Next Article in Journal
Thin Films and Coatings for Energy Storage and Conversion: From Supercapacitors and Batteries to Hydrogen Generators
Previous Article in Journal
Experimental Study on Bond Behavior between Steel Rebar and PVA Fiber-Reinforced Concrete
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dry Friction Properties of Friction Subsets and Angle Related to Surface Texture of Cemented Carbide by Femtosecond Laser Surface Texturing

1
College of Materials and Metallurgy, Guizhou University, Guiyang 550025, China
2
Guiyang High-Tech YiGe Electronic Co., Ltd., Guiyang 550022, China
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(4), 741; https://doi.org/10.3390/coatings13040741
Submission received: 4 March 2023 / Revised: 29 March 2023 / Accepted: 3 April 2023 / Published: 5 April 2023
(This article belongs to the Topic Laser Processing of Metallic Materials)

Abstract

:
This paper investigated the use of laser surface texturing (LST) to improve the tribological properties of YG6X cemented carbide. Three different spaced groove textures were processed on the surface of the YG6X carbide samples using a femtosecond laser. Friction experiments and friction simulations were performed under two friction subsets and two friction directions. The testing results showed that when the area density was 46%, the texture surface was beneficial when sliding against Si3N4, but not beneficial in reducing the coefficient of friction when sliding against Ti6Al4V titanium alloy. At area densities of 23% and 15.3%, the texture surface was beneficial when sliding against Si3N4, but not beneficial when sliding against the Ti6Al4V titanium alloy. When selecting the friction direction at 45° to the area density of 15.3%, the texture surface was not beneficial when sliding against the Si3N4 and Ti6Al4V titanium alloy. Sliding with Si3N4, the higher the stress value, the more easily the material was destroyed, leading to an elevated coefficient of friction and wear area. Sliding with Ti6Al4V titanium alloy, the higher the stress value of Ti6Al4V titanium alloy, the more easily the Ti6Al4V titanium alloy wore and generated a large number of abrasive chips.

1. Introduction

Titanium alloys have been widely used in the aerospace and biomedical industries due to their high specific strength, corrosion resistance, biocompatibility, and wide operating temperature range [1,2]. However, because of the low thermal conductivity of titanium alloys, the heat generated in the process of sliding with other materials cannot be effectively diffused, which has an adverse effect on the friction material. Second, titanium alloys weld easily with the frictional material during friction due to its high chemical reactivity, resulting in the premature failure of the frictional material [3]. Cemented carbide, mainly used for tool materials but also used as drawing dies and wear-resistant parts, is an alloy material made of hard compounds of refractory metals and viscous metals through powder metallurgy. Commonly used cemented carbides are divided into three types: tungsten-cobalt (YG), tungsten-cobalt-titanium (YT), and tungsten-titanium-tantalum (niobium) (YW). Among them, YG-type cemented carbides have good red hardness, high flexural strength, good thermal conductivity, and poor affinity with titanium alloys, which makes it suitable for processing titanium alloys, and its tribological properties are the key factor affecting its service life.
Surface texturing has remained an attractive area in recent years, which can further improve the wear resistance of materials [4]. There are various methods of micro-texturing such as electrical discharge machining (EDM) [5], lithography [6], micro-milling, and laser texturing (LST) [7]. EDM allows for the micro-fabrication of complex surfaces, but it can only machine conductive materials and has a slow machining speed. Lithography can obtain nanometer-level processing accuracy, however, it is not suitable for single-piece processing due to its high processing cost. Micro-milling can be used for the fine 3D surface machining of materials, but the size of the textures is generally in the micron level due to the size limitation of the milling cutter. Compared to other surface texturing techniques, laser texturing, especially femtosecond laser texturing, offers a smaller heat-affected zone, a precise processing area, flexible processing methods and high processing accuracy [8], and have been widely used for the surface texturing of metallic materials to improve their tribological properties.
A number of studies have found that texture morphology, geometric parameters, distribution, and the friction lubrication state have a significant effect on the performance of the texture. Fang et al. [9,10,11] presented different surface patterns with two different geometric structures that were processed on the contact surface of cemented carbide. The results showed that the linear pattern could have higher friction, while the pit pattern could effectively reduce the friction. Lian et al. [2] believed that a pit texture had better frictional properties than a groove texture, but that the pit texture could reduce the adhesion of titanium alloy under the condition using a solid lubricant. Mikhail et al. [12] found that the efficiency of tool turning of an Fe-Ni alloy could be improved by machining surface texturing on the cutting tool surface. Liu et al. [13] found that because the adhesion of the workpiece material on the surface of tools was obviously reduced, the cutting performance of micro-textured fluted carbide tools could be improved during the turning process of a Ti6Al4V titanium alloy. He et al. [14] studied the effect of circular pit, square pit, and ring groove micro-textured surfaces on the tribological properties of cemented carbide by the finite element simulation method and experiment. The results showed that the circular pit surface had a better friction performance compared to other types of surface micro-textures and smooth surfaces. Dan et al. [15] used a femtosecond laser to process four kinds of V-grooves on the surface of cemented carbide, with texture densities of 5.05%, 9.5%, 13.02%, and 15.2%, respectively. It was found that the surface friction coefficient of the sample with a texture density of 9.5% was the smallest and most stable. Zhang et al. [16] analyzed the tribological performance of YG8 with four types of groove-shaped textures under dry friction and solid lubricant, and corroborated that reasonable dimensional characteristics and distribution could reduce friction. Meng et al. [17] researched the dry machining of austenitic steels with cemented carbide tools and found that the groove texture captured wear debris under dry sliding conditions, thus reducing wear on the carbide. The beneficial effect of micro-scale grooves becomes higher with increasing groove density. Tong et al. [18] prepared micro-textures with variable distribution density on the carbide tool surface for high-speed milling tests with a titanium alloy and analyzed the effect of micro-texture parameters on tool wear, showing that the most determining factors were the diameter and spacing of the micro-textures. In the above study, it can be found that most of the studies considered only one kind of friction substrate and friction direction, but the process of friction is complex, and it is difficult to select only one kind of friction substrate [19] and friction direction [20] for an in-depth study of the actual mechanism of wear reduction.
This study prepared a variety of groove-textures on the surface of cemented carbide using a femtosecond laser, and the effects of texture spacing, friction balls, and friction directions on the tribological behavior of cemented carbide were analyzed to explore the influence mechanism of wear.

