in Table 6.
Figure 4 shows the X-ray diffraction (XRD) results of the laser cladding coatings as the HEA addition rate increases from 0% to 15%. The diffraction peak identification results indicated that four laser cladding layers are composed of a Laves phase, FCC phase, and BCC phase [33,34,35]. As one can see, the intensity of the diffraction peaks corresponding to the Laves phase and FCC phase (111) in the cladding coatings gradually strengthens with the increasing HEA addition rate, indicating a progressive increase in these phases. The lattice constants of each phase in the laser cladding coatings, determined using Equation (6), exhibit variation with different HEA addition rates. The calculated values are provided in Table 7 [36,37].
a = w 2 s i n θ h 2 + k 2 + l 2
where w represents the wavelength of the X-ray; h, k, and l are the Miller indices of the crystal plane; θ is the diffraction angle; and a denotes the lattice constant.
The results indicate that as the HEA addition rate increases, the lattice constant of the Laves phase initially decreases and subsequently increases, while the FCC and BCC phases display an opposite trend. Specifically, at an HEA addition rate of 5%, the lattice constants of both the Laves and FCC phases reach their maximum values, whereas the lattice constant of the BCC phase peaks at a 10% HEA addition rate. This phenomenon suggests that within the 0% to 15% range of HEA addition, a diffusion process involving solute elements with varying atomic radii occurs between the Laves phase and the dual-phase solid solution of BCC and FCC, leading to lattice distortion in each phase [38,39]. Consequently, the variation in coating performance is determined by the combined effects of the increased presence of the Laves and FCC phases, as well as the associated lattice distortion.

3.1. Analysis of Microhardness

The highest microhardness among the 16 experiments was achieved in the 13th run, characterized by process parameters set to A4B1C4D2E3 (HEA addition rate of 15%; laser power at 750 W; scanning speed of 11 mm/s; powder feed rate of 5 r/min; substrate tilt angle of 20°). Table 8 presents the signal-to-noise response of the microhardness data. The Delta values indicate that the order of factors affecting microhardness is as follows: powder feed rate > scanning speed > substrate tilt angle > laser power > HEA addition rate.
Figure 5 shows the main effects plot for the signal-to-noise ratio of microhardness. Within the figure, solid lines connect the mean values within each category, while a dashed reference line denotes the overall mean. The main effect of each categorical variable was assessed by comparing its solid plotted line with the dashed reference line. A horizontal solid line indicates the absence of a main effect for that parameter, whereas any deviation indicates its presence. A steeper slope suggests a stronger influence of the parameter on the results.
Figure 5 illustrates that the powder feed rate exerts the most significant effect on microhardness, consistent with the findings in Table 8. The optimal microhardness was achieved at a powder feed rate of 2 r/min. The powder feed rate also exhibited a substantial influence on microhardness in the investigation conducted by Dong et al. [40]. Increasing the powder feed rate during laser cladding facilitated greater entry of metal powder into the melt pool. Initially, lower feed rates contributed to strengthened composite coatings due to factors such as solution strengthening and dislocation strengthening, resulting in elevated microhardness [41]. However, further increments in the powder feed rate intensified shielding effects on the laser, resulting in thicker coatings, and reduced fusion depth [42]. Furthermore, the increased introduction of powder into the substrate reduces the energy input, which, in turn, diminishes the influx of elements into the melt pool [40,43]. Therefore, the microhardness is relatively low.
Similar patterns were observed for laser power, scanning speed, and substrate tilt angle. Initially, laser power decreased microhardness from 750 to 1100 W, possibly attributed to abrupt structural and compositional changes [44]. Subsequently, an increase in microhardness was observed from 1100 to 1450 W, stemming from enhanced melting and denser crystal growth. However, within the power range of 1450–1800 W, grain coarsening led to a reduction in microhardness [45].
Scanning speed influenced microhardness by regulating heat input and thermal effects. Low speeds (5 to 7 mm/s) resulted in reduced densification, coarser structures, and decreased microhardness. In contrast, intermediate speeds (7 to 9 mm/s) enhanced cooling rates, leading to increased microhardness due to thinner fused layers and higher levels of amorphous structure [46]. High speeds (9 to 11 mm/s) advanced solidification but could induce pore formation, thereby reducing coating strength.
Increasing the substrate tilt angle from 0 to 10° enhanced molten metal flow, prolonged melting times, and promoted grain coarsening, resulting in reduced hardness. Angles between 10 and 20° improved the adhesion of powder particles and reduced coating defects, thereby gradually increasing hardness [47]. However, tilt angles from 20 to 30° increased heat input, slowed cooling rates, and further coarsened the grain structure, leading to decreased coating hardness.
As illustrated in Figure 4, with increase in the HEA addition rate from 0% to 10%, there is a slight fluctuation in the proportion of the Laves, FCC, and BCC phases in the laser cladding coatings. The Vickers hardness of the BCC phase is higher than that of the FCC phase, and the presence of the Laves phase can enhance the microhardness of coating. The synergistic effects of these three phases lead to an overall increase in microhardness. However, when the HEA addition rate is elevated to 15%, the proportion of the BCC phase in the coating significantly decreases, resulting in a reduction in microhardness. Additionally, as indicated in Table 7, the variations in the lattice constants of the BCC, FCC, and Laves phases influence grain refinement strengthening, which subsequently contributes to the increase in microhardness.
Based on the findings from the above analyses, the optimal processing parameters for maximizing microhardness were as follows: 1450 W laser power, 9 mm/s scanning speed, 2 r/min powder feed rate, and 20° substrate tilt angle.

3.2. Analysis of Dilution Rate

The optimal dilution rate was observed in the 9th run with the process parameters set to A3B1C3D4E2 (HEA addition of 10%; laser power of 750 W; scanning speed of 9 mm/s; powder feed rate of 11 r/min; substrate tilt angle of 10°). Table 9 presents the signal-to-noise response data for the dilution rate. The Delta values indicate the order of priority for the factors as follows: powder feed rate > laser power > substrate tilt angle > scanning speed > HEA addition rate.
Figure 6 illustrates the main effect plot, highlighting that the powder feed rate significantly influenced the dilution rate of the coating. As the powder feed rate increased, there was a corresponding decrease in the dilution rate. This finding is consistent with previous study [43,48]. This phenomenon arose because higher rates of powder delivery introduced more metal powder into the melting process, which reduced the width and depth of substrate melting but increased the height of melting. Higher powder delivery rates could induce a “thermal shielding” effect on the substrate surface, thereby diminishing the energy input of the laser [49]. Firstly, this approach reduces the depth of the melting area [49]; conversely, it minimizes the ingress of elements into the substrate melting area, ultimately resulting in a decreased dilution rate [48].
Conversely, higher laser power was associated with an increased dilution rate. Elevated laser power expanded the width of the cladding layer, thereby enhancing the dilution rate. This phenomenon stemmed from the heightened temperature and increased energy absorption in the melt pool, which enlarged both the area of substrate melting and the amount of energy absorbed by the powder per unit time. Consequently, as the melting and solidification processes progressed, both the cladding width and dilution rate increased [50].
An increase in the substrate tilt angle from 0° to 10° significantly influenced the liquid-solidification process within the melt pool. This adjustment reduced both the height and depth of the melt gradually, altering the flow of liquid along the substrate surface due to gravitational effects, and ultimately decreasing the overall dilution rate [51]. As the substrate tilt angle increased from 10° to 30°, the width of the deposition layer progressively enlarged, accompanied by a greater offset in layering. Within this range, the dilution rate showed a positive correlation with the tilt angle.
Increasing the scanning speed from 5 mm/s to 9 mm/s resulted in a more pronounced reduction in melt height compared to melt depth, thereby increasing the dilution rate. Further increasing the scanning speed from 9 mm/s to 11 mm/s significantly decreased the interaction time between the laser beam, powder, and substrate surface per unit length. This reduction led to a decrease in the volume of melted powder, shallower depth of the melt pool, and subsequently lowered the dilution rate [52].
The XRD results in Figure 4 indicate that the trend in the FCC phase proportion aligns with the trend in dilution rate, implying that variations in dilution rate may be associated with the formation of the FCC phase and the phase transformation between FCC and BCC [53]. Furthermore, variations in the proportion of HEA powder added to the process could influence the overall density of metal powder, potentially leading to fluctuations in dilution rate.
These findings aligned with the ranking of factors based on the signal-to-noise ratio response presented in Table 9, indicating their respective levels of importance. According to the analyses, when focusing solely on minimizing dilution rate, the optimal processing parameters included a laser power of 750 W, scanning speed of 5 mm/s, powder feed rate of 11 r/min, exclusion of HEA powder, and a substrate tilt angle of 10°.

