Next Article in Journal
Plasma-Sprayed Osseoconductive Hydroxylapatite Coatings for Endoprosthetic Hip Implants: Phase Composition, Microstructure, Properties, and Biomedical Functions
Next Article in Special Issue
Comparison of Corrosion Behavior of a-C Coatings Deposited by Cathode Vacuum Arc and Filter Cathode Vacuum Arc Techniques
Previous Article in Journal
Design and Performance Study of Carbon Fiber-Reinforced Polymer Connection Structures with Surface Treatment on Aluminum Alloy (6061)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Research on Improving Wear Resistance of Titanium Alloys

1
No. 3 Oil Production Plant, PetroChina Changqing Oilfifield Company, Yinchuan 750006, China
2
School of Material Science and Engineering, Xi’an Shiyou University, Xi’an 710065, China
3
State Key Laboratory of Rolling and Automation, Northeastern University, Shenyang 110819, China
4
State Key Laboratory of Tribology in Advanced Equipment, Tsinghua University, Beijing 100084, China
5
Nation Center for Materials Service Safety, University of Science and Technology Beijing, Beijing 100083, China
6
Henan International Joint Laboratory of Dynamics of Impact and Disaster of Engineering Structures, Nanyang Institute of Technology, Nanyang 473004, China
7
Petroleum Systems Engineering, Faculty of Engineering and Applied Science, University of Regina, Regina, SK S4S 0A2, Canada
8
Shaanxi Coal Industry New Energy Technology Co., Ltd., Xi’an 710199, China
9
China Railway First Survey and Design Institute Group Co., Ltd., Xi’an 710043, China
*
Authors to whom correspondence should be addressed.
Coatings 2024, 14(7), 786; https://doi.org/10.3390/coatings14070786
Submission received: 7 June 2024 / Revised: 18 June 2024 / Accepted: 20 June 2024 / Published: 24 June 2024

Abstract

:
Titanium alloy is widely used as oil drill pipe material because of its light weight, high strength, good toughness, corrosion resistance, fatigue resistance, and good process performance. However, due to its low hardness, poor wear resistance, serious oxidation at high temperature (700 °C), and difficulty in lubrication, in oil and gas field exploration and development drilling, especially in deep wells, high displacement wells, horizontal wells, and highly deviated wells, wear and tear are prone to occur. The application and development of titanium alloys are greatly limited. This paper introduces the research status of the common surface modification technologies of titanium alloys, such as laser cladding, magnetron sputtering, plasma spraying, micro arc oxidation, etc. It points out the improvement effect of various modification technologies on the wear resistance and high-temperature oxidation resistance of titanium alloys and discusses the advantages and disadvantages of various modification technologies. A proposed method for enhancing the wear resistance and high-temperature oxidation resistance of titanium alloys was finally introduced, and its potential for future development was investigated.

1. Introduction

The main energy and important chemical raw materials in the world today have always been oil and natural gas. Their products are widely used in various fields of human social activities, deeply penetrate into all aspects of people’s lives, and are almost ubiquitous. Oil Country Tubular Goods (OCTGs), also known as special oil pipes, are the foundation of the petroleum industry. Casing, tubing, and drill string components (drill pipe, drill collar, kelly, etc.) are collectively referred to as oil well pipes. When the material of oil well tubing is exposed to aggressive environments, oil well pipe materials are prone to corrosion and wear [1]. Therefore, wellhead equipment, oil pipelines, and drilling tools face many problems [2]. Titanium alloy is considered as a candidate material for oil drill pipes and oil well pipes because of its light weight, good toughness, high strength, corrosion resistance, fatigue resistance, and excellent process performance (cutting, grinding, extrusion, and forging). Compared to highly alloyed nickel-based corrosion-resistant alloys, the price of titanium alloys is lower, and with the industrialization of titanium alloys in recent years, the current price is only about one-third that of ordinary carbon steel. According to the research results on the mechanical properties of high-performance drill pipes made of different materials (as shown in Table 1), it can be seen that titanium alloy has a significant advantage in its strength to gravity ratio.
However, titanium alloys have a lower hardness (the Vickers hardness is usually between 250 and 350) and poor wear resistance, and they are prone to wear and tear in oil and gas field exploration and development drilling [3,4,5]. Corrosion and wear seriously affect the safety and durability of titanium alloys. The wear of a drill pipe will cause significant damage to the oil and gas field; for example, (1) the production oil pipe cannot be put into the well normally; (2) when the damaged part of the casing is located in the water layer or sand layer, the oil well will produce a large amount of water or sand; (3) oil well stimulation measures cannot be carried out; (4) it is easy to cause blowout outside the casing, and the well can be scrapped. Although tubing, casing, delivery pipes, and other titanium products have been successfully developed, there is little research and application work involving these products in China’s oil and gas wells [6,7]. It is of great significance to reduce the friction coefficient of titanium alloys and improve their wear resistance in order to expand their industrial applications.
At present, people have explored many surface treatment methods and technologies, through the use of surface laser remelting, laser alloying, laser cladding, chemical vapor deposition, physical vapor deposition, nitriding, carburizing, plasma spraying, and other processes to improve and enhance the surface properties and wear resistance of titanium alloys [8]. However, these methods can only obtain a thinner coating after a long treatment time, and the coating does not have enough durability or load bearing capacity. At present, a large number of surface treatment methods and technologies such as carburizing, coating, and plasma nitriding [9] have been studied to improve the surface properties of titanium alloys and, at the same time, provide application prospects for improving the surface wear resistance of materials.
The research progress of the surface modification of titanium alloys at home and abroad is reviewed in this paper and summarizes the differences between various technologies. This provides effective information for future researchers to improve the wear resistance of titanium and its alloys. On the basis of introducing the research status and advantages and disadvantages of surface carburizing/nitrogen/boron technology, pulsed magnetic field treatment technology, laser cladding technology, wear-resistant surfacing technology, the application prospects of surface modification technology such as additive manufacturing technology, magnetron sputtering technology, and micro arc oxidation technology are also discussed in this paper. The application trend of surface modification technology of various titanium alloys is analyzed to improve their wear resistance and meet the requirements of the modern manufacturing industry.