2. Experiment

2.1. Materials and Laser Surface Textured

In this experiment, silicon nitride ((Si3N4) balls, Ti6Al4V (TC4)) titanium alloy balls, and YG6X cemented carbide were used. The diameter of the balls was 6 mm, and the size of the bulk cemented carbide was 10 mm × 10 mm × 4 mm. The main components of YG6X cemented carbide are tungsten carbide (WC) and cobalt (Co). Before the micro-nano processing of the alloy surface, the surface of the YG6X alloy was abraded using 2000-mesh sandpaper, and then polished into mirror using diamond abrasive paste. The surface textures were processed by the femtosecond laser system (FemtoYL-50, Anyang Laser Technology Co., Ltd., Anyang, China). Grooves were created on the samples using a femtosecond laser with a maximum average power of 88.3 W, a pulse width of 362 (fs), a wavelength of 1035.43 nm, and a spot size of 73 μm. The laser processing parameters were optimized by regulating the frequency, speed, and processing times to obtain straight grooves and reduce the influence of the groove edge effect. The orthogonal test parameters are listed in Table 1. The groove profiles were observed by an optical microscope (GX51+DP26, Olympus, Tokyo, Japan). Under the optimized process parameters, groove-textures with different spacing were machined on the YG6X alloy, and their texture distribution is shown in Figure 1. We set the texture spacing (L1) to 0.25 mm, 0.5 mm, and 0.75 mm. L2 was the width of the machined grooves. The samples corresponding to these three texture spacings were named G1, G2, and G3 in that order. The untextured sample was named G0. Samples were ultrasonically cleaned with absolute ethanol before and after laser treatment, and then dried with compressed air.

2.2. Characterization of Mechanical Properties

A hardness tester (HV-1000, AICEYI Opto-electronic Technology Co., Ltd., Shenzhen, China) was used to measure the hardness of all samples. The detection force was 0.2 N, the loading time was 10 s, and four points were measured for each sample and the average value was taken. Tribological investigations were performed using a friction and wear testing machine (HSR-2M, Lanzhou Zhongke Kaihua Technology Development Co., Ltd., Lanzhou, China) with a ball-on-disc contact configuration adopted under unlubricated ambient reciprocating sliding conditions. The silicon nitride balls, TC4 titanium alloy balls, and YG6X cemented carbide were subjected to friction experiments. The loading force was 50 N, the friction rate was 66.67 mm/s, and the diameter was 4 mm. The friction test lasted for 30 min with 15,000 reciprocations. The samples before and after the friction experiment were ultrasonically cleaned with absolute ethanol for 5 min. The morphology of the abrasion marks was observed by laser scanning confocal microscopy (LSCM) (OLS5000, Olympus, Tokyo, Japan) and scanning electron microscopy (SEM) (SUPRA 40, Zeiss, Oberkochen, Germany).

