Next Article in Journal
A Review on Preparation of Palladium Oxide Films
Previous Article in Journal
Surface Tribological Properties Enhancement Using Multivariate Linear Regression Optimization of Surface Micro-Texture
Previous Article in Special Issue
Aerosol-Deposited 8YSZ Coating for Thermal Shielding of 3YSZ/CNT Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure and Mechanical Properties of FeCoCrNiAl + WC Composite Coating Formed by Laser Cladding on H13

1
Department of Mechanical Engineering, Northeast Electric Power University, No. 169 Changchun Road, Jilin 132012, China
2
School of Mechanical Electrical Engineering, Jilin Institute of Chemical Technology, Jilin 132022, China
*
Author to whom correspondence should be addressed.
Coatings 2024, 14(10), 1259; https://doi.org/10.3390/coatings14101259
Submission received: 4 September 2024 / Revised: 28 September 2024 / Accepted: 30 September 2024 / Published: 1 October 2024

Abstract

:
To enhance high-temperature wear resistance of H13 steel, laser cladding was used to prepare a high-entropy alloy + carbide composite coating. The microstructure and high-temperature wear resistance of the composite coating were systematically analyzed. The results indicate that the FeCoCrNiAl + WC composite coating had a phase structure of BCC + FCC solid solutions, with a small amount of CFe15.1. The microstructure of the composite coating consisted of columnar and equiaxed grains. The microhardness of the FeCoCrNiAl + WC composite coatings was approximately 3.0–3.4 times that of H13. At wear temperatures of 823 K, compared with H13 steel, the wear volumes of composite coatings with different WC contents were reduced by 73.4%–80.2%. Among these, the FeCoCrNiAl + 10% WC composite coating showed the lowest wear volume. Furthermore, when wear temperatures increased from 623 K to 823 K, compared with H13 steel (108.37%), the increase in the wear volume of the FeCoCrNiAl + 10% WC coating was reduced to 90.82%, which indicates the FeCoCrNiAl + 10% WC coating had better high-temperature wear resistance. The wear mechanisms of the composite coating were abrasive and oxidative wear, while H13 steel exhibited abrasive wear, oxidative wear and fatigue wear.