2. Surface Modification Technology of Titanium Alloy

2.1. Surface Carburizing/Nitrogen/Boron Technology

Carburizing the surface of a titanium alloy is a technique used to enhance its wear resistance and hardness by creating a hardened layer of TiC through the carburization process, as shown in Figure 1. TiC has a cubic crystal structure. It is an important metal carbide that can be used in metal ceramic materials and cutting materials. It has a high thermal stability, high melting point, and a high hardness and Young’s modulus. It has not only the advantages of high temperature resistance and low friction coefficient, but it can also improve the biological compatibility of titanium alloys. The TiC layer formed by carburizing treatment thickens with the increase in temperature and does not separate from the subsurface layer. Most titanium alloys have a hexagonal close-packed (HCP) structure. Due to the large lattice size of carbon atoms in the carburizing process, it is not easy to enter the sub surface of titanium alloys in the diffusion process dominated by the gap mechanism. As the combination of oxygen atoms and titanium alloys is easier than that of carbon atoms, the carburizing treatment of titanium alloys should be carried out under the condition of isolating oxygen and high temperature. Compared with other alloys such as steel, titanium alloy carburizing requires more rigorous process conditions and a more complex carburizing mechanism [10,11,12,13]. Hydrogen embrittlement is inevitable due to the introduction of the hydrogen element in traditional carburizing, so hydrogen-free carburizing is usually adopted [14].
The common carburizing methods of titanium alloys include gas carburizing, laser melting carburizing, and plasma carburizing. The method of gas carburizing is similar to plasma carburizing, and the microstructure of the carburized layer is also similar. For gas carburizing, CO and methane are generally selected as the gas carbon source [15]. Carbon atoms are obtained from the atmosphere through the material transfer at the gas–solid interface, and the carbon is infiltrated into the workpiece through the principle of thermal diffusion to form TiC in the titanium or titanium alloy matrix. The gas carburizing operation is relatively simple, the carburizing speed is fast, and the whole carburizing process takes less time. However, since hydrogen will cause hydrogen embrittlement, which will affect the performance of titanium alloy, the carburizing process requires a high gas composition in the furnace. The carbon source of gas carburizing is usually hydrocarbon gas, which is divided into hydrocarbon gas carburizing and ion gas carburizing. When methane is used as the carburizing gas, a combination of titanium carbide and amorphous hexagonal carbon with a high hardness, high melting point, and thermochemical stability can be obtained. The gas content, gas composition, gas flow rate, and the exposure time of the sample in the gas have a significant impact on the mechanical properties and microstructure of the titanium carbide layer [16]. If the methane content is too high, solid carbon deposition will be formed on the sample surface, which will hinder the further generation of TiC [17]. As the affinity between titanium and oxygen is higher than that between titanium and carbon, thin titanium oxide films have been formed before most experiments, preventing the further diffusion of carbon ions [18]. Therefore, ion etching can be carried out on titanium alloys during gas carburizing to eliminate the influence of oxidized film on carburizing.
Yang et al. [19] found that after 820 °C × 10 h of low-pressure nitriding treatment, the surface phase of the Ti6Al4V (TC4) titanium alloy is mainly composed of TiN, TiAlN, and Ti3Al phases. Its surface hardness reaches 800–900 HV, nearly three times higher than that of the substrate without nitriding treatment. The wear resistance of the Ti6Al4V (TC4) titanium alloy is greatly improved due to its complete film layer, shallow furrow, and no tear marks.
It is possible to improve the wear resistance and hardness of titanium alloys by forming a boronized layer on their surface [20]. Thus far, many methods of boronizing have been carried out, such as gas boronizing, molten salt boronizing, and laser pulse boronizing [21,22,23]. Among them, due to its simple process and easy formation of low-cost and dense boride layers, the boronization method has become the most commonly used method to improve the surface properties of titanium alloys [24]. The Ti5-5Mo-5V-8Cr-3Al titanium alloy (TB2 alloy) is a metastable material with an excellent cold working performance, impact toughness, welding performance, and plasticity. It is a β type titanium alloy, with a tensile strength greater than 1400 MPa [25]. Because of the excellent performance of the TB2 alloy, it has a very broad application prospect in important parts related to aerospace [26]. By boriding, the TB2 alloy’s low hardness and inadequate wear resistance can be remedied. The research shows that the surface hardness of the TB2 alloy with boride layer can be increased to 27.5 GPa (nano indenter microhardness test), and the friction coefficient can be reduced to 0.34 [27]. These improvements can significantly improve the performance of the TB2 alloy, so that it can be used in important spacecraft and aircraft parts. In addition, by adding CeO2, lattice distortion can be generated to improve the diffusion of boron during boronizing [28]. In order to promote the diffusion of boron, rare earth elements can also increase the concentration of vacancies in the displacement mechanism [29] and obtain a boride layer with a relatively high surface hardness and low friction coefficient [30].
From the room temperature wear morphology of the TB2 alloy as shown in Figure 2, the surface of the TB2 alloy shows numerous deep grooves after the wear test. This is attributed to the low surface hardness of the TB2 alloy, which results in the detachment of surface matrix particles when cut by the Al2O3 ball. Furthermore, the reciprocating movement between the Al2O3 ball and the TB2 alloy leads to a temperature increase, causing the formation of additional TiO2 particles on the alloy’s surface. The presence of adhesion between TiO2 particles and matrix particles brings about the emergence of deep, elongated grooves on the TB2 alloy’s surface, indicative of a combination of abrasive wear and adhesive wear. In contrast, the boron-plated TB2 alloy exhibits a smoother surface with no prominent scratches, as the boride layer protects it from oxidation even under elevated contact temperatures. The presence of boride microbumps allows for the formation of abrasive particles during the wear process, highlighting the pivotal role of abrasive wear in the friction and wear mechanisms. While adhesive wear remains the primary cause of TB2 workpiece failure, the application of boron coating treatment enhances the wear resistance of the alloy, ultimately extending the service life of its crucial components. The experiment shows that when TB2 alloy is used in a relatively high-temperature environment (250 °C), the boride particles can reduce the surface abrasion of alumina balls and titanium dioxide particles. In addition, abrasive wear plays an important role after boronizing at 1100 °C, effectively improving the wear resistance of the boronized TB2 alloy [31].
Bloyce et al. [32] developed a simple and effective surface modification technology, namely palladium-treated thermal oxidation (PTO). The wear resistance of surface engineering titanium-based materials was studied by conventional plasma nitriding (PN), thermal oxidation (TO), and the newly developed PTO process. Compared with PN-treated and untreated materials, TO- and PTO-treated materials have a significantly excellent corrosion resistance in boiling HCl solution. The breakdown life of a protective surface layer of titanium treated with TO in boiling 20% HCl solution is about 13 times that of titanium treated with PN, while the life of the material treated with PTO is 2.6 times longer than that of the material treated with TO. Under the condition of oil lubrication, the material treated by PTO has a better scratch resistance than that treated by TO. Glow discharge spectroscopy, X-ray diffraction, and scanning electron microscopy (SEM) were used to characterize the surface layers treated with TO and PTO. The results showed that the technology effectively improved the wear resistance of titanium-based materials.

2.2. Pulse Magnetic Field Treatment

Pulse magnetic field treatment refers to a method to refine the alloy structure during the solidification process of the alloy, as shown in Figure 3. After applying a pulsed magnetic field during the casting process, the electromagnetic effect causes the melt to generate an induced current. The interaction between the induced current and the pulse magnetic field will generate an electromagnetic force (Lorentz force), which acts on the melt and causes movement in the melt, thus refining the grain structure during the solidification process. The enhancement in properties comes from the improvement in the remanence and hardness during pulse magnetization, especially the synergism between remanence and Fe3O4 nanoparticles plays a key role. Fe3O4 nanoparticles are attracted by the magnetic ball, which increases the concentration of the friction contact area, weakens the adhesion of titanium alloy, forms a film containing Fe3O4 nanoparticles, and improves the wear resistance.
Yang et al. [33] conducted a pulsed magnetic field treatment on cemented carbide (WC–12Co) to improve its wear resistance when contacting the TC4 titanium alloy, as shown in Figure 4. Based on the magnetic response after magnetic particle testing, ferromagnetic Ni and Fe3O4 nanoparticles and non-ferromagnetic Cu were selected as additives to prepare lubricating fluid. The results show that the magnetic particle inspection improves the wear resistance of a cemented carbide/titanium alloy contact under the lubrication of a cutting fluid. In the cutting fluid lubricant, compared with the untreated ball, the workpiece wear of the coupling decreased by 33.9% compared to before, and the ball wear marks also decreased by 36.4%. This enhanced wear resistance is mainly due to the magnetic response of Fe3O4 to the treated magnetic ball and the improvement of hardness. The adsorption of Fe3O4 nanoparticles onto the treated magnetic spheres reduces the adhesion of TC4 and forms a titanium alloy film containing Fe3O4 nanoparticles.
Figure 4b,c show the surface morphology of TC4 samples after magnetic field treatment (MT); b1, b2, c1, and c2 are magnetic domains, and the domain size statistics are determined. The standard deviation of the untreated samples is relatively low, and their size is small and uniform. However, the MT samples are significantly larger than the untreated samples. This is because the MT samples are magnetized by a pulsed magnetic field. During this process, small magnetic domains merge into large ones, and the domain walls migrate, reducing the number [34]. In short, the combination of magnetic domains, the migration of magnetic domain walls, and the magnetostriction caused by pulse magnetic field magnetization create conditions for an improvement in hardness. This is beneficial to reduce “contact severity” and improve wear resistance. The cooperation between magnetized remanence and ferromagnetic Fe3O4 further improves the wear resistance. The attraction of the magnetic sphere to the nanoparticles increases the concentration of the particles in the contact area. Fe3O4 nanoparticles can prevent more titanium alloys from adhering and strengthen the titanium alloy film.