3.3. Analysis of Average Contact Angle

The 14th experimental run produced the smallest mean contact angle, indicating optimal powder-substrate “wettability” under those conditions [54]. The process parameters for this run were set as A4B2C3D1E4 (HEA addition ratio of 15%; laser power of 1100 W; scanning speed of 9 mm/s; powder feed rate of 2 r/min; substrate tilt angle of 30°). Table 10 presents the signal-to-noise response data for the average contact angle. The Delta value reveals the priority order of factors as follows: powder feed rate > substrate tilt angle > laser power > HEA addition ratio > scanning speed.
The main effect plot (Figure 7) illustrates that the powder feed rate exerts the most significant influence on the average contact angle of the coating. The impact of the powder feed rate on the average contact angle contrasted with that of the dilution rate. Increasing the powder feed rate from 2 to 5 r/min enhanced the volume of powder beam melting, thereby enlarging the area of the cladding layer and resulting in a continuous rise in the average contact angle. Within the 5–8 r/min range, the average contact angle stabilized due to constraints within the process window. From 8 to 11 r/min, higher powder feed rates increased the cladding layer height, further elevating the contact angle.
As the substrate tilt angle increased from 0° to 30°, the downward flow of liquid in the molten pool along the substrate surface intensified due to gravity, leading to a decrease in the average contact angle influenced by the Marangoni effect [55].
The laser energy input also significantly affected the flight speed and temperature of molten droplets. Raising the laser power from 750 W to 1450 W facilitated easier spreading of molten metal droplets, resulting in an increased contact angle of the coating. The contact angle reached its peak at 1450 W and remained relatively stable between 1450 W and 1800 W [56].
An increase in the HEA proportion from 0% to 5% reduced the wettability of the metal powder due to the addition of aluminum, which resulted in a higher average contact angle. However, as the HEA proportion rose from 5% to 15%, the synergistic effects of elements such as Fe, Cu, Ni, Cr, and Al altered the solid–liquid interface adsorption behavior and microstructural formation mechanism of the coating. Specifically, the atomic radii of the aforementioned metal elements differ, and as the HEA addition rate increases from 5% to 15%, more solute elements separate from the BCC and FCC dual-phase structure and precipitate to form the Laves phase, which might enhance the interface wettability, consequently decreasing the average contact angle [57,58].
The main effect plot indicated that scanning speed fluctuated around the reference dashed line, suggesting it had minimal influence on the average contact angle.
The analyses were consistent with the prioritization of factors based on the signal-to-noise ratio response in Table 10. For optimizing the average contact angle, the optimal processing parameters were determined to be a laser power of 750 W, scanning speed of 7 mm/s, powder feed rate of 2 r/min, HEA addition ratio of 15%, and a substrate tilt angle of 30°.

3.4. Analysis of Mean Difference of Contact Angles

The 13th experimental run exhibited the smallest mean difference in contact angles, indicating optimal consistency in the coating. The process parameters were configured as A4B1C4D2E3 (15% HEA addition ratio; 750 W laser power; 11 mm/s scanning speed; 5 r/min powder feed rate; 20° substrate tilt angle). Table 11 presents the signal-to-noise response data for the mean difference in contact angles. The Delta values indicate the prioritization of factors as follows: laser power > HEA addition rate > scanning speed > powder feed rate > substrate tilt angle.
The main effect plot (Figure 8) illustrates that laser power has the greatest influence on the mean difference of contact angles in the coating. As the laser power increased from 750 W to 1450 W, the contact angle difference amplified. This effect was attributed to the higher laser power, which intensified energy input and consequently led to greater variation in the mean difference of contact angles. Further increasing the laser power from 1450 W to 1800 W enhanced the energy input, promoted complete melting of the metal powder, and improved the fluidity of the molten pool. This resulted in a slight reduction in the mean difference in contact angle due to surface tension [59].
The mean difference of contact angles followed an increasing-then-decreasing trend with rising HEA addition rates, mirroring the effect observed in average contact angle variations influenced by HEA proportions. At 0%–5% HEA addition rates, the introduction of aluminum reduced the wettability of the metal powder and the mobility of the molten pool, thereby increasing the mean difference of contact angles. Conversely, in the 5% to 15% range, synergistic effects between HEA and 316L hybrid powder components improved wettability and solid–liquid interfacial adsorption behavior, leading to a decrease in the mean difference of contact angles [60].
Increasing the scanning speed from 5 mm/s to 7 mm/s resulted in negligible change in the mean difference of contact angles. Within the range of 7 mm/s to 9 mm/s, higher scanning speeds reduced the laser energy density, weakened the flow capacity of the liquid molten cladding layer, and consequently accelerated the increase in the mean difference of contact angles. Despite the consistent amount of metal powder melting per unit time, scanning speeds from 9 mm/s to 11 mm/s significantly reduced the area of the cladding layer, leading to a sharp decline in the contact angle difference [61].
An increase in the powder feed rate from 2 r/min to 5 r/min escalated the amount of powder beam melting and expanded the area of the fused cladding layer, thereby increasing the mean difference in contact angles. Within the range of 5 r/min to 11 r/min, higher feed rates raised the height of the molten cladding layer, altering the distribution of surface tension effects. As the height increased, surface tension exerted a significant influence on the morphology of the melted cladding layer, resulting in a decrease in the mean difference in contact angles [62]. Importantly, higher feed rates led to more pronounced attenuation of powder interaction with the laser beam, which slowed the reduction trend of the contact angle difference in the 8 r/min to 11 r/min region, as evidenced in the main effects plot.
From the main effect plot, it was evident that the substrate tilt angle fluctuated around the reference dashed line, indicating minimal impact on the mean difference of contact angles.
The results from the aforementioned analyses corroborated the prioritization of factors as indicated by the signal-to-noise ratio response in Table 11. Considering solely the mean difference of contact angles, the optimal processing parameters identified were as follows: 750 W laser power, 11 mm/s scanning speed, 5 r/min powder feed rate, and an HEA addition rate of 15%.