1. Introduction

H13 hot work tool steel is commonly used for manufacturing dies that experience significant impact loads, such as forging dies, hot extrusion dies, precision forging dies, and die-casting molds for aluminum, copper, and their alloys. It has excellent hardenability, toughness, and crack resistance, with moderate high-temperature wear resistance. However, during high-temperature processing, the hardness of H13 steel decreases rapidly, reducing the service life of the molds. To improve the surface properties of H13 steel and extend the service life of the molds, researchers have applied various surface treatment techniques for its enhancement [1,2,3].
Wu et al. [4] used thermal spraying technology to prepare a Cr3C2-NiCr + 10%Ni/MoS2 composite coating on H13 steel. The study found that under high-temperature conditions, the lubricating effect of MoS2 and the protective effect of the oxide layer effectively improved the high-temperature wear resistance of the substrate surface. Chiu et al. [5] applied electroplating technology to produce a hard Cr coating to improve the surface properties of H13, finding that the hardness of hard Cr coating reached 890HV0.01, providing good long-term wear resistance. Although traditional surface treatment techniques can enhance the surface properties of H13 steel, they have certain limitations, such as low material utilization rates, as well as poor coating adhesion for thermal spraying technology, and environmental pollution for electroplating technology.
Compared to traditional surface treatment techniques, laser cladding technology offers advantages such as rapid melting and cooling rates, a small heat-affected zone, and high material utilization rates. This technology can produce dense coatings with excellent properties, making it widely used in the surface strengthening of H13 steel [6,7,8]. Since its introduction, the concept of high-entropy alloys (HEAs) has gained widespread attention due to their outstanding properties. The four major effects of high-entropy alloys (high-entropy effect, sluggish diffusion effect, lattice distortion effect, and cocktail effect) endow them with excellent comprehensive properties. Additionally, the design concept allows for a wide variety of element combinations, enabling the design of high-entropy alloys with different properties for various applications, making them an important choice for laser cladding coating materials. Researchers have conducted extensive studies on laser cladding high-entropy alloy coatings on H13 steel [9,10,11]. Shu et al. [12] designed a CoCrFeNiBSi amorphous high-entropy alloy coating, which primarily consisted of FeNi3, α-Co, and Cr2C. At a line energy of 140 J/mm, the average hardness of the coating exceeded 1000HV0.2, but the high-temperature wear resistance decreased with increasing line energy. Wang et al. [13] prepared an Al0.1CoCrFeNi coating on H13, which demonstrated better wear resistance than the substrate under different loads, with adhesive wear and slight oxidative wear being the main wear mechanisms. Shi et al. [14] studied the microstructure and properties of NiCoCrMnFe coatings, finding that the coating consisted of a single FCC phase, with a hardness of 500HV0.1, and a wear rate reduced by 63.2% compared to the substrate.
Many studies have shown that the addition of ceramic phases improves the performance of coatings [15,16,17]. Shrey Bhatnagar et al. [18] studied the effect of TiC particles with different grain sizes on Inconel 625 coatings, and found that coarse-grained TiC particles increased the laser absorption rate, leading to higher dilution rates and thus poorer coating performance. Zhang et al. [19] reinforced Ni60 coatings with graphene and NbC, finding that the addition of graphene improved the flow of the molten pool, promoted the dissolution and re-solidification of NbC, and increased the carbon content combined with chromium to form new hard phases, creating more nucleation points and refining the grains, with the coating hardness reaching up to 1048HV2. Hao et al. [20] used Si3N4 to reinforce MoNbTaWTi high-entropy alloy coatings. The decomposition of Si3N4 resulted in a solid solution of Si in the BCC phase and the formation of a TiN hard phase with Ti, improving the performance of the coating. The coating with 2 wt.% Si3N4 had a maximum microhardness of 892HV0.1 and a wear volume of only 22.35% of the substrate.
WC has the characteristics of high melting point and high hardness, and is often used in the manufacturing of cemented carbide and cutting tools. In recent years, WC has been widely studied as a strengthening phase in coatings to improve the wear resistance of coatings [21,22,23,24,25]. Wu et al. [26] designed and prepared a CoCrFeNiMo0.2Nb0.2/WC composite coating, and studied the influence of WC content on the high-temperature wear resistance of the coating. When the WC content was 60%, the coating had the lowest wear rate, and the wear mechanisms were mainly adhesive wear, abrasive wear and oxidative wear. Karmakar et al. [27] prepared Stellite 6 and Stellite 6 + 30% WC coatings on the surface of H13 steel, and compared the surface properties of H13 steel with laser remelting and the two coatings. After adding WC, the maximum hardness of coating reached up to 3000HV0.01, and the wear resistance of H13 steel at high temperature (450, 550, 650 °C) was significantly improved. Li et al. [28] prepared an AlCoCrFeNi2.1/WC composite coating, and found that with the increase in WC content, carbides (Cr7C3, Cr21W2C6) precipitated from the coating increased. When the WC content was 30%, the coating had the highest average microhardness of 572HV0.1 and the lowest wear volume of 5.44 mm3.
In summary, most of the current research on high-entropy alloy coatings on H13 steel has focused on wear resistance under room temperature conditions, with less discussion on high-temperature wear resistance. Research on ceramic phase-reinforced laser cladding coatings has mainly concentrated on traditional alloy coatings, with less focus on adding ceramic phases to high-entropy alloys. In this study, laser cladding technology was used to prepare the FeCoCrNiAl + WC composite coating with different WC contents (5%, 10%, 15%, 20%). The effect of WC content on the phase composition, hardness, high-temperature wear resistance and wear mechanism of the composite coating were analyzed. The study aims to provide some experimental support for the high-temperature performance of WC-enhanced FeCoCrNiAl coatings, and to provide a reference for the future development of high-temperature wear-resistant coatings on H13 die steel formed by laser cladding.

2. Materials and Methods

2.1. Experimental Materials

H13 steel was selected as the substrate material in this study, with substrate dimensions of 60 mm × 50 mm × 5 mm. The chemical composition of H13 steel is shown in Table 1. Powder materials used for the laser cladding experiment were Fe, Co, Cr, Ni, Al, and WC powders (Xingtai Xinnai Metal Materials Co., Ltd, Xingtai, China), with particle sizes of 50–80 μm and a purity of 99.9%. The particle shape was non-spherical and irregular. The weighing method for the powders was as follows: According to the mass ratio in Table 2, the Fe, Co, Cr, Ni, and Al powders were weighed using an electronic balance and mixed. The mixed FeCoCrNiAl high-entropy alloy powder was weighed and mixed with WC according to the mass ratio in Table 3. To ensure uniform mixing of the powders, a planetary ball mill was used to mix the powders thoroughly for 1 h at a speed of 240 r/min, with a ball-to-powder ratio (mass ratio) of 2:1.

2.2. Laser Cladding Experiment

Before the experiment, the surface of the H13 steel was polished and dried using 240-grit sandpaper to remove rust and oil, which improved laser absorption and enhanced the quality of the cladding layer. The powders were prepared on H13 by using a die with a size of 50 mm × 30 mm × 1.0 mm.
A DL-HL-T200 CO2 cross-flow laser (Shenyang Continental Laser Complete Equipment Co., Ltd., Shenyang, China) was used for the laser cladding experiment, with argon gas as the shielding gas. The parameters of the laser experiment are shown in Table 4. After the experiment, the sample was naturally cooled at room temperature, and was cut into 10 mm × 10 mm × 5 mm for subsequent use using wire cutting. The flow chart of laser cladding experiment is shown in Figure 1.