2.3. Laser Cladding Technology

Laser cladding technology [35], also known as direct energy deposition, refers to a new surface modification technology that melts the metallurgical combination of the matrix and powder at the same time to form a particle reinforced composite layer on the metal matrix, as shown in Figure 5. Under the irradiation of a laser, the cladding powder will rapidly melt and form a molten pool on the surface of the substrate. When the laser is removed, the molten pool will rapidly cool and solidify, and the coating will be metallurgically bonded with the substrate. Due to the high cooling rate, the coating structure will be refined to give the coating an excellent toughness [36]. Compared with other surface modification technologies, laser cladding technology has the following advantages: dense microstructure cladding layer can be prepared; it can combine two kinds of completely different matrix materials and cladding materials, which are almost unlimited by the types of matrix materials and cladding layers; the laser used has the excellent characteristics of a high energy density, clean and pollution-free, easy to realize automation, etc.
Yuan et al. [37] performed laser surface texturing (LST)–thermal oxidation (TO) double treatment on the Ti6Al4V alloy to improve its mechanical properties and wear properties, and they characterized the microstructure, bonding strength, microhardness, and nano indentation behavior of TO-Ti6Al4V. Under dry sliding conditions, the wear properties of the original Ti6Al4V, LST-Ti6Al5V, TO-Ti6Al4V, and DT-Ti6Al4V-related samples after being double treated (DT) were compared and evaluated. The results show that the thickness of 17 μm for a gradient titanium oxide coating was prepared, the content of O presents a gradient distribution on the whole thermal oxidation coating, and the thermal oxidation treatment improves the surface hardness and elastic modulus of Ti6Al4V. The specific wear rate results of the four treated samples are as follows: DT-Ti6Al4V<TO-Ti6Al4V<LST-Ti6Al4V<Ti6Al4V, which indicates that DT-Ti6Al4V has the best wear performance. It has been proved that Ti6Al4V can obtain significantly enhanced wear properties by using LST and TO treatment.
Proper surface texturing provides a new method to reduce the friction and wear of titanium alloys [38]. Laser processing technology discards the traditional texture processing technology. At the same time, it has become a promising surface manufacturing technology because of its advantages such as a competitive repeatability, high precision, no pollution to the environment, and a fast processing speed [39]. It has been proved that the specific texture size and configuration on the contact interface can help to reduce the vibration and noise caused by friction [40], and it can also be used as a micro memory for wear debris to reduce the three body wear caused by wear debris during sliding friction [41]. Many scholars have devoted themselves to determining the influence of surface texture unit shape on tribological properties. Wu et al. [42] successfully treated the surface texture of pits on the Ti6Al4V alloy by laser processing. It was pointed out that the dimple texture plays a positive role in reducing the wear rate of friction pairs. Kashyap et al. [39] proved that the groove texture on the polymer surface of the Ti6Al4V alloy has an excellent effect on reducing adhesion wear. Zhang et al. [43] prepared a petal-like texture on a Co–Cr–Mo alloy, proving that the relevant mechanism for improving the tribological properties is the formation of micro reservoirs between petals. Therefore, the pit surface texture unit is not affected by the sliding direction, And the process is easy to operate. It can change the state of the contact surface [44,45], which can significantly improve the wear performance of titanium alloys.
However, the laser-based manufacturing method produces a large temperature gradient near the molten pool due to its high energy density input. The non-uniform thermal expansion and contraction in the heat-affected zone results in the formation of residual stresses in the finished workpiece [46]. When laser cladding heterogeneous materials, the effect of tensile stress on the cladding quality is particularly obvious. Using materials with a similar thermal expansion coefficient is an effective method to reduce the tensile stress caused by a thermal mismatch. The thermal expansion coefficient of titanium alloys is similar to WC, so WC particles are considered as ideal materials to improve the surface properties of titanium alloys [47]. Because the wettability between ceramic and metal materials is usually poor, the bonding between the particles and the substrate can cause serious problems. Wang et al. [48] prepared a Stellite/WC composite coating on an AISIH13 hot working tool steel by laser cladding WC (WC-12Co) particles and W-CrCo alloy powder. The results show that WC-12Co particles can further improve the wear resistance of the coating compared with only WC particles as the coating material. Fan et al. [49] used a 4 kW fiber laser to deposit a cobalt-based coating with 40% WC on the surface of a 15MnNi4Mo steel cone bit. The coating is metallurgically bonded with the substrate, and the coating is compact without cracks, pores, and other defects.
Jiang and Zhang, along with their colleagues [47], employed WC-17Co spherical powder as the primary material for producing WC-Co composite coatings on the Ti-6Al4V alloy surface through laser cladding technology. They observed a direct correlation between the WC grain diameter and the microhardness of the composite coatings. Composite coatings with smaller grain diameters have a higher number of grain boundaries, which results in an increased resistance to dislocation movement and a reduced likelihood of plastic deformation. There is a lot of spalling on the worn surface. With an increase in laser power, the spalling conditions are improved. The main reasons for the spalling of composite coatings are tensile stress and plastic deformation under external load. In this study, a three-dimensional impact abrasion test was performed on the samples to examine the wear behavior of the WC-Co composite coating. The experimental results are depicted in Figure 6. Initially, the micro bulges on the coating surface created numerous small-area contact points during the experiment. At the same time, under the impact of high energy, the particles which not firmly bonded on the coating surface fall off, and the contact points are bonded and worn seriously, with a high wear rate. As the raised micro peaks are gradually rubbed off, the coating surface becomes smooth. Under the effect of surface hardening, the wear rate of the coating decreases rapidly until it becomes stable. It can be seen that the slope of 1400 W composite coating is the largest, indicating that the most serious wear occurs at this time. After that, the 2000 W composite coating first enters the stable wear stage, and the wear rate stabilizes at about 1.5 g/h. At the end of the experiment, the wear rate is slightly different under a different laser beam power. The results show that 2000 W has the lowest wear rate and the highest wear resistance. It is worth noting that the reduction in wear rate means an improvement in wear resistance, and the laser power plays a key role in wear resistance. However, there are few reports on WC-Co coatings in recent years, and there is still a lack of systematic research on the influence of laser power on the microstructure and wear resistance of WC-Co composite coatings.
Laser cladding technology, as a modern surface treatment technology, has the characteristics of a short production time, high bonding strength, low dilution ratio, small heat-affected zone, and strong design flexibility [50,51,52]. In the process of laser cladding, the composite coating undergoes repeated rapid heating and cooling cycles, leading to complex non equilibrium solid phase transitions, which will significantly affect the microstructure and performance, thus affecting the service life of parts under harsh working conditions [53].

2.4. Plasma Spraying Technology

Plasma spraying is considered as one of the key technologies of advanced remanufacturing technology, as shown in Figure 7. It has a high-temperature flame flow, high-speed particles, and adaptability to manufacturing high-performance coatings. Arc plasma-sprayed films and coatings exhibit excellent mechanical properties in corrosion protection, wear resistance, radiation resistance, thermal isolation, and heat conduction. These coatings have been used for metal parts of aircraft engines, automobile engines, gas turbine engines, diesel engines, nuclear power equipment, and oil refining equipment. The microstructure and properties of the deposited coatings vary with the plasma spraying processing parameters.
In general, the deposition of coatings by plasma spraying involves different types of material flow and atomization. Under the constant power of metal or ceramic powder, various spraying distances are used to deposit coatings. Wang and Li et al. [54] studied plasma-sprayed Al2O3-13 wt.% TiO2 coatings’ microstructure and nanostructure, which were successfully deposited on a Ti6Al4V titanium alloy substrate, respectively. To enhance the density and bonding strength of the coating, the researchers employed a CO2 laser to remelt the sprayed coating. The study employed scanning electron microscopy, Vickers microhardness testing, and X-ray diffraction to investigate the impact of laser remelting on the microstructure, mechanical properties, and phase composition of ceramic coatings. The results indicated that the coating had a more uniform and compact structure, as well as a robust metallurgical bond with the substrate after undergoing laser remelting. During laser remelting, the main γ-Al2O3 phase is converted into stable α-Al2O3. The microhardness values of conventional Metco 130 coatings and nanostructured Al2O3-13 wt.% TiO2 coatings are between 700 and 1000 HV0.3, while the microhardness values of the corresponding remelted coatings are increased to 1000–1350 HV0.3 and 1100–1800 HV0.3, respectively. As the laser scanning speed decreases, the microhardness and wear resistance also increase. Figure 8 shows the relationship between the distribution of microhardness and the coating thickness of the sprayed coating and laser-remelted coatings (LRmCs). Obviously, the microhardness value of the received titanium alloy substrate is in the range of 350–400 HV0.3, while that of the sprayed microstructure and nanostructure coating is in the range of 700–1000 HV0.3. After laser remelting, the microhardness of ceramic laser-remelted coatings (C-LRmCs) increases by about 30%, reaching the range of 1000–1350 HV0.3. Accordingly, the microhardness of nanostructured laser-remelted coatings (N-LRmCs) increased by about 60%, reaching the range of 1100–1800 HV0.3. In addition, the microhardness of LRmCs typically increases with a decrease in the laser beam scanning speed. This trend is more obvious in N-LRmCs.

2.5. Additive Manufacturing Technology

Additive manufacturing technology is a technology that can combine surface texture with a solid lubricant coating, as shown in Figure 9. Compared to traditional manufacturing processes, additive manufacturing technology has more advantages, such as producing complex shapes and an extremely high dimensional accuracy, and is regarded as one of the ideal technologies for manufacturing complex parts [55,56]. The honeycomb structure has a high adhesion, minimum material requirements, and maximum available space, which is the best topology covering a two-dimensional plane [57,58]. Due to their excellent lubrication, chemical stability, corrosion resistance, low cost, light weight, and ease of manufacture, honeycomb structures are often used as coatings [59]. Polytetrafluoroethylene (PTFE) was pressed into a honeycomb structure by hot pressing sintering as a solid lubricant. This composite coating has zero wear performance and a low friction coefficient in various service environments such as the atmosphere, deionized water, seawater, and acidic media. Compared with other preparation methods of similar coatings, the preparation method is fast and low-cost [60], and the method of constructing self-lubricating coating is not limited to TC4, but can be used for iron-based alloys, nickel-based alloys, cobalt chromium alloys, aluminum alloys, copper alloys, and other metals.
The TEM image of the worn cross-section of the sample after plasma electrolytic oxidation treatment of a magnetron sputtering aluminum-plated Ti6Al4V alloy is shown in Figure 10 [61]. Two dense layers are tightly bonded to the surface of the zirconia ball, with a total thickness of 50 nm (Figure 10a). EDS point analysis (Figure 10b) shows that the first layer is amorphous carbon. The adjacent second layer is composed of four elements, i.e., C, O, F, and Zr. The high-resolution TEM image (Figure 10c) and Inverse Fast Fourier Transform (IFFT) analysis (Figure 10 d–f) show that there are zirconia nanocrystals with an atomic spacing of 0.29 nm and 0.30 nm. In addition, a large area of typical amorphous substances surround zirconia nanocrystals (Figure 10g) and form very thin carbon layers (10~15 nm) mainly due to the defluorination of PTFE during friction. In addition, the Raman spectra results of the friction ball wear surface also show that PTFE and carbon exist on the wear surface. The defluorination phenomenon is more likely to occur in the formation process of the PTFE transfer film. A thin and uniform transfer film is more beneficial to the friction resistance and wear resistance of PTFE composites than a thick and discontinuous transfer film. Therefore, this technology has great potential in improving the friction and wear properties of metal and alloy parts.