2.3. Finite Element Analysis of the Tribological Behavior of Groove-Textured YG6X Alloy

The tool was replaced by a cemented carbide, and the workpiece was replaced by a titanium alloy ball. The ball-block test model was established, and the finite element analysis of the ball-block test was carried out, and the stress distribution diagram was obtained. Here, the volume of the YG6X and TC4 simulation models was consistent with the actual volume of each material in the friction experiment, the groove width was set to 115 μm, and the groove width was set to 40 μm. The TC4 simulation model used a C3D4T four-node thermally coupled cell, and the YG6X simulation model used a C3D8RT eight-node thermally coupled hexahedral cell using reduced integration. The element size of the specimen and the ball were 0.4 and 0.3, respectively. The YG6X simulation model kept the bottom surface fixed and defined the TC4 simulation model’s pressure and velocity. Pure elasticity calculation was conducted using an explicit formula. Their computations were made under the purely elastic deformation assumption and stresses may have be overestimated. The coefficient of friction was 0.4 and was obtained from friction experiments with TC4 by non-textured carbide. The focus of this study was to explore the stress on the contact surface during the friction process, the simulation of different grinding ball materials to ensure that the contact stress had the same law, so only the TC4 and YG6X friction pairs were selected for simulation. The simulation models of YG6X cemented carbide with different textured spacing are shown in Figure 2. F denoted the applied load, R denoted the radius of the grinding ball, and V denoted the sliding speed of the grinding ball.
The finite element simulation of the friction process was carried out using Abaqus software. The material properties are shown in Table 2. The temperature-displacement coupling analysis step was established in the analysis step module, and the total time was set to 0.0042 s. The whole process was set up so that the friction ball moved from one end of the groove textures to the other and moved a small distance over the original carbide surface. Then, the YG6X was fixed and restrained, and a pressure of 1.67 MPa was applied to the surface of the titanium alloy hemisphere. Finally, a velocity/angular velocity analysis type was established for TC4, and the velocity was set to 66.67 mm/s according to this analysis step. Previous work by numerous researchers has shown that sliding orientation plays a vital role in tribological performance. We selected a sample according to the simulation results to study the effect of a friction direction of 45° on wear resistance.

3. Results and Discussion

3.1. Optimization of Laser Processing Parameters

A series of groove textures prepared with different frequencies, speeds, and processing times are shown in Figure 3.
The grooves shown in Figure 3b were straight with a clear edge shape, while the grooves shown in Figure 3a were straight with a blurred edge shape, which was a problem of formability. The grooves shown in Figure 3c,d were not straight with a blurred edge shape. The optimized laser texturing parameter corresponding (100 kHz, 500 mm/s, processed 7 times) to the T7 groove was selected here due to the straight groove with a clear edge. The average width value of the T7 groove was 115 μm and the depth of the T7 groove was 4–5 μm, as shown in Figure 4a. The hardness value of 1903.4 HV0.2 was obtained near the edge of the groove and 1800.0 HV0.2 was obtained away from the edge of the groove, respectively, indicating that there was an increase in hardness at the edge of the groove during the laser texturing. This was considered here as laser processing creates a heat-affected zone in the material, melting and regenerating and creating a hard layer of the material at the edges of the grooves [21]. Figure 4c displays the morphology of the T7 groove. It could be seen that after femtosecond laser processing, there was no obvious convex structure on the edge of the groove and many micro/nano particle structures were formed inside the trench.
In the process of groove texturing, there was a problem of groove formability, so the morphological values were quoted here to analyze this orthogonal test. The evaluation of the shape value was divided into two parts: the groove textures were straight and the edge of the groove textures was clear. If one of them was met, the value was assigned to 1. If both were met, the value was assigned as 2. The morphological values corresponding to the orthogonal tests are collated in Table 3. The experimental results were analyzed visually using the analysis of range differences to obtain the mean response table listed in Table 4. The K average value represents the average value of the morphology value at each level of each factor, reflecting the influence of different levels of the same factor on the test results. R is equal to the maximum value of the K average value minus the minimum value, reflecting the magnitude of change in the test index when the factor level is changed. As shown in Table 4, the range in frequency (0.75) was greater than the range in the scanning speed (0.25) and processing times (0.25).
It can be concluded that the laser processing frequency had the dominant influence on the morphological values of the groove textures and it was easier to obtain groove textures that met the requirements at 100 kHz. Afterward, if further study of other parameters of the grooves is needed, the frequency can be fixed at 100 kHz, thus adjusting the speed and the processing times to achieve it.