2.3. Microstructure and Properties Testing

To analyze the phase composition of the FeCoCrNiAl + WC coating, the surface of the sample was ground using 400# sandpaper. An X-ray diffractometer equipped with a Cu target (TD-3500X, Dandong Tongda Technology Co., Ltd., Dandong, China) was used for phase detection. The parameters were set as follows: a diffraction angle of 20–95°, a sampling time of 1 s, and a step size of 0.068°.
A Zeiss Sigma300 scanning electron microscope (SEM, Zeiss, Oberkochen, Germany) was used to observe the microstructure of the coating, and a Smart EDX energy dispersive spectrometer was used to detect the elemental distribution in the coating. The cross-section of the sample was ground using 400#–2000# sandpaper and polished, then etched by HNO3:HCl = 1:3.
A Vickers hardness tester (HXD-1000TMC/LCD, Wuxi Metes Precision Technology Co., LTD, Wuxi, China) was used to measure the hardness of the sample. On the cross-section of the sample, points were spaced 0.1 mm apart from the top of the coating to the substrate, with a load of 0.98 N and a loading time of 15 s. The average value of three points in the same layer was taken as the hardness of the coating.
An MGW-02 linear reciprocating friction and wear testing machine (Jinan Yihua Tribology Testing Technology Co., Ltd., Jinan, China) was used to conduct the high-temperature friction and wear experiment. The grinding ball was a ZrO2 ceramic ball with a diameter of 6.5 mm. Due to the rapid deterioration property of H13 steel above 813 K, in order to compare the wear resistance of the coating and substrate at different temperatures, testing temperatures of 623 K and 823 K were selected. The experimental parameters were as follows: load, 20 N; frequency, 5 HZ; and duration, 30 min. After the experiment, the wear scar size and wear morphology were measured by using SEM.

3. Results

3.1. Macroscopic Morphology of the FeCoCrNiAl + WC Composite Coating

Figure 2 shows the macroscopic morphology of the FeCoCrNiAl + WC composite coating. It can be seen that the coatings have a good surface shape with no crack defects. When the WC content is 5%, the cladding track of the coating is relatively continuous, but there are some small pits (Figure 2a). When the WC content is 10% and 15%, the coating forms a scaly cladding morphology, and the overlapping areas are smooth (Figure 2b,c). When WC content increased to 20%, a black WC ceramic phase helps the coating absorb heat, leading to heat accumulation in the coating, blurring the cladding track, and deteriorating the morphology (Figure 2d). It can be seen from Figure 2d that parts of the coating that have failed laser cladding and that the surface of the coating is extremely uneven, making it difficult to conduct performance testing experiments. In addition, the unformed coating has no practical significance. Therefore, the sample with 20% WC was not analyzed in the subsequent structure and performance studies.

3.2. Phase Analysis of FeCoCrNiAl + WC Composite Coating

The phase compositions of the FeCoCrNiAl + WC composite coatings are shown in Figure 3. It is clear that the coating without the addition of WC mainly consists of BCC and FCC phases. When WC is added, because the high mixing entropy of the composite coating inhibits the formation of intermetallic compounds, the main phase structures of the coating remain BCC and FCC solid solutions. However, due to the high energy density of the laser, the WC decomposes, and the mixing enthalpy between carbon and iron is lower than that of other elements, resulting in the formation of a small amount of CFe15.1 phase. Furthermore, as illustrated in Figure 3, the diffraction peak of CFe15.1 is strongest when the WC content is 15%. The enlarged image of the main diffraction peak of the coating is shown in Figure 3b. It can be seen that after the addition of WC, the diffraction peak of the coating shifts, which indicates that the coatings all have lattice distortion. In addition, WC is not found in the XRD results, which indicates that the WC particles were decomposed during the laser cladding process. The W atom was dissolved in solid solution as a solute, and the solid solution-strengthening effect of the FeCoCrNiAl coating was further enhanced. This finding is similar to those of Xu et al. [29]. When the WC content is 10%, the diffraction peak offset is the largest, indicating that the lattice distortion is the largest and that the solid solution-strengthening effect is the best.

3.3. Microstructure Analysis of the FeCoCrNiAl + WC Composite Coating

Figure 4 shows the microstructure of the FeCoCrNiAl + 10% WC composite coating. As shown in Figure 4a, the coating structure is dense, with a narrow fusion line formed at the interface between the coating and substrate, and no pores or cracks are observed. Figure 4b illustrates an enlarged view of the fusion line, where a 15 μm thick planar crystal is formed at the junction. Due to the reheating effect of the laser, the surface layer of the substrate at the bottom of the fusion zone (heat-affected zone) undergoes quenching heat treatment, forming the lath martensitic structure. The enlarged microstructure of the top of the cladding layer is shown in Figure 4c, where the cladding layer has a mixture of columnar and equiaxed crystals. Due to the rapid solidification and cooling characteristics of the laser, the temperature gradient at the solid–liquid interface is large at the beginning of solidification, forming fine equiaxed crystals at the solid–liquid interface. As the solid–liquid interface moves toward the molten pool, with fine equiaxed crystals as the core, the growth direction of the crystals is perpendicular to the interface. The grains have difficulty growing laterally, forming columnar crystals. Meanwhile, as the temperature in other regions of the molten pool decreases, the directionality of heat dissipation is lost, and impurities floating on the surface of the liquid phase provide nucleation sites for the formation of new phases, allowing the crystals to grow freely and form equiaxed crystals.