2.6. Magnetron Sputtering Technology

Magnetron sputtering technology uses Ar+ to bombard the target surface with a certain amount of energy under the acceleration of a high-voltage electric field which causes molecules or atoms on the target surface to escape from the target surface, achieving the effect of sputtering, as shown in Figure 11. Molecules or atoms escaping from the target surface will fly to the substrate surface and deposit on it to form a thin film. Magnetron sputtering is a promising technology. Through this technology, films with a good surface uniformity and adhesion can be prepared at a relatively low substrate temperature. At the same time, this method has another advantage; that is, it has a fast growth rate and can be used for large-scale production. The hard film prepared by magnetron sputtering technology has an excellent wear resistance and corrosion resistance [62,63,64].
Kang et al. [65] proposed that the wear resistance of Ti6Al4V alloy can be greatly improved by the combination of magnetron sputtering and plasma electrolysis oxidation (PEO). His research, for the first time, obtained an ~13 μm pure aluminum layer, deposited on the Ti6Al4V alloy by magnetron sputtering, and then further treated with PEO. The PEO of aluminum-coated Ti6Al4V was carried out in aluminate (32g L −1) and silicate (16g L−1) electrolytes, and the wear properties of the samples with chromium steel balls were tested under an external load of 10 N. The coating formed in 32g L−1 aluminum electrolyte shows an excellent wear performance, its sliding time to the steel ball lasts 1800 s, and the wear rate is not high. On the contrary, the coating formed in the silicate electrolyte for 15 min has been destroyed in the tribological test, showing 3.9 × 10−4 mm3/(N·m). The excellent wear performance of coatings formed in aluminate electrolyte can be attributed to its high growth rate and homogeneity of microstructure. Figure 12 shows the 3D wear scar display of different specimens after a dry sliding test (confocal method). Figure 12a,b show that the samples of an Al/Ti6Al4V two-phase system and the coating formed in silicate for 15 min have been seriously worn. The coating formed in the aluminate electrolyte for 1 min shows a wide wear track (Figure 12c) after sliding for about 1200 s; however, the wear track is shallower than that in Figure 12a,b. The coating formed in aluminate electrolyte for 4 min shows the shallowest wear depth, which means that the coating has a good wear resistance. Figure 12e shows the three-dimensional morphology of the wear scar of the PEO film of aluminized Ti6Al4V alloy treated in 32 g·L−1 NaAlO2 for 1 min after 1200 s wear test. It can be seen from the figure that the color of the wear scar on the film surface is significantly lighter than that of the film in Figure 12c,d, which indicates that the wear resistance of the PEO film of aluminized Ti6Al4V alloy prepared by magnetron sputtering in 32 g·L−1 NaAlO2 is higher than that of the film prepared in the sodium silicate electrolyte. Figure 12f shows the three-dimensional morphology of the wear scars of a PEO film prepared by magnetron sputtering an aluminized Ti6Al4V alloy in 32 g·L−1 NaAlO2 for 4 min after 1800 s wear test. It can clearly be seen from the figure that the wear scar morphology of the film is the lightest among all the films, and no obvious wear scar profile can be observed, indicating that the PEO film prepared under this condition has an excellent wear resistance.
Magnetron sputtering technology is simple in equipment, easy to control, and can easily prepare large-area, uniform, and dense hard films; The target space of magnetron sputtering is very wide. According to the requirements of different equipment, many materials can be made into target materials, including various metals, semiconductors, ferromagnetic materials or insulating oxides, ceramics, polymers, etc. In the process of magnetron sputtering, the required reaction gas can also be mixed into the sputtering chamber. Nitrogen and oxygen are the most common gases. Through further processing, the performance of the hard film prepared by magnetron sputtering will continue to improve. Because of these remarkable characteristics of magnetron sputtering technology, it has become one of the main industrial coating technologies. The hard films prepared by magnetron sputtering technology have a high economic value.

2.7. Micro Arc Oxidation Technology

Micro arc oxidation (MAO), also known as plasma electrolytic oxidation (PEO), is the in situ conversion of the surface atoms of valve metals (such as Al, Mg, and Ti) into their oxides through micro arc discharge under a pulsed electric field in the electrolyte, thus greatly improving the corrosion resistance and wear resistance of a titanium substrate [66,67,68], as shown in Figure 13. In addition, the technology is simple and environmentally friendly. The optimization of the friction reduction and wear resistance of titanium alloys via a micro arc oxidation film mainly focuses on the following three aspects: electrolyte, electrical parameters, and composite treatment [69].
The core process parameter of micro arc oxidation is the electrolyte. Therefore, the formation mechanism of a micro arc oxidation film on the surface of a titanium alloy varies in different electrolyte systems. The prepared micro arc oxidation films have a different chemical composition and structure, which will also affect their tribological properties. The hardness of the film is closely related to its tribological properties. A film with a high hardness often has a better wear resistance [70]. The main components of the titanium alloy participating in the oxidation film formation reaction are rutile and anatase crystalline TiO2, both of which have a hardness of less than 600 HV, which is not significantly improved compared with the hardness of the titanium alloy itself (about 400 HV) [71]. Therefore, if one wants to further improve the wear resistance of the film, the electrolyte needs to provide other film-forming components. The common oxidation liquid silicate is the most widely used, because SiO2 has an extremely strong adsorption capacity, and it is extremely easy to be adsorbed on the surface of titanium-based materials to form impurity discharge centers during the oxidation process. On one hand, it can achieve stable film formation in a wide temperature and current range. On the other hand, due to the deposition of a large amount of colloidal silica, the film is relatively thicker, up to 90 μm. The hardness of the oxide film with amorphous SiO2 as the main component is generally only 600~800 HV, and the wear rate is high during the wear process of the friction pair with high hardness, so the ion concentration is the key to affecting the wear resistance of the oxide film.
The oxidation solution of a single system cannot present the best film characteristics, so the oxidation film characteristics of each system can be reasonably used, and the oxidation film with an optimized performance can often be obtained by using the combination method [72]. When studying the influence of the electrolyte system on the film formation mechanism of micro arc oxidation of titanium alloys, Li and Yang et al. [73] found that adding an appropriate amount of phosphate in silicate electrolyte can promote the formation of a passive film at the initial stage of oxidation. Phosphide will preferentially enrich on the surface of titanium alloys and grow toward the interior of the matrix, while silicon-containing oxides grow outward, mainly distributed in the outer loose layer. The adhesion and density of the composite film are better than those of the single Si or P system. Yerokhin et al. [74] compounded aluminate and phosphate to prepare an oxide film with a dense structure and smooth surface. In the counter grinding experiment with SAE52100 steel, the volume wear rate is only 3.4 × 108 mm3/(N.m), which not only solves the problem of insufficient thickness of the oxide film of the single P system, but also significantly improves the wear resistance of the film due to the rich hard phases such as Al2O3 and AlTiO5 in the film.
Quintero et al. [75] compared the effects of constant current and constant voltage modes on the wear resistance of a titanium alloy micro arc oxidation film in three electrolyte systems. The results indicate that in the friction test, from the perspective of energy consumption, the unit energy consumption of the constant pressure mode is about half of that of the constant current mode. In addition, the mass wear rate of coatings prepared under constant pressure is lower than that prepared under constant current mode. Therefore, when the power supply allows, the constant voltage operation mode has more advantages. Liu and Blawert et al. [9] carried out micro arc oxidation on Ti6Al4V titanium alloy drill pipe materials. Grafting graphene into the micro arc oxidation coating formed on the Ti6Al4V alloy can effectively improve the wear resistance. Using micro arc oxidation surface treatment technology combined with graphene nanosheets is an innovative and promising method to improve the wear resistance of Ti6Al4V alloy drill pipes.
Lan et al. [76] manufactured a filtered cathodic vacuum arc system on the TC4 titanium alloy by an MAO/tetrahedral amorphous carbon (ta-C) composite coating and studied the surface morphology, tribological properties, and corrosion resistance of the coating in simulated seawater. The results show that the MAO/ta-C composite coating is still porous and has a better adhesion than the single MAO coating or ta-C film. Figure 14a–c show that the MAO/ta-C coating exhibits the best electrochemical stability and corrosion resistance. Figure 14d indicates that MAO/ta-C composite coating has the lowest friction coefficient (0.13) and wear scar width (356.84 um) among all samples due to the close oxidation ceramic coating of the MAO coating and the self-lubricating of the ta-C film. In the simulated seawater solution, the lowest corrosion current density (8.321 × 109 A/cm2) and the highest corrosion potential (0.012 V) confirm that the MAO/ta-C coating provides the most effective corrosion protection. These results can provide important support for promoting the application of titanium. Figure 14 shows the friction coefficients between three TC4 surface coatings and a TC4 substrate and GCr15 in simulated seawater solution. It can be seen from the figure that the friction coefficient fluctuates greatly with the increase in wear time. The friction coefficient of the TC4 substrate and the ta-C film shows a strong fluctuation, which is kept at about 0.3, but the counterpart of the MAO coating and MAO/ta-C coating is kept between 0.10 and 0.25. Due to the poor combination of the ta-C film and TC4 substrate, the ta-C film was stripped from the TC4 substrate before wear in the simulated aqueous solution. As is known, the MAO coating contains high-hardness silicon dioxide and titanium oxide. The porous MAO coating can store simulated seawater to reduce friction. However, the loose and rough surface characteristics of the MAO coating cause obvious fluctuations in the friction coefficient. It should be noted that the friction coefficient of the MAO/taC coating is lower than that of a single MAO coating or ta-C film (about 0.1), which may be due to the high hardness of the MAO coating and the good combination of the ta-C film and the MAO coating.