3.5. Multi-Response Grey Relational Analysis

The objective of this study was to enhance both the microhardness and the quality of laser cladding layers simultaneously. Since the optimal outcomes for microhardness, dilution rate, average contact angle, and mean difference of contact angles were observed under different combinations of process parameters, further exploration was necessary for multi-response optimization. To achieve higher microhardness and improved quality of the cladding layer, multi-response grey relational analysis was employed to ascertain the optimal process parameters.
Developed by Julong Deng in the 1980s, grey relational analysis effectively integrates multiple objectives into a unified metric, addressing complex multi-response problems. This method offers an efficient approach to optimizing multiple parameters to achieve an optimal solution that meets diverse objectives [63,64]. Grey relational analysis involves three main steps [63,64,65,66]. The first step is data normalization, which adjusts for differences in units and ranges of the original data. Normalized data are expressed as values between 0 and 1. In this study, microhardness (H) was normalized using Equation (7), where larger values indicate better performance. The dilution rate (λ), the average contact angle ( a ¯ ), and the mean difference of contact angles (Δa) were normalized by Equation (8), where smaller values indicate superior performance.
Y i k = X i k min X i k max X i k min X i k
Y i k = max X i k X i k max X i k min X i k
where k denotes the kth response (k = 1, 2) in this study; i represents the ith experimental data (i = 1, 2, 3, …, 16); Yi(k) signifies the normalized response data; Xi(k) denotes the original response data; min Xi(k) and max Xi(k) indicate the minimum and maximum values of Xi(k).
The second step involves calculating the grey relational coefficient for each response using Equation (9).
G R C i k = Δ 0 m i n + ξ · Δ 0 m a x Δ 0 i k + ξ · Δ 0 m a x
where GRCi(k) represents the grey relational coefficient of the kth response in the ith run; ∆0i(k) indicates the deviation between the normalized value Yi(k) and the reference sequence {X0} = {1, 1, 1, 1}. The deviation is calculated as ∆0i(k) = Y 0 k Y i ( k ) ; ∆0min and ∆0max denote the minimum and maximum value of ∆0i(k), respectively; ξ is the distinguishing coefficient with ξ ∈ [0, 1]. A value of 0.5 is selected for ξ to ensure a moderate distinguishing effect and stability [67].
Following the normalization of microhardness, dilution rate, average contact angle, and mean difference of contact angles, the results of grey relational generation (Y) and their deviations (∆0) are presented in Table 12. In the final step, multi-response grey relational calculations are performed using Equation (10).
G R G i = 1 n k = 1 n G R C i k
where GRGi represents the grey relational grade (GRG) for the ith run and n is the number of responses, which in this study is 2.
Table 13 displays the grey relational coefficients (GRCs) and grey relational grades (GRG) for each run, where higher values indicate greater preference.
The analysis of the signal-to-noise (S/N) data in Table 14 reveals that laser power exhibits the largest absolute difference, underscoring its substantial impact on both microhardness and cladding layer quality. Utilizing Table 14, optimal processing parameter settings were identified by selecting the highest value in each category. The optimal combination of parameters identified was A4B1C2D1E4 (15% HEA addition rate; 750 W laser power; 7 mm/s scanning speed; 2 r/min powder feed rate; 30° substrate tilt angle). Since this specific parameter setting was not included in the original 16 runs, a validation experiment was conducted to verify the efficacy of these processing parameters.

3.6. Processing Parameters Optimization and Experimental Validation

The GRG prediction was employed to forecast the outcomes for the parameter setpoint (A4B1C2D1E4) prior to conducting the experiment. This predictive calculation was derived using Equation (11).
G R G p = G R G ¯ + j = 1 r ( G R G j ¯ G R G ¯ )
where GRGp is the predicted value of GRG for the selected set; G R G ¯ represents the overall mean of the GRG; r is the number of processing parameters, which is 5 in this study; and G R G j ¯ denotes the average GRG value at the selected level for the jth processing parameter.
Table 15 compares the optimal outcomes derived from the 16 orthogonal experiments, the predictions from grey relational analysis (GRA), and the results of the validation experiment. Specifically, it compares the best results achieved using configuration A4B1C4D2E3 in the orthogonal experimental design with the optimal parameter settings identified through GRA (A4B1C2D1E4). The microhardness increased significantly from 461.30 HV to 549.14 HV, marking a notable 19.04% improvement. Moreover, the average contact angle decreased impressively from 25.79° to 16.16°, representing a substantial 37.34% reduction. While there was a slight increase in both the dilution rate and the mean difference in contact angles, their impact on the quality of the coating remained negligible.
The Vickers hardness of the Q235 substrate measures 160 HV, while pure 316L stainless-steel registers 200 HV. In contrast, the laser cladding coating achieved a remarkable 549.14 HV—nearly 3.5 times the hardness of the Q235 substrate and 2.75 times that of pure 316L stainless-steel. This finding underscored the potential of the coating for creating robust and wear-resistant surfaces. The average contact angle of the laser cladding coating, measured at 16.16° under a 30° substrate tilt angle, indicated outstanding coating performance alongside its impressive hardness characteristics. In conclusion, the laser cladding technology employed in this study proves highly suitable for repairing and reinforcing parts with intricate geometries and specific tilt angles, offering substantial economic advantages in practical manufacturing applications.
Additionally, the discrepancy between the grey relational grade (GRG) prediction from Equation (11) and the experimental validation was less than one percent (0.95%), demonstrating the high accuracy of multi-response optimization via grey relational analysis. The methods proposed in this study could also assess uneven weighting distributions for metrics such as microhardness, dilution rate, average contact angle, and mean difference of contact angles. For instance, to achieve a laser melting coating with optimal dilution rate properties, the weighting factor in the GRA could be assigned as 60% for dilution rate, 20% for microhardness, and 10% for both the average contact angle and the mean difference of contact angles. Furthermore, this multi-response optimization approach using GRA could guide the design of large-scale experiments, which was crucial given the high cost and time requirements of actual experimentation. It could evaluate various responses, including aspect ratio, surface roughness, wear resistance, and cladding efficiency. Since GRA was designed for multi-response optimization, it could simultaneously optimize five or six of these responses.
Figure 9 shows the cross-sectional microscopic morphology of the laser cladding coatings prepared using the best combination of process parameters from the orthogonal design and the optimized parameter combination from the GRA. The figure illustrates that, compared to the optimal orthogonal parameter combination (a), the optimal GRA parameter combination (b) produced coatings that were devoid of pores and inclusions. The coating and the substrate formed a smooth and natural metallurgical bond, with the substrate being flat and free of undercuts. The surface of the coating was more uniform, rendering it more suitable for multi-pass laser cladding processes with overlapping joints. Additionally, Figure 9a shows that the ratio of the heat-affected zone depth to the cladding layer thickness (D/T) was 1.98, and the height-to-width ratio (H/W) was 0.158. In contrast, Figure 9b reveals that for the coating produced using the grey relational optimization parameters, the D/T value was 0.98 and the H/W value was 0.097. Generally, smaller D/T and H/W values are preferred [68,69]. In the grey relational optimization parameter group, the values of D/T and H/W were reduced by 25.8% and 38.6%, respectively, compared to those in the optimal parameter group identified from the orthogonal design. The lateral extensibility of single-pass coatings is enhanced as the H/W value decreases. Laser cladding parameters optimized via GRA are superior, resulting in laser cladding coatings that exhibit greater lateral extension, fuller morphologies, and smoother fusion with the substrate, ultimately leading to a reduction in the H/W value of the coating [69]. The D/T value quantitatively characterizes the extent of thermal impact on the substrate during processing. As the D/T value increases, the heat-affected zone in the composite coating enlarges, increasing the likelihood of defects such as fine cracks, microcracks, and penetrating cracks, which significantly threaten the coating’s service life. Due to the superior parameters identified by GRA, the heat input during the laser cladding process is more concentrated within the coating, reducing the thermal impact on the substrate and resulting in a lower D/T value in the metal matrix composite coating [68]. These results demonstrate the exceptional quality of the laser cladding coating produced using the optimal processing parameters derived from grey relational analysis.