3.2. Friction Simulation Result

Stress is an important factor causing surface wear during sliding, and surface texture affects stress distribution [3]. In order to study the effect of groove textures on the stress distribution of the samples, ABAQUS software (version 6.14) as used to simulate the stress field in the friction process between YG6X cemented carbide with different texture spacing and the TC4 titanium alloy. The simulation results of the stress field on the YG6X surfaces when the sliding direction was perpendicular to the groove textures are shown in Figure 5a–d. These simples were named G1-90°, G2-90°, and G3-90°. It was found that the stress and its distribution could be changed by preparing the groove texture on the surface.
The simulation results of the stress field on the surface of TC4 are shown in Figure 6a–d. The maximum stress of TC4 during the friction process also changed continuously, as shown in Figure 6e. The relationship between the areal density and the average stress of G-90°, G2-90°, and G3-90° is shown in Figure 6f. The average stress value first increased and then decreased with the decrease in areal density. The average stress value of G1-90°, with a 46% areal density, was the minimum. The larger the stress of TC4, the easier TC4 was removed, resulting in G2-90° and G3-90° having more wear debris than G1-90°, so G1-90° had the best wear resistance.
Subsequently, G3 was selected to research the influence of sliding orientation between the groove and grinding ball on the stress field of YG6X and TC4. When the sliding orientation was 45°, the simulation results of stress field of the groove-textured YG6X and TC4 ball can be seen as exhibited in Figure 7a,b, respectively. This sample was named G3-45°. The stress value during the friction process also changed continuously, as shown in Figure 7c,d, and the average stress values of YG6X and TC4 were 444.04 MPa and 286.15 MPa, respectively. Compared with the results of G3-90°, the average stress value of the groove-textured YG6X and TC4 ball both increased when the sliding orientation was 45°.

3.3. Friction Characteristics

In order to explore the friction and wear behavior of the textured samples under dry friction, using Si3N4 and TC4 as the sliding ball, friction and wear tests were carried out when the sliding direction was perpendicular to the groove textures. Figure 8a displays the coefficient of friction (COF) curves of samples sliding with the Si3N4 ball when the sliding direction was perpendicular to the groove textures. The COF is the ratio between the tangential and the normal load Ft/Fn. The COF of all samples initially increased, then remained stable with increasing time, and the average value of COF was G2-90° < G3-90° < G0 < G1-90°. The results demonstrate that the groove textures had the potential for reducing the friction coefficients under dry friction. Figure 8b,c demonstrates the wear cross section and the wear cross-sectional area of YG6X with different texture spacings, respectively. The wear area of G1-90° (8508.480 μm2) was the largest, which was increased by 38.17% compared to G0 (6157.948 μm2). However, the wear area of G2-90° (5510.097 μm2) and G3-90° (5858.555 μm2) were separately decreased by 10.52% and 4.86% compared to G0.
The SEM and LSCM morphologies of the worn YG6X surface, sliding with Si3N4, are shown in Figure 9. The results showed that the surface of YG6X started to peel off after several passes of low-speed reciprocating friction under high load and had the same worn-out appearance. This was mainly a tearing and peeling phenomenon, but the surface peeling phenomenon was more obvious in G1-90° and less so in G2-90° and G3-90°. This was consistent with the previous simulation results. During the wear process, by the action of pressure and shear force, G1-90°, with a high stress concentration value, was prone to sprout micro-cracks, and crack expansion led to the occurrence of spalling. Because of the wear debris-trapped effect [22], as shown in Figure 10, the depth of the wear scars of G2-90° and G3-90° was shallower than that of G0 because the groove texture with a certain surface density could effectively store the abrasive particles generated during the wear process and reduce the grinding on the surface of the material. When sliding with the Si3N4 ball and a sliding direction perpendicular to the groove textures, the groove textures, whose area densities were 15.3% and 23%, could obviously decrease the wear resistance of the material surface, while the groove texture improved the wear resistance of the material surface when the area density was 46%.
Figure 11a displays the COF curves of the samples sliding with the TC4 ball when the sliding direction was perpendicular to the groove textures. The average value of COF was G2-90° < G1-90° < G3-90° < G0. The results demonstrated that the groove textures had a potential for increasing the COF between YG6X and TC4 under dry friction. Figure 11b,c demonstrate the wear cross section and the wear cross-sectional area of YG6X under sliding with TC4 at different texture spacing conditions, respectively. The wear area of G2-90° (455,536.012 μm2) and G3-90° (34,781.901 μm2) were separately increased by 39.64% and 6.66% compared to G0 (32,610.684 μm2). The wear area of G1-90° (30,362.702 μm2) was decreased by 6.90% compared to G0. When sliding with the TC4 ball and a sliding direction perpendicular to the groove textures, the groove textures, whose area densities were 15.3% and 23%, could obviously improve the wear resistance of the material surface, while the groove texture decreased the wear resistance of the material surface when the area density was 46%.
The SEM and LSCM morphologies of the worn surface on the YG6X sliding with TC4, are shown in Figure 12. These mainly showed tearing and spalling phenomena, which was typical of adhesive wear characteristics. From the aforementioned simulation results, it could be seen that when the surface area density was 23% (G2), the stress concentration on the surface of TC4 ball was the largest, which might be susceptible to sprouting micro-cracks and crack expansion and cause the occurrence of spalling by the action of pressure and shear force in the sliding process. In addition, the soft and tough TC4 wear chips could not be collected effectively due to the lower area density and adhered to the surface of YG6X, producing an obvious tearing phenomenon.
From the results above-mentioned, G3 had better friction properties when sliding with Si3N4 and TC4 balls under a sliding direction perpendicular to the groove textures. Subsequently, the effect of sliding orientation on the friction and wear behavior of G3-90° was studied. Figure 13 shows the friction performance of G3 when it slid with Si3N4 under the sliding orientation of 45°. At the beginning of the friction, the COF increased more compared to G3-90°, as shown in Figure 13a. The wear area was 8481.196 μm2, greater than G3-90° (5858.555 μm2), as shown in Figure 13b. The YG6X surface had more serious peeling and produced multiple deep pits compared to G3-90°, as shown in Figure 13c. Figure 14 shows the friction performance of G3 when it slid with TC4 under the sliding orientations of 45° and 90°. The COF of G3-45° had an increasing trend to be greater than that of G3, as shown in Figure 14a. The wear area was 34,884.417 μm2 greater than G3-90° (34,781.901 μm2), as shown in Figure 14b. Figure 14c also showed a large number of TC4 abrasive chips adhered on the YG6X surface. It was also concluded that the friction properties of the textured samples did not improve with this friction orientation, but rather increased the wear of the material surface.