3.4. Hardness Analysis of the FeCoCrNiAl + WC Composite Coating

Figure 5 shows the hardness distribution curve of the FeCoCrNiAl + WC composite coating. As depicted in the figure, the hardness distribution is divided into three regions: cladding zone, bonding zone + heat-affected zone, and the substrate. The microhardness of the substrate is about 250HV0.1. In the upper layer of the coating, elemental burn-off and a loose structure lead to lower hardness. The middle and lower parts of the coating are denser, with hardness ranging from approximately 750HV0.1 to 850HV0.1, which is about 3.0–3.4 times that of the substrate.
The BCC phase formed in the coating has a relatively high hardness, which contributes to the increased hardness. Additionally, aluminum, which has the largest atomic radius among the five elements, increases the solid solution hardness. The incorporation of WC into the coating, facilitated by the high-energy laser beam, results in the decomposition of WC, with W and C atoms dissolving into the solid solution, further enhancing the solid solution strengthening effect. Moreover, Figure 5 shows that the heat-affected zone forms a martensitic structure (Figure 4b), resulting in a hardness of about 500HV0.1.

3.5. High-Temperature Friction and Wear Performance Analysis of the FeCoCrNiAl + WC Composite Coating

Due to the rapid deterioration properties of H13 steel above 813 K, in order to compare and analyze the high wear resistance of FeCoCrNiAl + WC coatings and H13 steel beyond the effective use temperature, high-temperature friction and wear experiments were conducted at 823 K. Figure 6 illustrates the friction force and coefficient curves of the FeCoCrNiAl + WC composite coatings at 823 K. Initially, the friction coefficient fluctuates significantly but stabilizes as wear progresses. The average friction coefficients of the FeCoCrNiAl + 5% WC, FeCoCrNiAl + 10% WC and FeCoCrNiAl + 15% WC coatings are 0.45, 0.5 and 0.3, respectively. The lowest friction coefficient is observed for the FeCoCrNiAl + 15% WC coating, which is attributed to the increased WC content that enhances the microhardness of the coating and resists plastic deformation and dislocation slip [30]. Furthermore, due to higher WC content, the composite coating has more of the CFe15.1 phase. The carbides are known for their excellent anti-friction properties, which also contributes to decreasing the friction coefficient of the coating with 15% WC.
Furthermore, the friction force initially increases with wear time and then stabilizes. This is because, during the initial stages of wear, the friction pair first contacts with the coating, leading to the lowest surface friction force. As the wear continues, the surface temperature rises, causing an increase in the volume of the friction pairs, which in turn, increases the friction force. However, once the friction pair temperature stabilizes, the friction force also stabilizes.
Figure 7 shows the wear volumes of FeCoCrNiAl + WC composite coatings and H13 steel. The wear volume of H13 steel is 0.946 mm3, while the composite coatings have wear volumes ranging from 0.187 mm3 to 0.251 mm3. Compared to H13 steel, the wear volumes of the composite coatings are reduced by 73.4%–80.2%. The high hardness of the BCC phase formed in the coating and a hard CFe15.1 phase generated by the addition of WC both improve the wear resistance of the coating. However, the compactness of the oxide in the worn surface reduced with the increase in the WC content, leading to easier oxidation inside the coating, resulting in a decrease in wear resistance. Therefore, the FeCoCrNiAl + 10% WC coating exhibits the smallest wear volume (0.187 mm3).
Figure 8 presents the wear morphology at 823 K for H13 steel and the composite coatings. Figure 8a shows the wear morphology of H13 steel, revealing severe surface wear, cracks, significant debris and deep grooves, indicating abrasive and fatigue wear as the primary wear mechanisms. However, the composite coatings comprise BCC and CFe15.1 with high microhardness, which prevents the extrusion between the friction ball and coating surface from causing plastic deformation and plowing effects [31]. The W atom resulting from the decomposition of WC has a larger atomic radius, which makes the degree of lattice distortion more significant. The solution-strengthening effect is improved, and the coating obtains a higher hardness. In addition, the phase structure of the composite coating is primarily FCC, in which edge dislocations are predominant. These dislocations offer superior toughness, which makes the coating difficult to detach during wear, so the wear resistance is enhanced. As a result, the wear morphology of the composite coatings with different WC contents (Figure 8b–d) shows no cracks, only small amounts of debris, and shallower grooves, with abrasive wear being the dominant wear mechanism.
Additionally, the EDS point scan analyses of the wear debris on the coatings and H13 steel are shown in Figure 9. As illustrated in Figure 9, the debris is composed of Fe oxides, indicating that both the coating and H13 steel undergo oxidative wear during the wear process.