3. Conclusions

  • Titanium alloys play a crucial role in the petroleum industry, especially in oil well tubing materials, because of their outstanding specific strength, toughness, corrosion resistance, and fatigue resistance. There is a need to enhance and advance the surface modification technology of titanium alloys, and to obtain more extensive applications and better economic benefits, increasing innovation is needed. This work reviews and summarizes the current methods for improving the wear resistance of titanium alloys through surface treatment, as shown in Figure 15. Table 2 provides a comprehensive comparison of the various methods. Through the review of this article, some conclusions can be summarized as follows.
  • In terms of economic benefits and the ease of use of equipment, the methods of pulse magnetic field treatment and additive manufacturing technology are relatively cheaper and more economical and the equipment preparation time is shorter.
  • Plasma spraying technology and laser cladding technology can form multiple coatings, but the roughness of the coating formed by plasma spraying technology is relatively high.
  • Additive manufacturing technology and micro arc oxidation technology are relatively environmentally friendly and cause less environmental pollution.

4. Areas for Further Research

Due to the broad application prospects of titanium alloy oil well pipes in the petroleum industry, surface modification methods for improving their wear resistance can be studied in the following areas in the future:
(1)
Micro arc oxidation is expected to remain the preferred method for improving the wear resistance of titanium alloys in the near future, due to its consideration of environmental protection, economic costs, and other factors. This technology also holds great potential for further development.
(2)
With further research on the preparation technology of wear-resistant micro arc oxidation films and the gradual improvement of the relevant theoretical basis, it is believed that micro arc oxidation coatings with controllable friction properties and a good comprehensive performance can be designed and developed at a certain energy, so as to broaden the application range of titanium alloys and promote their industrial application in more fields.

Funding

The present work has been financially supported by the National Natural Science Foundation of China Project (Grant No.: 12102340), Youth Innovation Team Construction in Shaanxi Universities (Shaanxi teaching letter 2023-997-29), Young Scientific Research and Innovation Team of Xi’an Shiyou University (Grant No.: 2019QNKYCXTD14), the Open Research Fund from the State Key Laboratory of Rolling and Automation, Northeastern University (Grant No.: 2022RALKFKT009), the Tribology Science Fund of State Key Laboratory of Tribology in Advanced Equipment (Grant No.: SKLTKF22B10), the Opening Project Fund of Materials Service Safety Assessment Facilities (Grant No.: MSAF-2021-101), Henan International Joint Laboratory of Dynamics of Impact and Disaster of Engineering Structures, Nanyang Institute of Technology (Grant No.: LDIDES-KF2022-02-02), and the China Scholarship Council Foundation (Grant No.: 202208615046).