4. Conclusions

Three different spacing groove textures were machined on the YG6X cemented carbide by a femtosecond laser. The effect of the groove texture spacing, friction subsets, and friction directions on friction wear was analyzed by ball-block reciprocating friction simulations and tests. The main conclusions obtained can be summarized as follows:
(1)
Using the optimized laser texturing parameter corresponding to 100 kHz, 500 mm/s, processed 7 times could obtain a straight groove with a clear edge on the surface of the YG6X carbide.
(2)
Sliding with hard material Si3N4 and the sliding direction perpendicular to the groove textures, when the area densities of the groove textures were 15.3% and 23%, the groove textures could reduce the friction coefficient and wear area of cemented carbide at the same time, while the friction coefficient and wear area of the YG6X cemented carbide increased when the area density of the groove texture was 46% compared with the smooth surface. When the area density of the groove texture was 23%, it reduced the wear resistance, changing the direction of friction to 45° compared with the original.
(3)
Sliding with the TC4 titanium alloy and a sliding direction perpendicular to the groove textures when the area densities of the groove textures were 15.3% and 23% showed that the groove textures could increase the friction coefficient and wear area of cemented carbide at the same time, while the groove textures could reduce the wear area of cemented carbide when the area density of the groove texture was 46% compared with the smooth surface. When the area densities of the groove texture were 23%, it reduced the wear resistance, changing the direction of friction to 45° compare with the original.
(4)
The simulation results were consistent with the experimental results. For the texture samples, the higher the average stress, the easier it was to wear.

Author Contributions

Conceptualization, H.C., F.Z. and Z.F.; methodology, H.C., F.Z. and Z.F.; software, H.C.; validation, H.C., F.Z. and Z.F.; formal analysis, H.C.; investigation, H.C.; resources, H.C.; data curation, H.C.; writing—original draft preparation, H.C.; writing—review and editing, H.C., F.Z. and Z.F.; visualization, H.C.; supervision, H.C., F.Z. and Z.F.; project administration, H.C.; funding acquisition, F.Z. All authors have agreed to take responsibility for the entire content of the research work. 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] grant number [No. 51965010], [the Nature Science Foundation of Guizhou Provincial Science and Technology Department] grant number [No. (2020)1Y202], and [the National Natural Science Foundation of China] grant number [No. (2019) 44]. The APC was funded by [Fang Zhou].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

LSCMLaser scanning confocal microscopy
SEMScanning electron microscope
COFCoefficient of friction
TC4Ti6Al4V
EDMElectrical discharge machining
LSTLaser texturing
G0Untextured YG6X carbide
G1YG6X carbide with 0.25 mm texture spacing
G2YG6X carbide with 0.50 mm texture spacing
G3YG6X carbide with 0.75 mm texture spacing
G1-90°The friction subsets slide perpendicular to the grooves of the G1 sample
G2-90°The friction subsets slide perpendicular to the grooves of the G2 sample
G3-90°The friction subsets slide perpendicular to the grooves of the G3 sample
G3-45°The friction subsets slide against the grooves of the G sample at 45 degrees