3.6. Effect of Heating Temperature on the Wear Resistance of the FeCoCrNiAl + 10% WC Coating and H13 Steel

To explore the characteristics of high-temperature wear performance with varying temperatures, 623 K and 823 K were selected as the test temperatures for the FeCoCrNiAl + 10% WC composite coating and H13 steel.
The wear volumes of the FeCoCrNiAl + 10% WC composite coating and H13 steel at 623 K and 823 K are shown in Figure 10. As illustrated in Figure 10, the wear volumes of the coating are 0.098 mm3 and 0.187 mm3, respectively, while the wear volumes of H13 steel are 0.454 mm3 and 0.946 mm3, respectively. These results indicate that at the same wear temperature, the wear volumes of the coatings are reduced by 78.4% and 80.2% compared to H13 steel. This demonstrates that the composite coating has greater high-temperature wear resistance than H13 steel. The improvement is mainly attributed to the unique high-temperature properties of the coating.
Figure 11 shows the wear morphology of the FeCoCrNiAl + 10% WC composite coating and H13 steel at 623 K and 823 K. Figure 11a,c represent the wear morphologies of the FeCoCrNiAl + 10% WC composite coatings, while Figure 11b,d show the wear morphologies of H13 steel. It can be seen that there are a number of grooves and cracks, as well as debris, on the wear surface of H13 steel at 623 K and 823 K. With the increase in temperature, more cracks and delamination can be seen, and the depth of grooves increases. This indicates that the high-temperature wear performance of H13 surface deteriorates further with the increase in temperature. The wear forms are mainly abrasive wear, fatigue wear and plastic deformation. The wear surface of the FeCoCrNiAl + 10% WC coating is mainly composed of debris and grooves, and the groove width increases with the increase in temperature, which means the wear degree increases. The wear form is mainly abrasive wear.
In addition, it is clear that when wear temperature increases from 623 K to 823 K, the increase in the wear volume of H13 steel is 108.37%, the increase in the wear volume of the FeCoCrNiAl + 10% WC coating is 90.82%, and the increase in the wear volume of the FeCoCrNiAl + 10% WC coating is much smaller than that of H13. This indicates that the high-temperature wear resistance of the FeCoCrNiAl + 10% WC coating is significantly improved compared with that of H13 steel.

4. Conclusions

This study used laser cladding technology to prepare FeCoCrNiAl + WC composite coatings on the surface of H13 steel, analyzed the microstructure and hardness changes in composite coatings with different WC contents, and focused on the high-temperature friction and wear mechanisms of FeCoCrNiAl + 10% WC high-entropy alloy composite coatings. The conclusions are as follows:
(1)
The coating is mainly divided into the fusion zone, bonding zone, heat-affected zone, and substrate. The coating and substrate are tightly bonded, and the decomposition of WC particles increases the dendritic nucleation rate of the coating, forming a composite structure of columnar and equiaxed crystals. The phases of the coatings have CFe15.1 metal compound besides BCC and FCC solid solutions. CFe15.1 compounds increase with the increase in WC content;
(2)
The microhardness of the composite coating under different WC contents is 750HV0.1–850HV0.1, which is about 3.0–3.4 times that of the substrate. The wear volume is 0.187 mm3–0.251 mm3, which is reduced by 73.4%–80.2% compared to H13. The FeCoCrNiAl + 10% WC composite coating has the smallest wear volume.
(3)
At wear temperatures of 623 K to 823 K, the increase in wear volume of the FeCoCrNiAl + 10% WC coating decreases from 108.37% to 90.82% compared with H13 steel, and the FeCoCrNiAl + 10% WC coating has the greatest high-temperature wear resistance among the composite coatings. The high-temperature wear types of the composite coatings are abrasive wear and oxidation wear. In addition to abrasive wear and oxidation wear, H13 steel also exhibits fatigue wear.

Author Contributions

Conceptualization, D.Z.; Methodology, Y.G., M.J. and D.Z.; Formal analysis, Y.G., S.B., M.J., D.Z. and Y.L.; Investigation, Y.L.; Resources, Y.G., M.J. and Y.L.; Data curation, S.B.; Writing—original draft, S.J.; Writing—review & editing, S.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Project of Science and Technology research of the Department of Education of Jilin Province (JJKH20240135KJ), and the Natural Science Foundation of the Science and Technology Department of Jilin Province (20230101335JC).

Data Availability Statement

Data are contained within the article.