Conflicts of Interest

Yazhou Chen, Honggang Zhang, Bitao Wang, Jianyong Huang, Meihong Zhou were employed by PetroChina Changqing Oilfifield Company. Ruifan Liu was employed by the company Shaanxi Coal Industry New Energy Technology Co., Ltd. Jiangtao Ji was employed by the company China Railway First Survey and Design Institute Group Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Lu, N.Y.; Fan, J.C.; Liu, D.; Liu, S.J.; Wen, M.; Liang, Z.W. Simulation study on leakage acoustic source characteristics of gas well tubing and casing. China Pet. Mach. 2017, 45, 79–82. [Google Scholar]
  2. Lin, N.M.; Liu, Q.; Zou, J.J.; Li, D.L.; Yuan, S.; Wang, Z.H.; Tang, B. Surface damage mitigation of Ti6Al4V alloy via thermal oxidation for oil and gas exploitation application: Characterization of the microstructure and evaluation of the surface performance. RSC Adv. 2017, 7, 13517–13535. [Google Scholar] [CrossRef]
  3. Wang, Y.M.; Jiang, B.L.; Lei, T.Q.; Guo, L.X. Microarc oxidation coatings formed on Ti6Al4V in Na2SiO3 system solution: Microstructure, mechanical and tribological properties. Surf. Coat. Technol. 2006, 201, 82–89. [Google Scholar] [CrossRef]
  4. Budinski, K.G. Tribological properties of titanium alloys. Wear 1991, 151, 203–217. [Google Scholar] [CrossRef]
  5. Yerokhin, A.L.; Leyland, A.; Matthews, A. Kinetic aspects of aluminium titanate layer formation on titanium alloys by plasma electrolytic oxidation. Appl. Surf. Sci. 2002, 200, 172–184. [Google Scholar] [CrossRef]
  6. Liu, Q.; Song, S.Y.; Li, D.J.; Bai, Q. Research and development of titanium alloy OCTG application in energy industry. Oil Field Equip. 2014, 43, 88–94. [Google Scholar]
  7. Wu, X.Y.; Zhang, H.; Xu, X.J.; Zhang, Z.S.; Guo, S.J. Application of titanium alloy in oil & gas exploration and development. Petrochem. Ind. Appl. 2016, 35, 105–113. [Google Scholar]
  8. Zhao, H.; Wang, B.T.; Du, C.Y.; Wang, S.R.; Meng, F.L.; Zhang, X. Research progress of surface modification technology of TC4 titanium alloy. J. Shenyang Ligong Univ. 2017, 2, 74–77. [Google Scholar]
  9. Liu, W.; Blawert, C.; Zheludkevich, M.L.; Lin, Y.; Talha, M.; Shi, Y.; Chen, L. Effects of graphene nanosheets on the ceramic coatings formed on Ti6Al4V alloy drill pipe by plasma electrolytic oxidation. J. Alloys Compd. 2019, 789, 996–1007. [Google Scholar] [CrossRef]
  10. Prabu, S.S.; Prathiba, S.; Asokan, M.A.; Akhil, V.S. Effect of titanium on dry sliding wear behavior of hot forged P/M low alloy steels. Mater. Today Proc. 2020, 22, 1380–1389. [Google Scholar] [CrossRef]
  11. Dong, Z.H.; Zhang, W.; Kang, H.W.; Xie, Y.J.; Ebrahimnia, M.; Peng, X. Surface hardening of laser melting deposited 12CrNi2 alloy steel by enhanced plasma carburizing via hollow cathode discharge. Surf. Coat. Technol. 2020, 397, 125976. [Google Scholar] [CrossRef]
  12. Yang, Y.H.; Wang, M.Q.; Chen, J.C.; Dong, H. Microstructure and Mechanical Properties of Gear Steels After High Temperature Carburization. J. Iron Steel Res. 2013, 20, 140–145. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Zhang, W.; Sun, W.; Li, Z.J.; Liu, Z. Preparation of Metakaolin Based Geopolymer and Its Three-dimensional Pore Structural Characterization. J. Wuhan Univ. Technol.-Mater. 2015, 30, 550–555. [Google Scholar] [CrossRef]
  14. Guo, L.N.; Chen, J.; Pan, J.D. Double Glow Plasma Hydrogen-Free Carburizing on the Surface of Purity Titanium. Adv. Mater. Res. 2011, 314–316, 108–111. [Google Scholar] [CrossRef]
  15. Jiang, Y.; Wu, Q.; Wang, Y.F.; Zhao, J.X.; Gong, J.M. Suppression of hydrogen absorption into 304L austenitic stainless steel by surface low temperature gas carburizing treatment. Int. J. Hydrogen Energy 2019, 44, 24054–24064. [Google Scholar] [CrossRef]
  16. Dewan, M.A.R.; Zhang, G.; Ostrovski, O. Carbothermal Reduction of Titania in Different Gas Atmospheres. Metall. Mater. Trans. B 2009, 40, 62–69. [Google Scholar] [CrossRef]
  17. Kusmanov, S.A.; Tambovskii, I.V.; Kusmanova, I.A.; Belkin, P.N. Improving the Wear Resistance of VT22 Titanium Alloy by Anodic Plasma Electrolytic Boriding. Surf. Eng. Appl. Electrochem. 2021, 57, 419–424. [Google Scholar] [CrossRef]
  18. Liu, Z.; Peng, Y.W.; Chen, C.M.; Gong, J.M.; Jiang, Y. Effect of surface nanocrystallization on low-temperature gas carburization for AISI 316L austenitic stainless steel. Int. J. Press. Vessel. Pip. 2020, 182, 104053. [Google Scholar] [CrossRef]
  19. Yang, C.; Liu, J.; Ma, Y.Q.; Hong, L.; Wang, H. Microstructure and Wear Resistance of Low Pressure Nitrided Layer on TC4 Titanium Alloy Surface. Mater. Mech. Eng. 2016, 40, 98–101. [Google Scholar]
  20. Peng, M.J.; Duan, Y.H.; Ma, L.S.; Shu, B.P. Characteristics of surface layers on Ti6Al4V alloy borided with CeO2 near the transition temperature. J. Alloys Compd. 2018, 769, 1–9. [Google Scholar] [CrossRef]
  21. Makuch, N. Influence of nickel silicides presence on hardness, elastic modulus and fracture toughness of gas-borided layer produced on Nisil-alloy. Trans. Nonferrous Met. Soc. China 2021, 31, 764–778. [Google Scholar] [CrossRef]
  22. Ma, L.S.; Duan, Y.H.; Li, P. Microstructure, growth kinetics and some mechanical properties of boride layers produced on pure titanium by molten-salt boriding. J. Mater. Eng. Perform. 2017, 26, 4544–4555. [Google Scholar] [CrossRef]
  23. Makuch, N.; Dziarski, P.; Kulka, M.; Piasecki, A.; Tulinski, M.; Majchrowski, R. Influence of niobium and molybdenum addition on microstructure and wear behavior of laser-borided layers produced on Nimonic 80A-alloy. Trans. Nonferrous Met. Soc. China 2019, 29, 322–337. [Google Scholar] [CrossRef]
  24. Atar, E.; Kayali, E.S.; Cimenoglu, H. Characteristics and wear performance of borided Ti6Al4V alloy. Surf. Coat. Technol. 2008, 202, 4583–4590. [Google Scholar] [CrossRef]
  25. Si, J.Y.; Wu, X.L.; Xia, K.N.; Zhang, J. Equal channel angular extrusion of TB2 alloy under different die designs by finite element method. Rare Met. Mater. Eng. 2014, 43, 1577–1581. [Google Scholar] [CrossRef]
  26. Wang, Z.M.; Chen, Z.Y.; Zhan, C.K.; Kuang, L.J.; Shao, J.B.; Wang, R.K.; Liu, C.M. Quasi-static and dynamic forced shear deformation behaviors of Ti-5Mo-5V-8Cr-3Al alloy. Mater. Sci. Eng. A 2017, 691, 51–59. [Google Scholar] [CrossRef]
  27. Li, P.; Liu, D.; Bao, W.Z.; Ma, L.S.; Duan, Y.H. Surface characterization and diffusion model of pack borided TB2 titanium alloy. Ceram. Int. 2018, 44, 18429–18437. [Google Scholar] [CrossRef]
  28. Huang, Y.G.; Chen, J.R.; Zhang, X.H.; Wang, H.Q.; Fang, Z.; Li, Q.Y. Effects of La2O3 on electrolytic boronising of titanium. Surf. Eng. 2015, 31, 570–574. [Google Scholar] [CrossRef]
  29. Zhu, Y.S.; Yin, Y.X.; Wu, J.; Liu, Y.F.; Lu, W.Z.; Zuo, D.W.; Xiao, H.P.; Cao, D.W.; Ko, T.J. Effect of RE on accelerating the kinetics of boride layer growth on titanium alloy. J. Alloys Compd. 2020, 844, 156091. [Google Scholar] [CrossRef]
  30. Duan, Y.H.; Liu, D.; He, B.L.; Ma, L.S.; Hu, Y.Y.; Li, X.Q. Experimental investigations of TB2 alloy by pack boriding with rare-earth oxides. Philos. Mag. Lett. 2019, 98, 521–526. [Google Scholar] [CrossRef]
  31. Qu, D.Y.; Liu, D.; Wang, X.Y.; Duan, Y.H.; Peng, M.J. Corrosion and wear properties of TB2 titanium alloy borided by pack boriding with La2O3. Trans. Nonferrous Met. Soc. China 2022, 32, 868–881. [Google Scholar] [CrossRef]
  32. Bloyce, A.; Qi, P.Y.; Dong, H.; Bell, T. Surface modification of titanium alloys for combined improvements in corrosion and wear resistance. Surf. Coat. Technol. 1998, 107, 125–132. [Google Scholar] [CrossRef]
  33. Yang, Y.F.; Qin, Y.; Yang, Y.; Wu, M.X.; Yang, G. Enhancing the wear resistance of a cemented carbide/titanium alloy under magnetofluid lubrication via the magnetic response. Wear Int. J. Sci. Technol. Frict. Lubr. Wear 2022, 500–501, 204370. [Google Scholar] [CrossRef]
  34. Awasthi, S.; Gaur, J.K.; Bobji, M.S. Advanced ferrogels with high magnetic response and wear resistance using carbon nanotubes. J. Alloys Compd. Interdiscip. J. Mater. Sci. Solid-State Chem. Phys. 2020, 848, 156259. [Google Scholar] [CrossRef]
  35. Jin, J.; Duan, H.J.; Li, X.H. The influence of plasma nitriding on microstructure and properties of CrN and CrNiN coatings on Ti6Al4V by magnetron sputtering. Vacuum 2016, 136, 112–120. [Google Scholar] [CrossRef]
  36. Trivedi, P.; Gupta, P.; Srivastava, S.; Jayaganthan, R.; Chandra, R.; Roy, P. Characterization and in vitro biocompatibility study of Ti-Si-N nanocomposite coatings developed by using physical vapor deposition. Appl. Surf. Sci. A J. Devoted Prop. Interfaces Relat. Synth. Behav. Mater. 2014, 293, 143–150. [Google Scholar] [CrossRef]
  37. Yuan, S.; Lin, N.M.; Zou, J.J.; Lin, X.Z.; Liu, Z.Q.; Yu, Y.; Wang, Z.X.; Zeng, Q.F.; Chen, W.G.; Tian, L.H.; et al. In-Situ fabrication of gradient titanium oxide ceramic coating on laser surface textured Ti6Al4V alloy with improved mechanical property and wear performance. Vacuum 2020, 176, 109327. [Google Scholar] [CrossRef]
  38. Yang, H.L.; Zheng, Y.; Wang, Y. Bio-tribology properties of imitation shark skin morphology on Ti6A14V surface. Hot Work. Technol. 2016, 45, 119–122. [Google Scholar]
  39. Kashyap, V.; Ramkumar, P. Feasibility study of micro-groove cross hatched surface texturing on Ti6Al4V for improved biotribological performance in metal-onpolymer hip implant. Tribol. Mater. Surf. Interfaces 2019, 13, 150–160. [Google Scholar] [CrossRef]
  40. Wang, D.W.; Mo, J.L.; Zhu, Z.Y.; Ouyang, H.; Zhu, M.H.; Zhou, Z.R. Debris trapping and space-varying contact via surface texturing for enhanced noise performance. Wear 2018, 396–397, 86–97. [Google Scholar] [CrossRef]
  41. Li, X.L.; Yue, W.; Huang, F.; Kang, J.J.; Zhu, L.N.; Tian, B. Tribological behaviour of textured titanium under abrasive wear. Surf. Eng. 2018, 35, 378–386. [Google Scholar] [CrossRef]
  42. Wu, Z.; Xing, Y.Q.; Huang, P.; Liu, L. Tribological properties of dimple-textured titanium alloys under dry sliding contact. Surf. Coat. Technol. 2017, 309, 21–28. [Google Scholar] [CrossRef]
  43. Zhang, H.; Qin, L.G.; Hua, M.; Dong, G.N.; Chin, K.S. A tribological study of the petaloid surface texturing for Co-Cr-Mo alloy artificial joints. Appl. Surf. Sci. 2015, 332, 557–564. [Google Scholar] [CrossRef]
  44. Koszela, W.; Pawlus, P.; Reizer, R.; Liskiewicz, T. The combined effect of surface texturing and DLC coating on the functional properties of internal combustion engines. Tribol. Int. 2018, 127, 470–477. [Google Scholar] [CrossRef]
  45. Yuan, S.; Lin, N.M.; Zou, J.J.; Liu, Z.Q.; Wang, Z.X.; Tian, L.H.; Qin, L.; Zhang, H.X.; Wang, Z.H.; Tang, B.; et al. Effect of laser surface texturing(LST) on tribological behavior of double glow plasma surface zirconizing coating on Ti6Al4V alloy. Surf. Coat. Technol. 2019, 368, 97–109. [Google Scholar] [CrossRef]
  46. Wu, C.; Yang, K.; Chen, Y.; Ni, J.; Yao, L.D.; Li, X.L. Investigation of friction and vibration performance of lithium complex grease containing nano-particles on rolling bearing. Tribol. Int. 2021, 155, 45–48. [Google Scholar] [CrossRef]
  47. Jiang, C.P.; Zhang, J.; Chen, Y.N.; Hou, Z.M.; Zhao, Q.Y.; Li, Y.; Zhu, L.X.; Zhang, F.Y.; Zhao, Y.Q. On enhancing wear resistance of titanium alloys by laser cladded WC-Co composite coatings. Int. J. Refract. Met. Hard Mater. 2022, 107, 105902. [Google Scholar] [CrossRef]
  48. Wang, K.P.; Wu, H.C.; Wang, H.D.; Liu, Y.H. Superior extreme pressure properties of different layer LDH nanoplatelets used as boundary lubricants. Appl. Surf. Sci. 2020, 530, 15–23. [Google Scholar] [CrossRef]
  49. Fan, S.G.; Gao, C.P.; Yang, G.B.; Zhang, Y.J.; Zhang, C.L.; Song, N.N.; Zhang, S.M.; Zhang, P.Y.; Zhang, Z.J.; Ke, S.J. Enhancing wear-resistance of Cu nanowires filled polyetheretherketone-based composites under water lubrication conditions via magnetic response to tribofilm. Tribol. Int. 2021, 153, 106601. [Google Scholar] [CrossRef]
  50. Ni, J.; Cui, Z.; Wu, C.; Sun, J.B.; Zhou, J.H. Evaluation of MQL broaching AISI 1045 steel with sesame oil containing nano-particles under best concentration. J. Clean. Prod. 2021, 320, 443–456. [Google Scholar] [CrossRef]
  51. Kossman, S.; Coorevits, T.; Iost, A.; Chicot, D. A new approach of the Oliver and Pharr model to fit the unloading curve from instrumented indentation testing. J. Mater. Res. 2017, 32, 2230–2240. [Google Scholar] [CrossRef]
  52. Wu, B.; Song, H.; Zhang, Q.Q.; Hu, X.G. Controllable synthesis and friction reduction of ZnFe2O4@C microspheres with diverse core-shell architectures. Tribol. Int. 2021, 153, 106614. [Google Scholar] [CrossRef]
  53. Zhang, Z.; Zhao, Y.; Shan, J.; Wu, A.; Tang, X. Evolution behavior of liquid film in the heataffected zone of laser cladding non-weldable nickel-based superalloy. J. Alloys Compd. 2021, 863, 158463. [Google Scholar] [CrossRef]
  54. Wang, Y.; Li, C.G.; Tian, W.; Yang, Y. Laser surface remelting of plasma sprayed nanostructured Al2O3-13wt%TiO2 coatings on titanium alloy. J. Alloys Compd. 2010, 503, 127–132. [Google Scholar] [CrossRef]
  55. Yu, G.; Li, Z.; Li, S.; Zhang, Q.; Hua, Y.; Liu, H.; Zhao, X.; Dhaidhai, D.T.; Li, W.; Wang, X. The select of internal architecture for porous Ti alloy scaffold: A compromise between mechanical properties and permeability. Mater. Des. 2020, 192, 108754. [Google Scholar] [CrossRef]
  56. Herzog, D.; Seyda, V.; Wycisk, E.; Emmelmann, C. Additive manufacturing of metals. Acta Mater. 2016, 117, 371–392. [Google Scholar] [CrossRef]
  57. Zhao, Y.; Mei, H.; Chang, P.; Yang, Y.; Huang, W.; Liu, Y.; Cheng, L.; Zhang, L. 3DPrinted Topological MoS2/MoSe2 Heterostructures for Macroscale Superlubricity. ACS Appl. Mater. Interfaces 2021, 13, 34984–34995. [Google Scholar] [CrossRef] [PubMed]
  58. Huang, Q.; Shi, X.; Xue, Y.; Zhang, K.; Wu, C. Optimization of bionic textured parameter to improve the tribological performance of AISI 4140 self-lubricating composite through response surface methodology. Tribol. Int. 2021, 161, 107104. [Google Scholar] [CrossRef]
  59. Li, S.; An, Y.; Zhao, X.; Zhou, H.; Chen, J.; Hou, G. Bioinspired Smart Coating with Superior Tribological Performance. ACS Appl. Mater. Interfaces 2017, 9, 16745–16749. [Google Scholar] [CrossRef]
  60. Harris, K.L.; Pitenis, A.A.; Sawyer, W.G.; Krick, B.A.; Blackman, G.S.; Kasprzak, D.J.; Junk, C.P. PTFE Tribology and the Role of Mechanochemistry in the Development of Protective Surface Films. Macromolecules 2015, 48, 3739–3745. [Google Scholar] [CrossRef]
  61. Zuo, Z.; Jin, X.; Li, C.; Zhang, Z.; Yang, Y. Chemisorption mechanism of defluorinated fluorine on bcc Fe surface during formation of PTFE transfer film. Appl. Surf. Sci. 2021, 567, 150777. [Google Scholar] [CrossRef]
  62. Xu, F.; Gong, D. Improved the elevated temperature mechanical properties of Al-Si alloy deposited with Al-Si coating by magnetron sputtering. Vacuum 2018, 150, 1–7. [Google Scholar] [CrossRef]
  63. Wu, B.H.; Leng, Y.X.; Huang, N.; Yang, W.M.; Li, X.Y. The plasma characteristics in high power pulsed impulsed magnetron sputtering (hipims) and its effect on films properties. Surf. Technol. 2018, 47, 245–255. [Google Scholar]
  64. Khrushchov, M.M.; Levin, I.S.; Reilyanu, M.D.; Obraztsova, E.A.; Atamanov, M.V. Structure and tribological behavior of chromium-carbon coatings obtained by magnetron sputtering. Phys. Met. Metallogr. 2019, 120, 204–209. [Google Scholar] [CrossRef]
  65. Kang, S.H.; Tu, W.B.; Han, J.X.; Li, Z.; Cheng, Y.L. A significant improvement of the wear resistance of Ti-6Al-4V alloy by a combined method of magnetron sputtering and plasma electrolytic oxidation (PEO). Surf. Coat. Technol. 2019, 358, 879–890. [Google Scholar] [CrossRef]
  66. Chi, G.F.; Yi, D.Q.; Liu, H.Q. Effect of roughness on electrochemical and pitting corrosion of Ti-6Al-4V alloy in 12 wt.% HCl solution at 35 °C. J. Mater. Res. Technol. 2020, 9, 1162–1174. [Google Scholar] [CrossRef]
  67. Cheng, Z.H.; Yang, W.; Xu, D.P.; Wu, S.K.; Yao, X.F.; Lv, Y.K.; Chen, J. Improvement of high temperature oxidation resistance of micro arc oxidation coated AlTiNbMo0.5Ta0.5Zr high entropy alloy. Mater. Lett. 2020, 262, 127192. [Google Scholar] [CrossRef]
  68. Correa, D.R.N.; Rocha, L.A.; Ribeiro, A.R.; Gemini-Piperni, S.; Archanjo, B.S.; Achete, C.A.; Werckmann, J.; Afonso, C.R.M.; Shimabukuro, M.; Doi, H.; et al. Growth mechanisms of Ca- and P-richMAO films in Ti-15Zr-xMo alloys for osseointegrative implants. Surf. Coat. Technol. 2018, 344, 373–382. [Google Scholar] [CrossRef]
  69. Guo, Q.; Xu, D.; Yang, W.; Guo, Y.; Yang, Z.; Li, J.; Gao, P. Synthesis, corrosion, and wear resistance of a black microarc oxidation coating on pure titanium. Surf. Coat. Technol. 2020, 386, 125454. [Google Scholar] [CrossRef]
  70. Aliofkhazraei, M.; Rouhaghdam, A.S. Wear and coating removal mechanism of alumina/titania nanocomposite layer fabricated by plasma electrolysis. Surf. Coat. Technol. 2011, 205, S57–S62. [Google Scholar] [CrossRef]
  71. Qi, Y.M.; Peng, Z.J.; Liang, J. Fabrication and wear resistance of hard micro arc oxidation coatings on Ti alloys. Surf. Technol. 2019, 48, 81–88. [Google Scholar]
  72. Chen, L.; Wei, K.; Yao, Q.; Tao, L.; Bei, C.; Liao, B.; Xue, W. Characterization of plasma electrolytic oxidation film on biomedical high niobiumcontaining β-titanium alloy. Surf. Coat. Technol. 2018, 352, 295–301. [Google Scholar] [CrossRef]
  73. Li, Q.B.; Yang, W.B.; Liu, C.C.; Wang, D.A.; Liang, J. Correlations between the growth mechanism and properties of micro-arc oxidation coatings on titanium alloy: Effects of electrolytes. Surf. Coat. Technol. 2017, 316, 162–170. [Google Scholar] [CrossRef]
  74. Yerokhin, A.L.; Nie, X.; Leyland, A.; Matthews, A. Characterization of oxide films produced by plasma electrolytic oxidation of a Ti-6Al-4V alloy. Surf. Coat. Technol. 2000, 130, 195–206. [Google Scholar] [CrossRef]
  75. Quintero, D.; Galvis, Q.; Calderon, J.A.; Gomez, M.A.; Castano, J.G.; Echeverria, F.; Habazaki, H. Control of the physical properties of anodic coatings obtained by plasma electrolytic oxidation on Ti6Al4V alloy. Surf. Coat. Technol. 2015, 283, 210222. [Google Scholar] [CrossRef]
  76. Lan, N.; Yang, W.; Gao, W.; Guo, P.; Zhao, C.; Chen, J. Characterization of ta-C film on micro arc oxidation coated titanium alloy in simulated seawater. Diam. Relat. Mater. 2021, 117, 108483. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of surface carburizing/nitrogen/boron technology.
Figure 1. Schematic diagram of surface carburizing/nitrogen/boron technology.
Coatings 14 00786 g001
Figure 2. Wear morphology of TB2 alloy and borated sample (1100 °C, 20 h, containing 4 wt.% La2O3) tested at different temperatures (adapted from [31]).
Figure 2. Wear morphology of TB2 alloy and borated sample (1100 °C, 20 h, containing 4 wt.% La2O3) tested at different temperatures (adapted from [31]).
Coatings 14 00786 g002
Figure 3. Schematic diagram of pulse magnetic field treatment.
Figure 3. Schematic diagram of pulse magnetic field treatment.
Coatings 14 00786 g003
Figure 4. (a) Wear volumes of TC4, (b,c) magnetic force microscope (MFM) images of untreated and MT samples (adapted from [33]).
Figure 4. (a) Wear volumes of TC4, (b,c) magnetic force microscope (MFM) images of untreated and MT samples (adapted from [33]).
Coatings 14 00786 g004
Figure 5. Schematic diagram of Laser cladding technology.
Figure 5. Schematic diagram of Laser cladding technology.
Coatings 14 00786 g005
Figure 6. (a) Cross-section microstructure of WC-Co composite coatings, (b) microhardness distribution in cross-section of WC-Co composite coatings, and (c) wear rate distribution of the three body impact abrasion test (adapted from [47]).
Figure 6. (a) Cross-section microstructure of WC-Co composite coatings, (b) microhardness distribution in cross-section of WC-Co composite coatings, and (c) wear rate distribution of the three body impact abrasion test (adapted from [47]).
Coatings 14 00786 g006
Figure 7. Schematic diagram of plasma spraying technology.
Figure 7. Schematic diagram of plasma spraying technology.
Coatings 14 00786 g007
Figure 8. Microhardness of the as-sprayed and laser-remelted coatings, and the corresponding surface micrographs (adapted from [54]).
Figure 8. Microhardness of the as-sprayed and laser-remelted coatings, and the corresponding surface micrographs (adapted from [54]).
Coatings 14 00786 g008
Figure 9. Schematic diagram of additive manufacturing technology.
Figure 9. Schematic diagram of additive manufacturing technology.
Coatings 14 00786 g009
Figure 10. (a) TEM image of zirconia ball wear surface, (b) element composition at the selected point in (a), (c) high-resolution TEM image of zirconia ball wear surface, and (dg) corresponding image and IFFT image of zirconia ball wear surface [61].
Figure 10. (a) TEM image of zirconia ball wear surface, (b) element composition at the selected point in (a), (c) high-resolution TEM image of zirconia ball wear surface, and (dg) corresponding image and IFFT image of zirconia ball wear surface [61].
Coatings 14 00786 g010
Figure 11. Schematic diagram of magnetron sputtering technology.
Figure 11. Schematic diagram of magnetron sputtering technology.
Coatings 14 00786 g011
Figure 12. Three-dimensional topography of wear scars of different samples after wear testing: (a) Ti6Al4V alloy; (b) magnetron sputtering aluminized Ti6Al4V alloy; (c) Na2SiO3, 4 min; (d) Na2SiO3, 15 min; (e) NaAlO2, 1 min; (f) NaAlO2, 4 min (adapted from [65]).
Figure 12. Three-dimensional topography of wear scars of different samples after wear testing: (a) Ti6Al4V alloy; (b) magnetron sputtering aluminized Ti6Al4V alloy; (c) Na2SiO3, 4 min; (d) Na2SiO3, 15 min; (e) NaAlO2, 1 min; (f) NaAlO2, 4 min (adapted from [65]).
Coatings 14 00786 g012
Figure 13. Schematic diagram of micro arc oxidation technology.
Figure 13. Schematic diagram of micro arc oxidation technology.
Coatings 14 00786 g013
Figure 14. (a) Open circuit potential of TC4 and three coatings, (b) Nyquist plots of TC4 and three coatings, (c) polarization curves of TC4 and three coatings, and (d) relationship between the friction coefficient and sliding time of TC4 with three different coatings (adapted from [76]).
Figure 14. (a) Open circuit potential of TC4 and three coatings, (b) Nyquist plots of TC4 and three coatings, (c) polarization curves of TC4 and three coatings, and (d) relationship between the friction coefficient and sliding time of TC4 with three different coatings (adapted from [76]).
Coatings 14 00786 g014
Figure 15. Surface treatment method summary of improving wear resistance of titanium alloys.
Figure 15. Surface treatment method summary of improving wear resistance of titanium alloys.
Coatings 14 00786 g015
Table 1. Comparison of related parameters of high-performance drill pipes.
Table 1. Comparison of related parameters of high-performance drill pipes.
Material TypeMaterial GradeYield Strength/MPaTensile Strength/kNGravity/NStrength to Gravity Ratio
SteelUD-165(VM-150)113745915039901
SteelV-150103441745039819
SteelZ-1409658955039765
SteelS-13593137565039737
Aluminum alloyAl-Zn-MgI32517783189557
Aluminum alloyAl-Zn-MgII48026263189823
Aluminum alloyAl-Cu-Mg-Si-FeIII34018603189583
Aluminum alloyAl-Zn-MgIV35019143189600
Titanium alloyTi6Al4V951382531301213
Table 2. Comparison of various surface technologies.
Table 2. Comparison of various surface technologies.
Comparison ContentSurface Carburizing/Nitrogen/Boron TechnologyPulse Magnetic Field TreatmentLaser Cladding TechnologyPlasma Spraying TechnologyAdditive Manufacturing TechnologyMagnetron Sputtering TechnologyMicro Arc Oxidation
Equipment and processExpensive equipment and complex processSimple equipment and complex processComplex equipment and processExpensive equipment and simple processSimple equipment and processExpensive equipment and complex processSimple equipment and process
Base materialMetallic materialsMetallic materialsMetals, ceramics and compositesMetal, ceramic and organic materialsMetallic materialsMetallic materialsMetallic materials
Matrix temperature800 °C900 °CLower100~260 °C500 °C40~50 °C<300 °C
Preparation timeA few hoursA few minutesA few minutesA few minutes to ten minutesA few secondsAbout 4 hA few minutes to ten minutes
Coating thicknessThin, 20~30 μm5 mm0.5~2.0 mmDozens to hundreds of microns0.1 mmAbout 6 nmThin, more than ten microns
Microhardness of coatingHighHighVery highHigherHighHighHigh
Coating roughnessLowerLowLowHigherLowLowLow
Substrate suitabilitySuitable for simple small size partsFinal partsAlloy powderSimple parts with wide size rangeHighly complex and functional partsPrecision instrumentSuitable for complex parts
Type of coatingSingleSingleVarietyVarietySingleSingleSingle
Working environmentVacuum state, environment-friendlyVacuum state, environment-friendlyVacuum state, environment-friendlyInert gas protection, loud noiseNo pollution to the environmentVacuum stateNo pollution to the environment
Preparation costHigh, suitable for small-scale productionLow, suitable for mass productionLow, widely usedLow, suitable for mass productionHigh, suitable for small-scale productionLow, suitable for mass productionLow, suitable for mass production
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

Chen, Y.; Zhang, H.; Wang, B.; Huang, J.; Zhou, M.; Wang, L.; Xi, Y.; Jia, H.; Xu, S.; Liu, H.; et al. A Review of Research on Improving Wear Resistance of Titanium Alloys. Coatings 2024, 14, 786. https://doi.org/10.3390/coatings14070786

AMA Style

Chen Y, Zhang H, Wang B, Huang J, Zhou M, Wang L, Xi Y, Jia H, Xu S, Liu H, et al. A Review of Research on Improving Wear Resistance of Titanium Alloys. Coatings. 2024; 14(7):786. https://doi.org/10.3390/coatings14070786

Chicago/Turabian Style

Chen, Yazhou, Honggang Zhang, Bitao Wang, Jianyong Huang, Meihong Zhou, Lei Wang, Yuntao Xi, Hongmin Jia, Shanna Xu, Haitao Liu, and et al. 2024. "A Review of Research on Improving Wear Resistance of Titanium Alloys" Coatings 14, no. 7: 786. https://doi.org/10.3390/coatings14070786

APA Style

Chen, Y., Zhang, H., Wang, B., Huang, J., Zhou, M., Wang, L., Xi, Y., Jia, H., Xu, S., Liu, H., Wen, L., Xiao, X., Liu, R., & Ji, J. (2024). A Review of Research on Improving Wear Resistance of Titanium Alloys. Coatings, 14(7), 786. https://doi.org/10.3390/coatings14070786

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