3.7. Economic, Energy, and Sustainability Analyses

Thermal spray is a traditional technique for producing metal matrix composite coatings. Among its methods, plasma spray has become particularly notable for its ability to enhance the surfaces of substrates or components in demanding and corrosive environments. This technique is distinguished by its high deposition rates and the flexibility to achieve varied coating properties through modifications in processing parameters and ionic characteristics [70,71]. This method draws a parallel with the emerging laser cladding technology. Consequently, this study will compare the economic and energy consumption aspects of both methods during the preparation phase, aiming to demonstrate the cost-effectiveness and sustainability potential of laser cladding coating preparation optimized through grey relational analysis (GRA).
A total of 1000 samples of metal matrix composite coatings from Wuhan, China, were analyzed. Wang et al. [72] conducted a comprehensive examination of the plasma spray process for preparing composite coatings. Therefore, this study compared it with the laser cladding process utilizing the optimal parameters identified through GRA.
In the first step, the labor time costs were analyzed. Calculations based on the time required for multi-pass spraying and the number of passes revealed that plasma spray technology averages approximately 3.17 s for a single pass of about 60 mm, with an additional 254 s allocated to preheating and preparation tasks. In contrast, laser cladding technology using the optimal parameters requires about 0.86 s for the same pass, with an additional 241 s for preheating and preparation. Consequently, the production of 1000 composite coating samples via plasma spraying technology necessitated 71.44 working hours, whereas laser cladding required 67.18 working hours.
In the subsequent phase, the analysis focused on energy costs. Wang et al. demonstrated that the energy consumption for the optimal performance of plasma spray coating was 3.86 × 104 kJ for 25 passes. According to this estimation, the energy consumption for a single-pass process is approximately 0.43 kW·h. In this study, the power consumption of each subsystem involved in the laser cladding process was systematically assessed by referencing the nominal power ratings provided on the equipment nameplates. This analysis determined that the energy consumption for a single-pass operation is approximately 0.29 kW·h. As a result, 430 kW·h of electrical energy was consumed to produce 1000 composite coating samples using plasma spraying technology, whereas laser cladding technology required 290 kW·h.
The third step entails an assessment of sustainability impacts. Carbon emission factors for electricity exhibit considerable variation based on the geographic location and the electricity generation conditions of each country. In this study, we selected the permanent members of the United Nations Security Council as the focus of our analysis and calculated the carbon emissions associated with the production of 1000 composite coating samples in China, the United Kingdom, France, the United States, and the Russian Federation. The detailed findings are presented in Table 16. Carbon emission factor data were derived from sources such as Climate Transparency (2022 Report), UK Govt—Defra/BEIS (2022), Association of Issuing Bodies (AIB) 2022, US Env Protection Agency (EPA), and other pertinent reports [73,74,75,76].
In summary, the production of 1000 metal matrix composite coating samples utilizing laser cladding technology optimized by GRA achieves a reduction in labor time by 4.26 h and a decrease in electricity consumption by 140 kW·h compared to traditional thermal spraying technology. The associated reduction in CO2 emissions ranges from 5.74 kg to 78.01 kg, depending on the electricity generation conditions and the geographic location in the respective countries. Therefore, the application of laser cladding technology for composite coating preparation offers significant advantages in terms of economic cost-effectiveness and sustainability.

4. Conclusions

This study employs an orthogonal experimental design to systematically investigate the impact of varying the HEA addition rate, laser power, scanning speed, powder feed rate, and substrate tilt angle on the microhardness, dilution rate, average contact angle, and mean difference of contact angles in the single-track laser cladding layer. Optimal parameter combinations were identified through signal-to-noise ratio calculations and grey relational analysis (GRA), which were subsequently validated by experimental verification. Based on this investigation, the following conclusions can be drawn:
  • The analysis of microhardness as the sole response variable indicates a notable decrease with higher powder feed rates. Moreover, variations in laser power, scanning speed, and substrate tilt angle initially lead to a reduction in microhardness, followed by subsequent improvements and subsequent declines.
  • The analysis of the dilution rate indicates a significant decrease with higher powder feed rates and an increase with greater laser power. As scanning speed increases, the dilution rate initially rises and then declines. Conversely, as the substrate tilt angle increases, the dilution rate first decreases and then rises. Regarding the average contact angle, there is a progressive increase with higher powder feed rates, while an increase in substrate tilt angle leads to a decrease. The addition of HEA initially raises the average contact angle and subsequently reduces it, whereas higher laser power consistently enhances it. The mean difference in contact angles shows an initial sharp increase followed by a slight decrease with increasing laser power. Increases in the HEA addition rate, scanning speed, and powder feed rate result in an initial rise followed by a decline in the mean difference of contact angles.
  • Grey relational analysis effectively discerns the processing parameters that optimize coating quality. The optimal combination comprises a 15% HEA addition rate, 750 W laser power, 7 mm/s scanning speed, 2 r/min powder feed rate, and a 30° substrate tilt angle. Coatings produced under these conditions exhibit markedly enhanced microhardness and average contact angle compared to those achieved using the optimal parameters identified in the orthogonal test.
  • The validation experiment, conducted using optimal parameter settings identified through grey relational analysis, demonstrates a minimal error of 0.95% relative to the predicted value, underscoring the efficacy of GRA in optimizing laser cladding parameters. The preparation of laser cladding coatings utilizing the optimal parameters derived through GRA demonstrates enhanced economic efficiency and superior environmental sustainability compared to conventional methods. The methodologies presented in this study can also be applied to optimize laser processing parameters in selective laser melting of titanium alloys and laser powder bed fusion of aluminum alloys. These methodologies offer a robust framework for advancing laser cladding technology in industrial applications, thereby facilitating the fabrication of wear-resistant coatings and the enhancement or repair of component surfaces. Specifically, laser cladding coatings optimized through GRA can be used for corrosion resistance treatment of ships in the maritime industry, surface enhancement of specialized parts in the aerospace industry, and service protection of cooling tower components in the power industry.