Acknowledgments

Special thanks go to the Laser Laboratory of Northeast Electric Power University for the help with this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Movassagh-Alanagh, F.; Abdollah-Zadeh, A.; Zolbin, M.A.; Nemati, N.; Aghababaei, R. Plasma-enhanced chemical vapor deposition of TiB2 and TiBN hard coatings using BBr3. Tribol. Int. 2023, 179, 108137. [Google Scholar] [CrossRef]
  2. Shi, W.; Wu, Y.P.; Hong, S.; Yang, W.H.; Wei, Z. Microstructural characterization and evolution under high temperature oxidation of CoCrAlYTa-10%Al2O3 coating deposited by high-velocity oxygen fuel thermal spraying. Mater. Res. Express 2020, 7, 016435. [Google Scholar] [CrossRef]
  3. Wang, X.Y.; Wang, J.H.; Gao, Z.M.; Xia, D.H.; Hu, W.B. Effect of Process Parameters on Electrodeposited Nanocrystalline Chromium Coatings Investigated by an Orthogonal Experiment. Prot. Met. Phys. Chem. Surf. 2020, 56, 857–866. [Google Scholar] [CrossRef]
  4. Wu, H.Y.; Ye, Y.M.; Lu, H.Q.; Yang, J.; Kong, D.J. Tribological behavior of laser thermal sprayed Cr3C2-NiCr +10%Ni/MoS2 composite coating on H13 hot work mould steel. Mater. Res. Express 2020, 7, 016599. [Google Scholar]
  5. Chiu, L.H.; Yang, C.F.; Hsieh, W.C.; Cheng, A.S. Effect of contact pressure on wear resistance of AISI H13 tool steels with chromium nitride and hard chromium coatings. Surf. Coat. Technol. 2002, 154, 282–288. [Google Scholar] [CrossRef]
  6. Chen, W.; Peng, Y.C.; Wang, Y.F.; Cao, P.; Zhu, Y.X.; Guo, Y.Q. Research on high-temperature friction and wear performances of Stellite 12 laser cladding layer against coated boron steels. Wear 2023, 520, 204665. [Google Scholar] [CrossRef]
  7. Yang, Y.H.; Li, J.H.; Yao, F.P.; Li, X.Y. Effect of laser remelting on the organization and properties of WC/Ni-based coatings generated in situ by laser cladding. J. Manuf. Process 2023, 102, 501–512. [Google Scholar] [CrossRef]
  8. Yang, C.; Jing, C.N.; Fu, T.L.; Lin, T.; Guo, W.M.; Liu, N.N. Effect of TiC addition on the microstructure and properties of laser cladding Ni60A coatings on H13 steel surface. Mater. Today Commun. 2024, 38, 107904. [Google Scholar] [CrossRef]
  9. Ding, L.; Wang, H.X.; Quan, X.M. Effect of aging treatment on microstructure and wear resistance of CoCrFeNiTiNbB1.25 high entropy alloy coatings by laser cladding. Appl. Phys. A Mater. 2022, 128, 235. [Google Scholar] [CrossRef]
  10. Li, Z.T.; Jing, C.N.; Feng, Y.; Wu, Z.L.; Lin, T.; Zhao, J.R. Phase evolution and properties of AlxCoCrFeNi high-entropy alloys coatings by laser cladding. Mater. Today Commun. 2023, 35, 105800. [Google Scholar] [CrossRef]
  11. Yang, C.; Jing, C.N.; Fu, T.L.; Lin, T.; Guo, W.M.; Liu, N.A. Effect of CeO2 on the microstructure and properties of AlCoCrFeNi2.1 laser cladding coatings. J. Alloy Compd. 2024, 976, 172948. [Google Scholar] [CrossRef]
  12. Shu, F.Y.; Wang, B.; Zhao, H.Y.; Tan, C.W.; Zhou, J.L.; Zhang, J. Effects of Line Energy on Microstructure and Mechanical Properties of CoCrFeNiBSi High-Entropy Alloy Laser Cladding Coatings. J. Therm. Spray Technol. 2020, 29, 789–797. [Google Scholar] [CrossRef]
  13. Wang, X.J.; Lan, Z.; Liu, Y.J.; Liu, N.; Fan, J.C.; Niu, M.Y.; Qiao, Y.X.; Liu, B. Laser assisted synthesis of Al0.1CoCrFeNi High Entropy Alloy Coating: Microstructures and Properties. Int. J. Electrochem. Sci. 2022, 17, 22088. [Google Scholar] [CrossRef]
  14. Shi, F.K.; Zhang, Q.K.; Xu, C.; Hu, F.Q.; Yang, L.J.; Zheng, B.Z.; Song, Z.L. In-situ synthesis of NiCoCrMnFe high entropy alloy coating by laser cladding. Opt. Laser Technol. 2022, 151, 108020. [Google Scholar] [CrossRef]
  15. Ding, L.; Hu, S.S.; Quan, X.M.; Shen, J.Q. Microstructure and high temperature tribological performance of Co-based laser cladded coatings reinforced with in-situ TiN-VC. Vacuum 2022, 198, 110894. [Google Scholar] [CrossRef]
  16. Guo, C.H.; Xu, S.C.; Chen, Z.B.; Gao, H.B.; Jiang, G.R.; Sun, W.Y.; Wang, X.H.; Jiang, F.C. Effect of B4C content and particle sizes on the laser cladded B4C/Inconel 625 composite coatings: Process, microstructure and corrosion property. J. Mater. Res. Technol. 2024, 30, 6278–6290. [Google Scholar] [CrossRef]