References

  1. Zhu, S.; Liu, J.; Deng, X. Modification of strain rate strengthening coefficient for Johnson-Cook constitutive model of Ti6Al4V alloy. Mater. Today Commun. 2021, 26, 102016. [Google Scholar] [CrossRef]
  2. Avcu, Y.Y.; Iakovakis, E.; Guney, M.; Çalım, E.; Özkılınç, A.; Abakay, E.; Sönmez, F.; Koç, F.G.; Yamanoğlu, R.; Cengiz, A.; et al. Surface and Tribological Properties of Powder Metallurgical Cp-Ti Titanium Alloy Modified by Shot Peening. Coatings 2023, 13, 89. [Google Scholar] [CrossRef]
  3. Lian, Y.; Mu, C.; Wang, L.; Yao, B. Numerical simulation and experimental investigation on friction and wear behaviour of micro-textured cemented carbide in dry sliding against TC4 titanium alloy balls. Int. J. Refract. Met. Hard Mater. 2018, 73, 121–131. [Google Scholar] [CrossRef]
  4. Wang, Z.; Li, Y.B.; Bai, F.; Wang, C.W. Angle-dependent lubricated tribological properties of stainless steel by femtosecond laser surface texturing. Opt. Laser Technol. 2016, 81, 60–66. [Google Scholar] [CrossRef]
  5. Rajbongshi, S.K.; Sarma, D.K. Performance parameters studies in machining of AISI D2 steel with dot-textured, groove-textured & non-textured cutting tool at the flank face. Int. J. Refract. Met. Hard Mater. 2019, 83, 104970. [Google Scholar] [CrossRef]
  6. Sarkar, S.S.; Katiyar, A.K.; Sarkar, A.; Dhar, A.; Rudra, A.; Khatri, R.K.; Ray, S.K. Germanium growth on electron beam lithography patterned Si3N4/Si (001) substrate using molecular beam epitaxy. Appl. Surf. Sci. 2018, 437, 144–151. [Google Scholar] [CrossRef]
  7. Sawant, M.S.; Jain, N.; Palani, I.A. Influence of dimple and spot-texturing of HSS cutting tool on machining of Ti-6Al-4V. J. Mater. Process. Technol. 2018, 261, 1–11. [Google Scholar] [CrossRef]
  8. Ródenas, A.; Gu, M.; Corrielli, G.; Paiè, P.; John, S.; Kar, A.K.; Osellame, R. Three-dimensional femtosecond laser nanolithography of crystals. Nat. Photonics 2019, 13, 105–109. [Google Scholar] [CrossRef] [Green Version]
  9. Fang, S.; Herrmann, T.; Rosenkranz, A.; Gachot, C.; Marro, F.G.; Mücklich, F.; Llanes, L.; Bähre, D. Tribological performance of laser patterned cemented tungsten carbide parts. Procedia CIRP 2016, 42, 439–443. [Google Scholar] [CrossRef] [Green Version]
  10. Fang, S.; Llanes, L.; Klein, S.; Gachot, C.; Rosenkranz, A.; Bähre, D.; Mücklich, F. Frictional performance assessment of cemented carbide surfaces textured by laser. IOP Conf. Ser. Mater. Sci. Eng. 2017, 258, 012006. [Google Scholar] [CrossRef]
  11. Fang, S.; Soldera, F.; Rosenkranz, A.; Herrmann, T.; Bähre, D.; Llanes, L.; Mücklich, F. Microstructural and metallurgical assessment of the laser-patterned cemented tungsten carbide (WC-CoNi). Procedia Manuf. 2018, 26, 198–204. [Google Scholar] [CrossRef]
  12. Stebulyanin, M.; Ostrikov, E.; Migranov, M.; Fedorov, S. Improving the Efficiency of Metalworking by the Cutting Tool Rake Surface Texturing and Using the Wear Predictive Evaluation Method on the Case of Turning an Iron–Nickel Alloy. Coatings 2022, 12, 1906. [Google Scholar] [CrossRef]
  13. Liu, X.; Liu, Y.; Li, L.; Tian, Y. Performances of micro-textured WC-10Ni3Al cemented carbides cutting tool in turning of Ti6Al4V. Int. J. Refract. Met. Hard Mater. 2019, 84, 104987. [Google Scholar] [CrossRef]
  14. He, C.; Yang, S.; Zheng, M. Analysis of synergistic friction reduction effect on micro-textured cemented carbide surface by laser processing. Opt. Laser Technol. 2022, 155, 108343. [Google Scholar] [CrossRef]
  15. Dan, L.; Xuefeng, Y.; Chongyang, L.; Jian, C.; Shouren, W.; Yanjun, W. Tribological characteristics of a cemented carbide friction surface with chevron pattern micro-texture based on different texture density. Tribol. Int. 2020, 142, 106016. [Google Scholar] [CrossRef]
  16. Zhang, Z.; Lu, W.; He, Y.; Zhou, G. Research on optimal laser texture parameters about antifriction characteristics of cemented carbide surface. Int. J. Refract. Met. Hard Mater. 2019, 82, 287–296. [Google Scholar] [CrossRef]
  17. Meng, R.; Deng, J.; Liu, Y.; Duan, R. Improving tribological performance of cemented carbides by combining laser surface texturing and WSC solid lubricant coating. Int. J. Refract. Met. Hard Mater. 2018, 72, 163–171. [Google Scholar] [CrossRef]
  18. Tong, X.; Shen, J.; Su, S. Properties of variable distribution density of micro-textures on a cemented carbide surface. J. Mater. Res. Technol. 2021, 15, 1547–1561. [Google Scholar] [CrossRef]
  19. Magnus, C.; Gulenc, I.T.; Rainforth, W.M. Ambient dry sliding friction and wear behaviour of laser surface textured (LST) Ti3SiC2 MAX phase composite against hardened steel and alumina. Wear 2022, 490, 204184. [Google Scholar] [CrossRef]
  20. Huang, J.; Yang, S. Investigation on anisotropic tribological properties of superhydrophobic/superlipophilic lead bronze surface textured by femtosecond laser. Appl. Surf. Sci. 2022, 579, 152223. [Google Scholar] [CrossRef]
  21. Huang, J.; Guan, Y.; Ramakrishna, S. Tribological behavior of femtosecond laser-textured leaded brass. Tribol. Int. 2021, 162, 107115. [Google Scholar] [CrossRef]
  22. Lu, L.; Zhang, Z.; Guan, Y.; Zheng, H. Comparison of the effect of typical patterns on friction and wear properties of chromium alloy prepared by laser surface texturing. Opt. Laser Technol. 2018, 106, 272–279. [Google Scholar] [CrossRef]
Figure 1. Texture spacing diagram, L1 is the texture spacing and L2 is the width of the machined grooves.
Figure 1. Texture spacing diagram, L1 is the texture spacing and L2 is the width of the machined grooves.
Coatings 13 00741 g001
Figure 2. The simulation models of (a) G0, (b) G1, (c) G2, (d) G3.
Figure 2. The simulation models of (a) G0, (b) G1, (c) G2, (d) G3.
Coatings 13 00741 g002
Figure 3. Optical micrographs of the groove textures (a) T1–T4, (b) T5–T8, (c) T9–T12, (d) T13–T16.
Figure 3. Optical micrographs of the groove textures (a) T1–T4, (b) T5–T8, (c) T9–T12, (d) T13–T16.
Coatings 13 00741 g003
Figure 4. (a) The width of the groove textures. (b) The hardness of the edge of the groove textures. (c) The SEM images of the groove textures.
Figure 4. (a) The width of the groove textures. (b) The hardness of the edge of the groove textures. (c) The SEM images of the groove textures.
Coatings 13 00741 g004
Figure 5. Simulation results of the stress field of the (a) G0, (b) G1, (c) G2, and (d) G3 samples. (e) Maximum stress distribution curves and (f) the relationship between the areal density and average stress of samples when the sliding direction was perpendicular to the groove textures.
Figure 5. Simulation results of the stress field of the (a) G0, (b) G1, (c) G2, and (d) G3 samples. (e) Maximum stress distribution curves and (f) the relationship between the areal density and average stress of samples when the sliding direction was perpendicular to the groove textures.
Coatings 13 00741 g005
Figure 6. Simulation results of the stress field of TC4 sliding with the (a) G0, (b) G1, (c) G2, and (d) G3 samples. (e) Maximum stress distribution curves and (f) the relationship between the areal density and average stress of TC4 when the sliding direction was perpendicular to the groove textures.
Figure 6. Simulation results of the stress field of TC4 sliding with the (a) G0, (b) G1, (c) G2, and (d) G3 samples. (e) Maximum stress distribution curves and (f) the relationship between the areal density and average stress of TC4 when the sliding direction was perpendicular to the groove textures.
Coatings 13 00741 g006
Figure 7. When the sliding orientations was 45°, (a,c) show the simulation results of YG6X. (b,d) are the simulation results of TC4.
Figure 7. When the sliding orientations was 45°, (a,c) show the simulation results of YG6X. (b,d) are the simulation results of TC4.
Coatings 13 00741 g007
Figure 8. The COF (a) and wear section (b) and its area (c) of YG6X with different texture spacings under the sliding direction of the Si3N4 ball perpendicular to the groove texture condition.
Figure 8. The COF (a) and wear section (b) and its area (c) of YG6X with different texture spacings under the sliding direction of the Si3N4 ball perpendicular to the groove texture condition.
Coatings 13 00741 g008
Figure 9. The SEM and LCM morphologies of the worn YG6X surface of (a) G0, (b) G1, (c) G2, and (d) G3 under the sliding direction of the Si3N4 ball perpendicular to the groove texture condition.
Figure 9. The SEM and LCM morphologies of the worn YG6X surface of (a) G0, (b) G1, (c) G2, and (d) G3 under the sliding direction of the Si3N4 ball perpendicular to the groove texture condition.
Coatings 13 00741 g009
Figure 10. Wear mechanism analysis of the (a) smooth surface and (b) textured surface.
Figure 10. Wear mechanism analysis of the (a) smooth surface and (b) textured surface.
Coatings 13 00741 g010
Figure 11. The COF (a) and wear section (b) and its area (c) of YG6X with different texture spacings under the sliding direction of the TC4 ball perpendicular to the groove texture condition.
Figure 11. The COF (a) and wear section (b) and its area (c) of YG6X with different texture spacings under the sliding direction of the TC4 ball perpendicular to the groove texture condition.
Coatings 13 00741 g011
Figure 12. The SEM and LCM morphologies of the worn YG6X surface of (a) G0, (b) G1, (c) G2, and (d) G3 when sliding with TC4.
Figure 12. The SEM and LCM morphologies of the worn YG6X surface of (a) G0, (b) G1, (c) G2, and (d) G3 when sliding with TC4.
Coatings 13 00741 g012
Figure 13. The (a) COF, (b) wear section and area, and (c) SEM and LCM morphologies of G3 in the 45° and 90° friction direction when sliding with Si3N4.
Figure 13. The (a) COF, (b) wear section and area, and (c) SEM and LCM morphologies of G3 in the 45° and 90° friction direction when sliding with Si3N4.
Coatings 13 00741 g013
Figure 14. The (a) COF, (b) wear section and area; (c) SEM and LCM morphologies of G3 in the 45° and 90° friction directions when sliding with TC4.
Figure 14. The (a) COF, (b) wear section and area; (c) SEM and LCM morphologies of G3 in the 45° and 90° friction directions when sliding with TC4.
Coatings 13 00741 g014
Table 1. Laser texture parameters.
Table 1. Laser texture parameters.
GroupFrequency (kHz)Speed (mm/s)Processing Times (Times)
T1251251
T2252503
T3255005
T42510007
T51001253
T61002501
T71005007
T810010005
T93001255
T103002507
T113005001
T1230010003
T138001257
T148002505
T158005003
T1680010001
Table 2. Material properties of the block and ball samples.
Table 2. Material properties of the block and ball samples.
Materials Name
Material Properties
Block Sample
(YG6X)
Ball Sample
(Ti6Al4V)
Density (kg/m3)1.485 × 10204.43 × 1021
Modulus of elasticity (GPa)640114
Poisson’s ratio0.220.33
Conductivity (W/m·k)79.66.6
Thermal expansion coefficient (×10−6/°C)4.79
Specific heat capacity (J/(kg·°C))176670
Table 3. Shape values (the evaluation of the shape value was divided into two parts: the groove textures were straight and the edge of the groove textures was clear; the value 1 was assigned if one criteria was met, and the value 2 was assigned if both criteria were met).
Table 3. Shape values (the evaluation of the shape value was divided into two parts: the groove textures were straight and the edge of the groove textures was clear; the value 1 was assigned if one criteria was met, and the value 2 was assigned if both criteria were met).
GroupT1T2T3T4T5T6T7T8T9T10T11T12T14T15T16
Shape values111121221111111
Table 4. Analysis of range.
Table 4. Analysis of range.
LevelFrequencySpeedProcessing Times
K average value1 time--1.00
3 times--1.25
5 times--1.25
7 times--1.25
25 kHz1.00--
100 kHz1.75--
300 kHz1.00--
800 kHz1.00--
125 mm/s-1.25-
250 mm/s-1.00-
500 mm/s-1.25-
1000 mm/s-1.25-
R0.750.250.25
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

Cheng, H.; Zhou, F.; Fei, Z. Dry Friction Properties of Friction Subsets and Angle Related to Surface Texture of Cemented Carbide by Femtosecond Laser Surface Texturing. Coatings 2023, 13, 741. https://doi.org/10.3390/coatings13040741

AMA Style

Cheng H, Zhou F, Fei Z. Dry Friction Properties of Friction Subsets and Angle Related to Surface Texture of Cemented Carbide by Femtosecond Laser Surface Texturing. Coatings. 2023; 13(4):741. https://doi.org/10.3390/coatings13040741

Chicago/Turabian Style

Cheng, Hang, Fang Zhou, and Zihao Fei. 2023. "Dry Friction Properties of Friction Subsets and Angle Related to Surface Texture of Cemented Carbide by Femtosecond Laser Surface Texturing" Coatings 13, no. 4: 741. https://doi.org/10.3390/coatings13040741

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