Author Contributions

Conceptualization, N.L. and G.Z.; data curation, S.G. and Q.F.; formal analysis, S.G., Q.F. and C.C.; funding acquisition, B.Y. and G.Z.; investigation, S.G., M.L. and C.C.; methodology, S.G., M.L., L.H. and G.Z.; project administration, N.L., B.Y. and G.Z.; resources, M.L.; software, S.G. and Q.F.; supervision, N.L., B.Y. and G.Z.; validation, S.G., Q.F. and L.H.; visualization, Q.F., L.H. and N.L.; writing—original draft, S.G.; writing—review and editing, S.G. and M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (52371073) and the Shenzhen Science and Technology Program (JCYJ20220530140601002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jiang, Y.; Tang, H.; Li, Z.; Zheng, D.; Tu, Y. Additive manufactured Mg-Gd-Y-Zr alloys: Effects of Gd content on microstructure evolution and mechanical properties. Addit. Manuf. 2022, 59, 103136. [Google Scholar] [CrossRef]
  2. Tan, H.; Fan, W.; Qian, Y.; Chen, Y.; Liu, S.; Liu, X. Influence of inclined substrate on process characteristics of directed energy deposition. Opt. Laser Technol. 2020, 129, 106288. [Google Scholar] [CrossRef]
  3. Zhang, X.Q.; Yang, D.; Jia, Y.; Wang, G.; Prashanth, K.G. Microstructure evolution and tensile property of high entropy alloy particle reinforced 316 L stainless steel matrix composites fabricated by laser powder bed fusion. J. Alloys Compd. 2023, 965, 171430. [Google Scholar] [CrossRef]
  4. Kahya, H.; Gurun, H.; Kucukturk, G. Experimental and Analytical Investigation of the Re-Melting Effect in the Manufacturing of 316L by Direct Energy Deposition (DED) Method. Metals 2023, 13, 1144. [Google Scholar] [CrossRef]
  5. Wang, J.; Fang, J.; Zhang, P.; Wang, S.; Wang, D. Research on Geometrical Features of FGH95 Nickel-based Superalloy by Coaxial Single Track Laser Cladding. J. Phys. Conf. Ser. 2022, 2390, 012027. [Google Scholar] [CrossRef]
  6. Yang, S.; Xu, Z.; Peng, S.; Cao, S.; Liu, W.; Wang, F. Investigation on the relationships among process parameters, molten pool characteristics, and fabrication quality of Ti-10V-2Fe-3Al by laser direct energy deposition. Optik 2023, 275, 170521. [Google Scholar] [CrossRef]
  7. Zhang, C.; Zhu, J.; Ji, C.; Guo, Y.; Fang, R.; Mei, S.; Liu, S. Laser powder bed fusion of high-entropy alloy particle-reinforced stainless steel with enhanced strength, ductility, and corrosion resistance. Mater. Des. 2021, 209, 109950. [Google Scholar] [CrossRef]
  8. Gu, B.; Chu, J.; Wang, Y.; Xu, G.; Gao, L.; Yang, Y. Effects of ultrasonic impact treatment on the corrosion resistance of laser-cladded CrMnFeCoNi high-entropy alloy coatings. Surf. Coat. Technol. 2024, 489, 131102. [Google Scholar] [CrossRef]
  9. Wilson, J.M.; Piya, C.; Shin, Y.C.; Zhao, F.; Ramani, K. Remanufacturing of turbine blades by laser direct deposition with its energy and environmental impact analysis. J. Clean. Prod. 2014, 80, 170–178. [Google Scholar] [CrossRef]
  10. Guo, C.; Zhou, Y.; Xu, Z.; Li, Y.; Li, G.; Shi, Z. Macro Co-Al-W-based superalloy single-crystal fabrication using selective laser melting. Scr. Mater. 2024, 251, 116213. [Google Scholar] [CrossRef]
  11. Yadroitsev, I.; Krakhmalev, P.; Yadroitsava, I. Selective laser melting of Ti6Al4V alloy for biomedical applications: Temperature monitoring and microstructural evolution. J. Alloys Compd. 2014, 583, 404–409. [Google Scholar] [CrossRef]
  12. Zhang, S.; Sun, Y.; Cheng, W.; Chen, Y.; Gu, J. High temperature oxidation behavior of CoCrFeNiMo0.2 high-entropy alloy coatings produced by laser cladding. Mater. Today Commun. 2024, 39, 108639. [Google Scholar] [CrossRef]
  13. Zia Ullah, A.; Muhammad Yasir, K.; Ehtsham ur, R. Laser-aided additive manufacturing of high entropy alloys: Processes, properties, and emerging applications. J. Manuf. Proc. 2022, 78, 131–171. [Google Scholar]
  14. Duchaniya, R.K.; Pandel, U.; Rao, P. Coatings based on high entropy alloys: An overview. Mater. Today Proc. 2021, 44, 4467–4473. [Google Scholar] [CrossRef]
  15. Lu, K.; Zhu, J.; Ge, W.; Hui, X. Progress on New Preparation Methods, Microstructures, and Protective Properties of High-Entropy Alloy Coatings. Coatings 2022, 12, 1472. [Google Scholar] [CrossRef]
  16. Meng, Y.; Yang, Z.; Shi, Y.; Liu, X.; Wang, L.; Zhang, Q. Microstructural evolution and enhanced properties of multi-layer TiZrNbCrCo high entropy alloy coatings laser-clad onto Ti–6Al–4V alloy. Intermetallics 2024, 173, 108407. [Google Scholar] [CrossRef]
  17. Yu, B.; Ren, Y.; Zeng, Y.; Ma, W.; Morita, K.; Zhan, S. Recent progress in high-entropy alloys: A focused review of preparation processes and properties. J. Mater. Res. Technol. 2024, 29, 2689–2719. [Google Scholar] [CrossRef]
  18. Ping, Z.; Ziwei, L.; Haomin, L.; Yuanbin, Z.; Huaixue, L.; Chuanwei, S. Recent progress on the microstructure and properties of high entropy alloy coatings prepared by laser processing technology: A review. J. Manuf. Process. 2022, 76, 397–411. [Google Scholar]
  19. Lian, G.; Xiao, S.; Zhang, Y.; Jiang, J.; Zhan, Y. Multi-objective optimization of coating properties and cladding efficiency in 316L/WC composite laser cladding based on grey relational analysis. Int. J. Adv. Manuf. Technol. 2021, 112, 1449–1459. [Google Scholar] [CrossRef]
  20. Gao, J.; Han, Q.; Soe, S.; Wang, L.; Zhang, Z.; Zhang, H. Laser additive manufacturing of TiB2-modified Cu15Ni8Sn/GH3230 heterogeneous materials: Processability, interfacial microstructure and mechanical performance. Mater. Sci. Eng. A 2024, 900, 146496. [Google Scholar] [CrossRef]
  21. Patil, V.V.; Mohanty, P.C.; Prashanth, G.K. Parametric Optimization of Selective Laser Melted 13Ni400 Maraging Steel by Taguchi Method. J. Manuf. Mater. Process. 2024, 8, 52. [Google Scholar] [CrossRef]
  22. Zhao, Y.; Wu, Y.; Liu, Y.; Zhang, Z.; Che, X.; Chen, H. Effect of Mn alloying on the microstructure and micro-shear properties of laser direct energy deposition Stellite21 cobalt-based coating and its process research. Surf. Coat. Technol. 2024, 483, 130786. [Google Scholar] [CrossRef]
  23. Javidrad, H.; Aydin, H.; Karakaş, B.; Alptekin, S.; Kahraman, A.S.; Koc, B. Process parameter optimization for laser powder directed energy deposition of Inconel 738LC. Opt. Laser Technol. 2024, 176, 110940. [Google Scholar] [CrossRef]
  24. Barekat, M.; Lashani Zand, M.; Shoja Razavi, R.; Erfanmanesh, M.; Ilanlou, M.; Reza Borhani, M. Experimental and numerical investigation of the effect of process parameters on crack formation and residual stresses in the laser coating process of stellite 6 alloy on X19CrMoNbVN11-1 steel substrate. Opt. Laser Technol. 2024, 176, 111038. [Google Scholar] [CrossRef]
  25. Khaleghi, S.; Azdast, T.; Doniavi, A.; Hasanzadeh, R. Performance optimization of acrylonitrile butadiene styrene/thermoplastic polyurethane composite foams blown with carbon dioxide using Taguchi technique. J. Appl. Polym. Sci. 2023, 141, e54996. [Google Scholar] [CrossRef]
  26. Udaya Prakash, J.; Ananth, S.; Jebarose Juliyana, S.; Cep, R.; Khedkar, N.; Salunkhe, S. Parametric optimization of wear parameters of hybrid composites (LM6/Bsub4/subC/fly ash) using Taguchi technique. Front. Mech. Eng. 2023, 9, 1279481. [Google Scholar] [CrossRef]
  27. Tai, Z.; Yang, Y.; Lv, J.; Zhou, H.; Trofimov, V.; Wei, Y. In-situ synthesis of high-performance TiN/TC4 sandwich structures via nitrogen-controlled laser directed energy deposition. Mater. Sci. Eng. A 2024, 896, 146231. [Google Scholar] [CrossRef]
  28. Thanumoorthy, R.S.; Sekar, P.; Bontha, S.; Balan, A.S.S. A study on the effect of process parameters and scan strategies on microstructure and mechanical properties of laser directed energy deposited IN718. J. Mater. Process. Tech. 2023, 319, 118096. [Google Scholar] [CrossRef]
  29. Sachin, A.; Chaitanya, V.; Bhargavi, A.; Ramesh, S. Modeling of Deposition Geometry in Laser Directed Energy Deposition over Inclined Surfaces for Restoration and Remanufacturing. Trans. Indian Natl. Acad. Eng. 2021, 6, 1057–1069. [Google Scholar]
  30. Cheng, F.; Wang, H.; Li, Z.; Cheng, X.; Zheng, D.; Zhang, S.; Hu, X.; Zhang, H.; Liu, M. Dynamic compression deformation behavior of laser directed energy deposited α + β duplex titanium alloy with basket-weave morphology. Addit. Manuf. 2023, 61, 103336. [Google Scholar] [CrossRef]
  31. Ilo, S.; Just, C.; Xhiku, F. Optimisation of multiple quality characteristics of hardfacing using grey-based Taguchi method. Mater. Des. 2012, 33, 459–468. [Google Scholar] [CrossRef]
  32. Yusuf, S.; Yakup, T. Experimental and statistical investigation of mechanical properties and surface roughness in additive manufacturing with selective laser melting of AlSi10Mg alloy. J. Braz. Soc. Mech. Sci. Eng. 2023, 45, 515. [Google Scholar]
  33. Cheng, K.-C.; Chen, J.-H.; Stadler, S.; Chen, S.-H. Properties of atomized AlCoCrFeNi high-entropy alloy powders and their phase-adjustable coatings prepared via plasma spray process. Appl. Surf. Sci. 2019, 478, 478–486. [Google Scholar] [CrossRef]
  34. Nagase, T.; Takemura, M.; Matsumuro, M.; Maruyama, T. Solidification microstructure of AlCoCrFeNi2.1 eutectic high entropy alloy ingots. Mater. Trans. 2018, 59, 255–264. [Google Scholar] [CrossRef]
  35. Jiang, H.; Han, K.; Li, D.; Cao, Z. Microstructure and Mechanical Properties Investigation of the CoCrFeNiNbx High Entropy Alloy Coatings. Crystals 2018, 8, 409. [Google Scholar] [CrossRef]
  36. Chu, M.; Xiao, H.; Ren, L.; Mo, T.; Lin, B. Microstructure and properties of Ti–Al–C composite coatings prepared by laser cladding. Ceram. Int. 2024, 50, 12498–12509. [Google Scholar] [CrossRef]
  37. Cui, Y.-H.; Zhang, Q.; Shao, Y.-X.; Yang, Y.; Ma, Y.-D.; Sun, W.-W. Microstructure and properties of in-situ ZrBsub2/sub-ZrC composite coatings by plasma spraying. Surf. Coat. Technol. 2021, 409, 126846. [Google Scholar] [CrossRef]
  38. Liang, J.; Li, G.; Ding, X.; Li, Y.; Wen, Z.; Zhang, T.; Qu, Y. Effect of C14 Laves/BCC on microstructure and hydrogen storage properties of (Ti32.5V27.5Zr7.5Nb32.5)1−xFex (x = 0.03, 0.06, 0.09) high entropy hydrogen storage alloys. J. Energy Storage 2023, 73, 108852. [Google Scholar] [CrossRef]
  39. Cui, P.; Ma, Y.; Zhang, L.; Zhang, M.; Fan, J.; Dong, W. Microstructure and mechanical behaviors of CoFeNiMnTixAl1−x high entropy alloys. Mater. Sci. Eng. A 2018, 731, 124–130. [Google Scholar] [CrossRef]
  40. Dong, Z.; Feng, L.; Long, H.; Lu, B.; Zhu, J.; Yan, X. A multi-objective optimization of laser cladding processing parameters of AlCoCrFeNi2.1 eutectic high-entropy alloy coating. Opt. Laser Technol. 2024, 170, 110302. [Google Scholar] [CrossRef]
  41. Li, B.; Zhang, L.; Xu, Y.; Liu, Z.; Qian, B.; Xuan, F. Selective laser melting of CoCrFeNiMn high entropy alloy powder modified with nano-TiN particles for additive manufacturing and strength enhancement: Process, particle behavior and effects. Powder Technol. 2020, 360, 509–521. [Google Scholar] [CrossRef]
  42. Gnanasekaran, S.; Padmanaban, G.; Balasubramanian, V.; Kumar, H.; Albert, S.K. Optimizing the laser parameters to attain maximum hardness in nickel based hardfacing surfaces. J. Mech. Behav. Mater. 2017, 26, 113–125. [Google Scholar] [CrossRef]
  43. Liu, Y.; Xu, T.; Li, G.; Li, Z.; Du, F.; Chen, G. A multi-objective optimization of laser cladding process of Ni-based coating on the preheated copper alloy. Mater. Today Commun. 2023, 35, 105614. [Google Scholar] [CrossRef]
  44. Nie, M.H.; Zhang, S.; Wang, Z.Y.; Zhang, C.H.; Chen, H.T.; Chen, J. Effect of laser power on microstructure and interfacial bonding strength of laser cladding 17-4PH stainless steel coatings. Mater. Chem. Phys. 2022, 275, 125236. [Google Scholar] [CrossRef]
  45. Yang, M.; Lu, J.; Wang, H.; Liu, Y.; Yang, H.; Zhang, W.; Tan, J.; Chen, J. Effect of the laser power on the microstructure and mechanical properties of the laser-MIG hybrid welding joints of the 2195 Al–Li alloy. Sci. Technol. Weld. Join. 2020, 26, 75–83. [Google Scholar] [CrossRef]
  46. Ratanapongpien, K.; Khantachawana, A.; Kondoh, K. Effect of Laser Scanning Speed and Fine Shot Peening on Pore Characteristics, Hardness, and Residual Stress of Ti-6Al-4V Fabricated by Laser Powder Bed Fusion. Metals 2024, 14, 250. [Google Scholar] [CrossRef]
  47. Ni, C.; Shi, Y.; Liu, J. Effects of inclination angle on surface roughness and corrosion properties of selective laser melted 316L stainless steel. Mater. Res. Express 2019, 6, 036505. [Google Scholar] [CrossRef]
  48. Ma, M.; Xiong, W.; Lian, Y.; Han, D.; Zhao, C.; Zhang, J. Modeling and optimization for laser cladding via multi-objective quantum-behaved particle swarm optimization algorithm. Surf. Coat. Technol. 2020, 381, 125129. [Google Scholar] [CrossRef]
  49. Rafiei, J.; Ghasemi, R.A. Development of thermo-mechanical simulation of WC/Inconel 625 metal matrix composites laser cladding and optimization of process parameters. Int. J. Therm. Sci. 2024, 198, 108883. [Google Scholar] [CrossRef]
  50. Zhang, H.; Pan, Y.; Zhang, Y.; Lian, G.; Cao, Q.; Yang, J. Influence of laser power on the microstructure and properties of in-situ NbC/WCoB–TiC coating by laser cladding. Mater. Chem. Phys. 2022, 290, 126636. [Google Scholar] [CrossRef]
  51. Shi, B.; Mu, X.; Zhan, H.; Deng, L.; Li, T.; Zhang, H. Crack Behavior of Ni60A Coating Prepared by Laser Cladding on a Tilted Substrate. Coatings 2022, 12, 966. [Google Scholar] [CrossRef]
  52. Li, Y.; Su, K.; Bai, P.; Wu, L. Microstructure and property characterization of Ti/TiBCN reinforced Ti based composite coatings fabricated by laser cladding with different scanning speed. Mater. Charact. 2020, 159, 110023. [Google Scholar] [CrossRef]
  53. Wang, C.; Yu, J.; Zhang, Y.; Yu, Y. Phase evolution and solidification cracking sensibility in laser remelting treatment of the plasma-sprayed CrMnFeCoNi high entropy alloy coating. Mater. Des. 2019, 182, 108040. [Google Scholar] [CrossRef]
  54. Al-Ruqeishi, M.S.; Mohiuddin, T.; Al-Amri, K.; Rohman, N. Graphene Surface Energy by Contact Angle Measurements. Arab. J. Sci. Eng. 2022, 48, 757–762. [Google Scholar] [CrossRef]
  55. Segura-Cardenas, E.; Ramirez-Cedillo, E.; Sandoval-Robles, J.; Ruiz-Huerta, L.; Caballero-Ruiz, A.; Siller, H. Permeability Study of Austenitic Stainless Steel Surfaces Produced by Selective Laser Melting. Metals 2017, 7, 521. [Google Scholar] [CrossRef]
  56. Zhao, S.; Yang, L.; Huang, Y.; Xu, S. Feasibility study of producing Ni/WC composite coatings by laser deposition with a tubular cored wire. J. Mater. Process. Tech. 2020, 285, 116745. [Google Scholar] [CrossRef]
  57. Luo, W.; Liu, Y.; Tu, C. Wetting behaviors and interfacial characteristics of molten AlxCoCrCuFeNi high-entropy alloys on a WC substrate. J. Mater. Sci. Technol. 2021, 78, 192–201. [Google Scholar] [CrossRef]
  58. Zhang, Y.; Han, T.-F.; Xiao, M.; Shen, Y.-F. Effect of process parameters on the microstructure and properties of laser-clad FeNiCoCrTi 0.5 high-entropy alloy coating. Int. J. Miner. Metall. Mater. 2020, 27, 630–639. [Google Scholar] [CrossRef]
  59. Ding, Y.; Gui, W.; Nie, B.; Bi, W.; Zhong, C.; Xue, Y.; Luan, B. Elimination of elemental segregation by high-speed laser remelting for ultra-high-speed laser cladding Inconel 625 coatings. J. Mater. Res. Technol. 2023, 24, 4118–4129. [Google Scholar] [CrossRef]
  60. Gao, Q.; Ling, W.; Zhou, Y.; Zhan, X.; Shen, H. Correlation between special bud-shaped area and molten pool flow in laser cladding high-entropy alloy on TC4 surface. Int. J. Therm. Sci. 2023, 183, 107825. [Google Scholar] [CrossRef]
  61. Sun, Z.; Chen, L.; Chen, X.; Meng, F.; Yu, T.; Zhao, J. Analysis and prediction of Cu-Sn-Ti alloy deposited on 316 L steel by coaxial laser cladding. Optik 2023, 282, 170839. [Google Scholar] [CrossRef]
  62. Liu, L.; Wang, G.; Ren, K.; Di, Y.; Wang, L.; Rong, Y.; Wang, H. Marangoni flow patterns of molten pools in multi-pass laser cladding with added nano-CeO2. Addit. Manuf. 2022, 59, 103156. [Google Scholar] [CrossRef]
  63. Deng, D.; Li, T.; Huang, Z.; Jiang, H.; Yang, S.; Zhang, Y. Multi-response optimization of laser cladding for TiC particle reinforced Fe matrix composite based on Taguchi method and grey relational analysis. Opt. Laser Technol. 2022, 153, 108259. [Google Scholar] [CrossRef]
  64. Yu, T.; Yang, L.; Zhao, Y.; Sun, J.; Li, B. Experimental research and multi-response multi-parameter optimization of laser cladding Fe313. Opt. Laser Technol. 2018, 108, 321–332. [Google Scholar] [CrossRef]
  65. Lian, G.; Yue, K.; Liu, Z.; Lu, H.; Wang, Y. Influences of the cladding method and overlapping rate on the multi-channel overlapping of composite coatings. Optik 2023, 283, 170896. [Google Scholar] [CrossRef]
  66. Agapovichev, A.V.; Khaimovich, A.I.; Smelov, V.G.; Kokareva, V.V.; Zemlyakov, E.V.; Babkin, K.D.; Kovchik, A.Y. Multiresponse Optimization of Selective Laser Melting Parameters for the Ni-Cr-Al-Ti-Based Superalloy Using Gray Relational Analysis. Materials 2023, 16, 2088. [Google Scholar] [CrossRef]
  67. Yue, H.; Lv, N.; Guo, C.; Zhai, J.; Dai, W.; Zhang, J. Multi-Objective Process Optimization of Laser Cladding Co-Based Alloy by Process Window and Grey Relational Analysis. Coatings 2023, 13, 1090. [Google Scholar] [CrossRef]
  68. Sneha, S.; Jaromír, K.; Petr, Š. Interfacial Adhesion of Thick NiTi Coating on Substrate Stainless Steel. Materials 2022, 15, 8598. [Google Scholar] [CrossRef]
  69. Li, B.; Wang, B.; Zhu, G.; Zhang, L.; Lu, B. Low-Roughness-Surface Additive Manufacturing of Metal-Wire Feeding with Small Power. Materials 2021, 14, 4265. [Google Scholar] [CrossRef]
  70. Meghwal, A.; Anupam, A.; Murty, B.S.; Berndt, C.C.; Kottada, R.S.; Ang, A.S.M. Thermal Spray High-Entropy Alloy Coatings: A Review. J. Therm. Spray Technol. 2020, 29, 857–893. [Google Scholar] [CrossRef]
  71. Santosh, K. Comprehensive review on high entropy alloy-based coating. Surf. Coat. Technol. 2024, 477, 130327. [Google Scholar]
  72. Wang, L.; Ran, X.; Li, Y.; Li, F.; Liu, J.; Du, J. Energy consumption model of plasma spraying based on unit process life cycle inventory. J. Mater. Res. Technol. 2020, 9, 15324–15334. [Google Scholar] [CrossRef]
  73. Minx, J.C.; Lamb, W.F.; Andrew, R.M.; Canadell, J.G.; Crippa, M.; Döbbeling, N.; Forster, P.M.; Guizzardi, D.; Olivier, J.; Peters, G.P.; et al. A comprehensive and synthetic dataset for global, regional, and national greenhouse gas emissions by sector 1970–2018 with an extension to 2019. Earth Syst. Sci. Data 2021, 13, 5213–5252. [Google Scholar] [CrossRef]
  74. Koide, R.; Lettenmeier, M.; Akenji, L.; Toivio, V.; Amellina, A.; Khodke, A.; Watabe, A.; Kojima, S. Lifestyle carbon footprints and changes in lifestyles to limit global warming to 1.5 °C, and ways forward for related research. Sustain. Sci. 2021, 16, 2087–2099. [Google Scholar] [CrossRef]
  75. Jack, T.; Bååth, J.; Heinonen, J.T.; Gram-Hanssen, K. Correction: How individuals make sense of their climate impacts in the capitalocene: Mixed methods insights from calculating carbon footprints. Sustain. Sci. 2024, 19, 1117. [Google Scholar] [CrossRef]
  76. Dominique, M.V.; Christos, Z. “Take your responsibility”: The politics of green sacrifice for just low-carbon transitions in rural Portugal. Sustain. Sci. 2024, 19, 1313–1326. [Google Scholar]
Figure 1. SEM of FeCuNiCrAl.
Figure 1. SEM of FeCuNiCrAl.
Coatings 14 01103 g001
Figure 2. Laser cladding system.
Figure 2. Laser cladding system.
Coatings 14 01103 g002
Figure 3. Cross-sectional geometry of the laser cladding layer.
Figure 3. Cross-sectional geometry of the laser cladding layer.
Coatings 14 01103 g003
Figure 4. X-ray diffraction of the coatings from the single-factor experiment on HEA addition rate.
Figure 4. X-ray diffraction of the coatings from the single-factor experiment on HEA addition rate.
Coatings 14 01103 g004
Figure 5. Main effects plot for the S/N analysis of microhardness.
Figure 5. Main effects plot for the S/N analysis of microhardness.
Coatings 14 01103 g005
Figure 6. Main effects plot for the S/N analysis of dilution rate.
Figure 6. Main effects plot for the S/N analysis of dilution rate.
Coatings 14 01103 g006
Figure 7. Main effects plot for the S/N analysis of average contact angle.
Figure 7. Main effects plot for the S/N analysis of average contact angle.
Coatings 14 01103 g007
Figure 8. Main effects plot for the S/N analysis of mean difference of contact angles.
Figure 8. Main effects plot for the S/N analysis of mean difference of contact angles.
Coatings 14 01103 g008
Figure 9. Comparison of cross-sectional microscopic morphology of laser cladding coatings between the optimal parameter set from the orthogonal design (a) and the optimal set from GRA (b).
Figure 9. Comparison of cross-sectional microscopic morphology of laser cladding coatings between the optimal parameter set from the orthogonal design (a) and the optimal set from GRA (b).
Coatings 14 01103 g009
Table 1. Chemical composition of Q235 substrate (wt.%).
Table 1. Chemical composition of Q235 substrate (wt.%).
CSiMnSPFe
≤0.22≤0.35≤1.4≤0.050≤0.045Bal.
Table 2. Chemical composition of FeCuNiCrAl HEA (wt.%).
Table 2. Chemical composition of FeCuNiCrAl HEA (wt.%).
FeNiCrAlCu
22.1522.4119.7210.39Bal.
Table 3. Chemical composition of 316L stainless-steel (wt.%).
Table 3. Chemical composition of 316L stainless-steel (wt.%).
CSiMnSPCrNiMoFe
<0.03<1.00<2.00<0.03<0.045<18<14<3Bal.
Table 4. Orthogonal experimental design.
Table 4. Orthogonal experimental design.
Processing
Parameter
NotationUnitLevels
1234
HEA Addition RateA%051015
Laser PowerBW750110014501800
Scanning SpeedCmm/s57911
Powder Feed RateDr/min25811
Substrate Tilt AngleE°0102030
Table 5. Orthogonal experimental design parameters and responses.
Table 5. Orthogonal experimental design parameters and responses.
RunA
(%)
B
(W)
C
(mm/s)
D
(r/min)
E
(°)
Parameter
Combination
H
(HV)
λ
(%)
a ¯
(°)
Δa
(°)
10750520A1B1C1D1E1439.8738.3224.260.634
2011007510A1B2C2D2E2181.6015.9251.850.600
3014509820A1B3C3D3E3364.0927.2457.7626.464
401800111130A1B4C4D4E4194.1318.6563.274.236
557507830A2B1C2D3E4190.0520.3125.213.076
65110051120A2B2C1D4E3184.728.3976.1517.584
75145011210A2B3C4D1E2460.2883.3239.4220.054
851800950A2B4C3D2E1457.0385.5090.8148.533
91075091110A3B1C3D4E2203.306.5559.061.460
101011001180A3B2C4D3E1257.8321.9352.391.029
111014505530A3B3C1D2E4407.1336.0733.942.513
121018007220A3B4C2D1E3424.4786.2717.954.194
131575011520A4B1C4D2E3461.3031.2625.790.312
141511009230A4B2C3D1E4445.0279.5312.672.010
151514507110A4B3C2D4E1180.0414.6473.755.192
161518005810A4B4C1D3E2209.8125.1555.202.743
Table 6. Single-factor experiment of high-entropy alloy (HEA) addition rate.
Table 6. Single-factor experiment of high-entropy alloy (HEA) addition rate.
Processing
Parameter
NotationUnitLevels
1234
HEA Addition RateA%051015
Laser PowerBW750750750750
Scanning SpeedCmm/s7777
Powder Feed RateDr/min2222
Substrate Tilt AngleE°30303030
Table 7. Lattice constants of phases in the laser cladding coatings at different HEA addition rates.
Table 7. Lattice constants of phases in the laser cladding coatings at different HEA addition rates.
CriterionPhaseHEA Addition Rate
0%5%10%15%
Lattice
constant
(nm)
Laves0.48910.48810.48850.4885
FCC0.41980.42060.42030.4203
BCC0.28830.28860.28880.2885
Table 8. Signal-to-noise ratio for microhardness.
Table 8. Signal-to-noise ratio for microhardness.
LevelABCDE
148.7649.4749.2152.9149.85
249.3447.9347.1150.9647.76
349.7950.4550.8947.8750.60
449.4549.4950.1345.5949.13
Delta1.032.523.787.322.84
Rank54213
Table 9. Signal-to-noise ratio for dilution rate.
Table 9. Signal-to-noise ratio for dilution rate.
LevelABCDE
112.5413.9912.683.309.89
29.5813.1611.959.0713.30
311.759.619.5812.5811.05
410.197.319.8619.129.82
Delta2.976.683.1015.823.48
Rank52413
Table 10. Signal-to-noise ratio for average contact angle.
Table 10. Signal-to-noise ratio for average contact angle.
LevelABCDE
1−33.31−29.85−32.7−26.69−34.65
2−34.19−32.09−31.19−33.07−34.12
3−31.38−33.78−32.97−33.12−31.54
4−30.62−33.78−32.64−36.61−29.18
Delta3.573.931.789.925.47
Rank43512
Table 11. Signal-to-noise ratio for mean difference of contact angles.
Table 11. Signal-to-noise ratio for mean difference of contact angles.
LevelABCDE
1−8.150.26−9.43−10.15−11.08
2−23.61−6.69−8.02−6.79−8.41
3−6.00−19.20−17.88−11.81−13.92
4−4.75−16.87−7.17−13.76−9.09
Delta18.8619.4610.716.975.51
Rank21345
Table 12. Normalized experimental data and corresponding deviations.
Table 12. Normalized experimental data and corresponding deviations.
RunY(H)0(H)Y(λ)0(λ) Y ( a ¯ )0 ( a ¯ ) Y ( Δ a )0 ( Δ a )
10.923810.076190.602260.397740.851700.148300.993320.00668
20.005550.994450.883310.116690.498640.501360.994030.00597
30.654380.345620.741530.258470.423010.577000.457660.54234
40.050100.949900.848180.151820.352470.647540.918630.08138
50.035590.964410.828110.171900.839600.160400.942680.05732
60.016640.983360.977420.022590.187610.812390.641820.35818
70.996370.003630.037640.962360.657750.342250.590590.40941
80.984820.015180.010040.989960101
90.082700.91730100.406300.593700.976190.02381
100.276580.723420.808030.191970.491710.508290.985130.01487
110.807400.192600.629860.370140.727830.272180.954360.04564
120.869050.13095010.932460.067550.919500.08050
13100.690090.309910.832190.1678110
140.942120.057880.085320.91468100.964790.03521
15010.899620.100380.218410.781590.898800.10120
160.105850.894160.766630.233380.455730.544270.949590.05041
Table 13. GRCs and GRG for each experimental run.
Table 13. GRCs and GRG for each experimental run.
RunGRC(H)GRC(λ) GRC ( a ¯ ) GRC ( Δ a )GRGRank
10.867770.556950.771250.986820.808832
20.334570.810780.499320.988200.706908
30.591280.659220.464260.479690.5425115
40.344850.767080.435720.860030.6411911
50.341430.744160.757110.897150.719207
60.337070.956780.380980.582630.5820414
70.992800.341910.593650.549810.5948313
80.970530.335580.333330.333330.4585416
90.352781.000000.457170.954550.736664
100.408690.722580.495890.971120.693129
110.721920.574620.647520.916350.733505
120.792460.333330.880990.861320.721566
131.000000.617350.748711.000000.848991
140.896250.353441.000000.934210.797753
150.333330.832810.390140.831670.6355612
160.358640.681780.478800.908410.6485310
Table 14. Mean of S/N (GRG) for each processing parameter.
Table 14. Mean of S/N (GRG) for each processing parameter.
Processing ParameterNotationLevelsAbsolute Value DifferenceRank
1234
HEA Addition RateA−3.51−4.71−2.84−2.771.942
Laser PowerB−2.20−3.22−4.11−4.312.111
Scanning SpeedC−3.25−3.16−4.18−3.251.014
Powder Feed RateD−2.79−3.48−3.78−3.7915
Substrate Tilt AngleE−3.93−3.49−3.57−2.841.093
Table 15. Comparison of optimization and experimental validation results.
Table 15. Comparison of optimization and experimental validation results.
OutputBest Parameter Set fromOrthogonal DesignGRA PredictionValidation on GRA Prediction
Parameter SetA4B1C4D2E3A4B1C2D1E4A4B1C2D1E4
H461.30 HV-549.14 HV
λ0.31-0.536
a ¯ 25.79°-16.16°
Δa0.312-1.69
GRG0.848990.943160.93427
Table 16. Carbon emissions associated with the production of 1000 composite coating samples in selected countries.
Table 16. Carbon emissions associated with the production of 1000 composite coating samples in selected countries.
Technical CategoriesChina/kgThe United Kingdom/kgFrance/kgThe United States/kgThe Russian
Federation/kg
Laser
cladding
161.5961.2111.88108.2193.32
Plasma spray239.6090.7617.62160.45138.37
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

Gao, S.; Fu, Q.; Li, M.; Huang, L.; Liu, N.; Cui, C.; Yang, B.; Zhang, G. Optimization of Laser Cladding Parameters for High-Entropy Alloy-Reinforced 316L Stainless-Steel via Grey Relational Analysis. Coatings 2024, 14, 1103. https://doi.org/10.3390/coatings14091103

AMA Style

Gao S, Fu Q, Li M, Huang L, Liu N, Cui C, Yang B, Zhang G. Optimization of Laser Cladding Parameters for High-Entropy Alloy-Reinforced 316L Stainless-Steel via Grey Relational Analysis. Coatings. 2024; 14(9):1103. https://doi.org/10.3390/coatings14091103

Chicago/Turabian Style

Gao, Senao, Qiang Fu, Mengzhao Li, Long Huang, Nian Liu, Chang Cui, Bing Yang, and Guodong Zhang. 2024. "Optimization of Laser Cladding Parameters for High-Entropy Alloy-Reinforced 316L Stainless-Steel via Grey Relational Analysis" Coatings 14, no. 9: 1103. https://doi.org/10.3390/coatings14091103

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

Article Metrics

Back to TopTop