  17. Zhang, H.; Zhang, Y. Effects of h-BN content and heat treatment temperature on the tribological behavior and toughness of laser-clad Mo2FeB2/WC/h-BN coatings. Ceram. Int. 2024, 50, 26097–26108. [Google Scholar] [CrossRef]
  18. Shrey, B.; Suvradip, M. A study on the influence of reinforcement particle size in laser cladding of TiC/Inconel 625 metal matrix composite. Opt. Laser Technol. 2023, 161, 109115. [Google Scholar]
  19. Zhang, Z.Y.; Chen, Z.Q.; Yang, Y.Y.; Zhang, W.H.; Cui, H.T.; Zheng, X.; Mu, X.Y.; Xu, C.H. Effect of graphene and niobium carbide on microstructure and mechanical properties of Ni60 composite coatings prepared by laser cladding. J. Mater. Res. Technol. 2024, 30, 8277–8286. [Google Scholar] [CrossRef]
  20. Hao, X.H.; Liu, H.X.; Zhang, X.W.; Tao, J.T.; Wang, Y.Y.; Yang, C.; Liu, Y.X. Effect of Si3N4 content on microstructure and frictional-wear properties of MoNbTaWTi refractory high entropy alloy composite coatings. Mater. Today Commun. 2024, 40, 109416. [Google Scholar] [CrossRef]
  21. Yang, X.Y.; Wang, L.L.; Gao, Z.N.; Wang, Q.; Du, M.Z.; Zhan, X.H. WC distribution, microstructure evolution mechanism and microhardness of a developed Ti-6Al-4 V/WC MMC coating fabricated by laser cladding. Opt. Laser Technol. 2022, 153, 108232. [Google Scholar] [CrossRef]
  22. Lyu, Y.Z.; Varriale, F.; Malmborg, V.; Ek, M.; Pagels, J.; Wahlström, J. Tribology and airborne particle emissions from grey cast iron and WC reinforced laser cladded brake discs. Wear 2024, 556–557, 205512. [Google Scholar] [CrossRef]
  23. Han, P.; Kong, D.J. Laser cladded NiCoCrAlY–xNi60WC coatings: Microstructure and high temperature tribological properties. Intermetallics 2024, 173, 108399. [Google Scholar] [CrossRef]
  24. Liang, H.; Hou, J.X.; Li, X.C.; Jiang, L.; Cao, Z.Q. Micro-structures and dry sliding wear properties of novel AlCrFeNiMo0.5-x(WC) high entropy alloy-ceramic composite coatings. Intermetallics 2024, 170, 108328. [Google Scholar] [CrossRef]
  25. Wang, H.; Liu, K.; Li, J.; Geng, S.N.; Jing, L.; Skuratov, V. Reinforcements/matrix micro-interface evolution and properties of in-situ Ni60A/WC coatings prepared by laser cladding. Surf. Coat. Technol. 2024, 484, 130834. [Google Scholar] [CrossRef]
  26. Wu, T.; Yang, C.; Yu, L.T.; Zheng, X.F.; Zhang, L.F.; Jiang, Y.T.; Xue, Y.P.; Lu, Y.H.; Luan, B.L. Microstructure characterization and high-temperature wear mechanism of high-entropy alloy matrix composite coating fabricated by laser cladding. Appl. Surf. Sci. 2024, 677, 161032. [Google Scholar] [CrossRef]
  27. Karmakar, D.P.; Muvvala, G.; Nath, A.K. High-temperature abrasive wear characteristics of H13 steel modified by laser remelting and cladded with Stellite 6 and Stellite 6/30% WC. Surf. Coat. Technol. 2021, 422, 127498. [Google Scholar] [CrossRef]
  28. Li, Z.S.; Xie, D.Q.; Liu, Y.; Lv, F.; Zhou, K.; Jiao, C.; Gao, X.S.; Wang, D.S.; Liu, Y.J.; Zu, H.C.; et al. Effect of WC on the microstructure and mechanical properties of laser-clad AlCoCrFeNi2.1 eutectic high-entropy alloy composite coatings. J. Alloy Compd. 2024, 976, 173219. [Google Scholar] [CrossRef]
  29. Xu, L.F.; Li, M.X.; Song, Z.H.; Li, F.; Guo, J.; Gao, M. WC-High Entropy Alloy Reinforced Long Life Self-Grinding Silage Knife Prepared by Laser Cladding. Nanomaterials 2022, 12, 1013. [Google Scholar] [CrossRef]
  30. Cao, Q.Z.; Fan, L.; Chen, H.Y.; Hou, Y.; Dong, L.H.; Ni, Z.W. Wear behavior of laser cladded WC-reinforced Ni-based coatings under low temperature. Tribol. Int. 2022, 176, 107939. [Google Scholar] [CrossRef]
  31. Wang, H.R.; Sun, Y.F.; Qiao, Y.Z.; Du, X.S. Effect of Ni-coated WC reinforced particles on microstructure and mechanical properties of laser cladding Fe-Co duplex coating. Opt. Laser Technol. 2021, 142, 107209. [Google Scholar] [CrossRef]
Figure 1. The flow chart of laser cladding experiment.
Figure 1. The flow chart of laser cladding experiment.
Coatings 14 01259 g001
Figure 2. The macroscopic morphology of FeCoCrNiAl + WC composite coating: (a) 5% WC; (b) 10% WC; (c) 15% WC; (d) 20% WC.
Figure 2. The macroscopic morphology of FeCoCrNiAl + WC composite coating: (a) 5% WC; (b) 10% WC; (c) 15% WC; (d) 20% WC.
Coatings 14 01259 g002
Figure 3. The X-ray diffraction results of the composite coatings: (a) 20–95°; (b) 42–44°.
Figure 3. The X-ray diffraction results of the composite coatings: (a) 20–95°; (b) 42–44°.
Coatings 14 01259 g003
Figure 4. Microstructure of FeCoCrNiAl + 10% WC composite coating: (a) cross-section, (b) bonding zone, and (c) top microstructure of the coating.
Figure 4. Microstructure of FeCoCrNiAl + 10% WC composite coating: (a) cross-section, (b) bonding zone, and (c) top microstructure of the coating.
Coatings 14 01259 g004
Figure 5. Microhardness curves of the composite coatings.
Figure 5. Microhardness curves of the composite coatings.
Coatings 14 01259 g005
Figure 6. The friction force and coefficient curves of composite coatings at 823 K: (a) 5% WC; (b) 10% WC; and (c) 15% WC.
Figure 6. The friction force and coefficient curves of composite coatings at 823 K: (a) 5% WC; (b) 10% WC; and (c) 15% WC.
Coatings 14 01259 g006
Figure 7. The wear volumes of composite coatings and substrate.
Figure 7. The wear volumes of composite coatings and substrate.
Coatings 14 01259 g007
Figure 8. The wear morphology of H13 steel and composite coatings at 823 K: (a) H13 steel; (b) 5% WC; (c) 10% WC; and (d) 15% WC.
Figure 8. The wear morphology of H13 steel and composite coatings at 823 K: (a) H13 steel; (b) 5% WC; (c) 10% WC; and (d) 15% WC.
Coatings 14 01259 g008
Figure 9. The point scanning results of the wear debris on composite coatings and substrate: (a) H13 steel; (b) 5% WC; (c) 10% WC; and (d) 15% WC.
Figure 9. The point scanning results of the wear debris on composite coatings and substrate: (a) H13 steel; (b) 5% WC; (c) 10% WC; and (d) 15% WC.
Coatings 14 01259 g009
Figure 10. The wear volumes of 10% WC coatings and H13 steel at 623 K and 823 K.
Figure 10. The wear volumes of 10% WC coatings and H13 steel at 623 K and 823 K.
Coatings 14 01259 g010
Figure 11. The wear morphology of H13 steel and FeCoCrNiAl + 10% WC composite coatings: (a) FeCoCrNiAl + 10% WC coating at 623 K; (b) H13 steel at 623 K; (c) FeCoCrNiAl + 10% WC coating at 823 K; and (d) H13 steel at 823 K.
Figure 11. The wear morphology of H13 steel and FeCoCrNiAl + 10% WC composite coatings: (a) FeCoCrNiAl + 10% WC coating at 623 K; (b) H13 steel at 623 K; (c) FeCoCrNiAl + 10% WC coating at 823 K; and (d) H13 steel at 823 K.
Coatings 14 01259 g011
Table 1. Chemical composition of H13 steel (wt.%).
Table 1. Chemical composition of H13 steel (wt.%).
ElementCSiMnCrMoVSPFe
Mass ratio0.32~0.450.80~1.200.20~0.504.75~5.501.10~1.750.80~1.20≤0.03≤0.03Bal.
Table 2. The mass ratio of Fe, Co, Cr, Ni, and Al powders (wt.%).
Table 2. The mass ratio of Fe, Co, Cr, Ni, and Al powders (wt.%).
ElementFeCoCrNiAl
Mass ratio22.1223.3420.6023.2510.69
Table 3. The mass ratio of FeCoCrNiAl high-entropy alloy and WC powders.
Table 3. The mass ratio of FeCoCrNiAl high-entropy alloy and WC powders.
PowdersMass Ratio of HEAMass Ratio of WC
FeCoCrNiAl + 5% WC955
FeCoCrNiAl + 10% WC9010
FeCoCrNiAl + 15% WC8515
FeCoCrNiAl + 20% WC8020
Table 4. The parameters of laser experiment.
Table 4. The parameters of laser experiment.
Laser Power (W)Scanning Speed (mm/min)Overlap Rate (%)Laser Spot Diameter (mm)Argon Flow Rate (L/min)
12002403035
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, Y.; Jiang, S.; Bai, S.; Jie, M.; Zhang, D.; Liu, Y. Microstructure and Mechanical Properties of FeCoCrNiAl + WC Composite Coating Formed by Laser Cladding on H13. Coatings 2024, 14, 1259. https://doi.org/10.3390/coatings14101259

AMA Style

Gao Y, Jiang S, Bai S, Jie M, Zhang D, Liu Y. Microstructure and Mechanical Properties of FeCoCrNiAl + WC Composite Coating Formed by Laser Cladding on H13. Coatings. 2024; 14(10):1259. https://doi.org/10.3390/coatings14101259

Chicago/Turabian Style

Gao, Yali, Shan Jiang, Sicheng Bai, Meng Jie, Dongdong Zhang, and Yu Liu. 2024. "Microstructure and Mechanical Properties of FeCoCrNiAl + WC Composite Coating Formed by Laser Cladding on H13" Coatings 14, no. 10: 1259. https://doi.org/10.3390/coatings14101259

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop