You are on page 1of 10

Proceedings of the International Conference on Mechanical Engineering and Mechatronics Ottawa, Ontario, Canada, 15-17 August 2012 Paper

No. 83 (The number assigned by the OpenConf System)

A Critical Review on Tool wear in Titanium machining


Narasimhulu Andriya, Venkateswara Rao P,
Indian Institute of Technology Delhi Department of Mechanical Enginering, New Delhi-110016, India narasimha.iitdelhi@gmail.com; pvrao@mech.iitd.ac.in;

Sudarsan Ghosh
Indian Institute of Technology Delhi, Department of Mechanical Engineering Hauz Khas, New Delhi-110016, India sudarsan.ghosh@gmail.com
Abstract - Titanium machining is attaining importance for machining of titanium alloys as it has several benefits over other materials. The major concern while machining of Titanium alloys is the rapid tool wear that takes place in almost all varieties of tool materials especially while machining at higher cutting speeds. There are several issues regarding tool wear, which should be understood and dealt with, to achieve successful performance of the process. Researchers have worked upon several aspects related to titanium machining. The present work is an effort to review some of these works and to understand the key issues related to process performance. The review reveals the effect of tool wear on the types of tool material and different types of coatings, cutting edge geometry and cutting parameters. The present work also aims to find out certain areas that can be taken up for further research in minimizing the tool wear during machining of Titanium alloys. Keywords: Titanium alloys, tool wear, machining.

1. Introduction
Titanium is a metal showing a high strength-weight ratio which is maintained even at elevated temperatures, and it has exceptional corrosion resistance. These characteristics were the main reasons for the rapid growth of the titanium industry over the last 70 years. The major application of the material is in the aerospace industry, both in airframes and engine components manufacturing. Non-aerospace applications take advantage mainly of their excellent strength properties, for example in steam turbine blades, superconductor missiles etc., or corrosion resistance, for example in marine services, chemical, petro-chemical, electronics industry, biomedical industry etc., However, despite the increased usage and production of titanium, they are expensive when compared to many other metals, because of the complexity of the extraction process, difficulty of melting and problems during fabrication and machining. Near net-shape methods such as casting, isothermal forging, and powder metallurgy have been introduced to reduce the cost of titanium components. However, most titanium parts are still manufactured by conventional machining methods. Virtually all types of machining operations, such as turning, milling, drilling, reaming, tapping, sawing, and grinding, are employed in producing aerospace components. For the manufacturing of gas turbine engines, turning and drilling are the major machining operations, whilst in airframe production; end milling and drilling are amongst the most important machining operations. The machinability of titanium and its alloys is generally considered to be poor owing to several inherent properties of the materials. Titanium is chemically very reactive and, therefore, has a tendency to weld to the cutting tool during machining, thus leading to chipping and premature tool failure. Its low thermal conductivity increases the temperature at the tool/workpiece interface, which affects the tool life adversely. Additionally, its high strength maintained at elevated temperature and its low modulus of
83-1

elasticity further impairs its machinability (Hong, Riga et al. 1993). The poor machinability of titanium and its alloys have led many large companies (for example Rolls-Royce and GE) to invest large sums of money in developing techniques to minimize machining cost. Reasonable production rates and excellent surface quality can be achieved with conventional machining methods if the unique characteristics of the metal and its alloys are taken into account.

2. Basic problems in machining of Titanium and its alloys


Machinability is defined as the ease (or difficulty) with which a material can be machined under a given set of operating conditions including cutting speed, feed rate and depth of cut. It can also be described as a measure of the response of material to be machined with a given tool material resulting in an acceptable tool life and at the same time providing good surface finish and acceptable functional characteristics of the components. Machinability rating depends on tool life, surface finish and power consumed during the machining operation. Component forces and chip shape also provide a good assessment of the machinability of material(Trent 2000). Machinability of any material is mainly assessed by measuring the tool life, surface generated and the magnitude of various cutting force components during machining operations. The favourable responses of all these measures lead towards the improvement of machining performance by selecting the appropriate process parameters, cutting tool and cutting fluid. Reasons for poor machinability Many authors (Komanduri and Von Turkovich 1981; Komanduri and Reed Jr 1983; Ezugwu and Wang 1997; Jawaid, Che-Haron et al. 1999; Yang and Richard Liu 1999; Hong, Ding et al. 2001; Hong, Markus et al. 2001; Ribeiro, Moreira et al. 2003; Li, Yang et al. 2004; Abele 2008; Jianxin, Yousheng et al. 2008; G. A. Ibrahim 2009; Bermingham, Kirsch et al. 2011; Li, Zhao et al. 2011) claim individually some of the following points as the reasons for the poor machinability of titanium: a. The high strength is maintained even at elevated machining temperatures and this opposes the plastic deformation needed to form a chip. b. Titaniums chip is very thin with an unusually small contact area with the tool [one third that of the contact area of steel at the same feed rate and depth of cut]. This causes high stresses on the tool, although cutting forces are reported to be similar to steel and hence the power consumption in machining is approximately the same. c. It has a high coefficient of friction between the chip and the tool face, although it is more or less comparable with that obtained during machining of many hardened steels. d. There is a strong chemical reactivity of titanium at the cutting temperature (>500 0C) with almost all tool materials available. e. Titanium chips are formed by the adiabatic or catastrophic thermoplastic shear process. Titaniums low volumetric specific heat and relatively small contact area along with the presence of a very thin flow zone between the chip and the tool [approximately 8 m compared with 50 m when cutting steels under the same cutting conditions] cause high tool-chip temperatures ranging around 11000C. f. Normally in titanium alloy machining built up edge does not occur that predominantly but some authors have confirmed its presence at low cutting speeds, and this could lead to a poor surface finish in some operations. g. Low modulus of elasticity which cause chatter, deflection and rubbing problems on the workpiece. h. Titanium alloys can have a tendency to ignite during machining because of the high temperatures involved and appropriate care needs to be taken. i. There is a high rate of work hardening, although some researchers like Zaltin and Child and Dalton have reported that it work-hardens to a lesser extent than steel.

83-2

3. Common tool materials, coatings and their geometry used for Titanium alloy machining
When it comes to cutting tools, coated indexable inserts are responsible for the major share of metal removed throughout machine shops in modern times. These are the tools that, since 1970, have continually increased manufacturing productivity and, consequently, material prosperity. With the advent of different types of coating materials, the applicability of these coated tools increased. About 80 percent of the inserts used in machining today are coated cemented carbide grades. These tool inserts earned and maintain that growing market share because of their broad application for removing large amounts of material while sustaining a long tool-life. The high material removal rates and long life of these tools are achieved through a balance of wear resistance and toughness. Relatively thin coatings in the order of 0.001mm. to 0.0178 mm. protect inserts from the heat, corrosion and abrasion that normally shorten their lives. The basic insert the substrate provides most of the required toughness for the cutting tool, while the coating adds wear protection and increased hardness. Most coatings are ceramic by nature, and if an insert was made solely of the coating material, it would be too brittle for general use. The advantage is obtained by correctly combining the most suitable coating material and substrate material. Coating materials: Most coating-materials used on todays inserts are classified as ceramic. Some of the most common coating materials along with their main properties are listed below: Titanium carbonitride - TiCN has very good abrasive wear resistance and good adherence to tungstencarbide substrate; Aluminum oxide (alumina) - Al2O3 provides very good outer thermal and chemical protection for the insert substrate; Titanium nitride - TiN mainly used for the golden color which provides clear wear detection; Zirconium oxide - ZrO2 gives very good thermal and chemical protection for the substrate; By using more than one layer of different coating materials it is possible to combine the benefits that each provides. Many insert grades have three layers of coatings to ensure good adherence between the insert substrate and coatings. Also, sometimes the laminating effect of the coatings provides added strength to the inserts. Coating processes: Two principal coating processes CVD (Chemical Vapor Deposition) and PVD (Physical Vapor Deposition)) are used for indexable inserts to provide cutting edges with fundamentally different properties for machining. The CVD method allows one to form coatings with high adhesion strength to substrate, high density, and compositional uniformity. One characteristic structure of such coatings is equiaxed grains, which adapt better to operations in continuous cutting conditions. However, to perform the gas phase CVD processes, a relatively high temperature and time duration are required. These conditions, in turn, become the reason for the softening of the coated tool and the decrease in their efficiency, especially for heavily loaded cutting operations, intermittent processes, and in cutting tough materials (Che-Haron, Ginting et al. 2007; Blinkov, Anikin et al. 2011). The best tool substrate and coating for machining titanium alloys and super alloys is a submicron substrate that is combined with a physical vapor deposition (PVD) TiAlN coating. The thin, smooth surface of the PVD coating, together with sufficient residual stress, enhances tool resistance to chipping and notching wear, so PVD coatings provide enhanced wear resistance, chemical stability and resistance to built-up edge. Machining problems that were seen in the past that arose from earlier coatings, no longer exist with PVD coatings because of the improved adhesion techniques and the uniformity of the coatings (Bouzakis, Michailidis et al. 2000; Bouzakis, Michailidis et al. 2001; Bouzakis, Hadjiyiannis et al. 2003; Bouzakis, Hadjiyiannis et al. 2004; Klocke, Michailidis et al. 2010; Jaffery and Mativenga 2011). In titanium machining the cutting edges of the tools should be kept sharp without rounding in order to avoid high cutting temperatures and forces. The tools should be replaced immediately when wear occurs (Guimu, Chao et al. 2003; Che-Haron, Ginting et al. 2007). Due to the low plasticity of titanium and the attenuation of the cutting edge by the additional grinding operation, chip breakers should not be used at all. The references and recommendations regarding the wedge angles are not consistent. While the
83-3

rake angles of -5 to 0 are recommended, Lopez et al. (Lpez de lacalle, Prez et al. 2000) showed that clearance angles of less than 18 cause much friction and heat; clearance angles larger than 20 weaken the cutting edge.

4. Various types of tool wear and their reasons


Different tool materials show different responses to different wear mechanisms when machining titanium alloys with notching, crater wear, and flank wear, chipping and catastrophic failure being the prominent failure modes (A.R. Machado 1990; Ezugwu and Wang 1997; Ezugwu, Bonney et al. 2003). However, the crater and the flank wear of the uncoated WC/Co cutting tools result mostly from dissolution-diffusion, while in the case of other tool materials, attrition is the major cause of flank wear (Dearnly PA 1986) Hartung and Kramer (Hartung, Kramer et al. 1982) have reported that most of the commonly used cutting tool materials suffer from chemical instability leading to rapid tool wear when machining titanium alloys and titanium from the chip material adheres to the tool hindering relative sliding at the chip-tool interface. This leads to the formation of a boundary layer of titanium at the interface. Repeated tearing and transport of this titanium layer by the chip underside results in tool material being pulled away causing increased crater wear. According to Hartung (Hartung, Kramer et al. 1982) who conducted experimental studies of tool wear in titanium machining, the main mechanism controlling the crater wear of cutting tool materials in the machining of titanium alloys is fundamentally different from that in the machining of steel and nickel-based alloys. It is suggested that tool wear is greatly reduced when adhesion occurs between the tool and the chip, preventing relative sliding at the tool/chip interface. This adhesion is promoted by chemical reaction at the interface. The thickness of the reaction layer is determined by the balance between the diffusion flux of tool constituents through the layer and the removal of tool constituents through chemical dissolution at the interface between the reaction layer and the titanium chip. In this way, for given cutting conditions, a characteristic thickness of the reaction layer is maintained and tool wear is limited by the rate of dissolution of the reaction layer into the titanium. The existence of a stable reaction layer of TiC on diamond and WC-based tools (the two most wear resistant tool materials) has been demonstrated, and the estimated diffusion flux correlates well with the observed wear rate. Jawaid et al (Jawaid, Che-Haron et al. 1999) have observed that the dominant wear mechanisms of cemented WC-Co tools were dissolution, diffusion and attrition, causing the plucking and pulling-out of carbide from the tool. Abrasion wear mechanisms dominated the wear type at the flank face and tool nose. As the substrates of the PVD-TiN and CVD-TiCN-Al2O3 coated tools were of the same composition, they are subjected to similar wear mechanisms throughout the cutting tests. Both tools experienced a variety of wear mechanisms that accelerated the tool wear when face milling of the Ti-6Al4V. Some of the observed wear mechanisms are delamination of coating, adhesion, attrition, diffusion and plastic deformation and cracks. Coating delamination, galling on the rake face and adhesion of work material at the cutting edge were responsible for the initial wear mechanism for both of the coated tools. Attrition and diffusion wear mechanisms were responsible for the flank and rake face wear of both of the coated tools. Evidence of the diffusion of cobalt and tungsten into the adhered workpiece was found at the flank and rake faces of the tools. The thermal cracks observed were thought to be responsible for the severe chipping and/or flaking of the inserts at both the rake and flank faces (Jawaid, Sharif et al. 2000; Nouari and Ginting 2006). Su et al (Su, He et al. 2006) have observed that flank wear was the dominant failure mode under the different cooling/lubrication conditions, including dry, nitrogen-oil-mist, CCNG(compressed cold nitrogen gas) and CCNGOM (compressed cold nitrogen gas and oil mist) for the cutting conditions of Cutting speed 400 m/min, Feed rate 0.1 mm/rev, Axial depth of cut 5.0mm, Radial depth of cut 1.0mm) and they have also found that TiN/TiC/TiN coating was removed from the cutting edge under all the cooling/lubrication conditions employed. Adhesion of workpiece material onto the flank face of the tool was observed under all the cooling/lubrication conditions. Fig. 1 shows an example of adhered workpiece material on the flank face of the tool, and it indicates that there is a strong bond at the tool workpiece
83-4

interface. Many researchers have reported that the temperature at the cutting edge of carbide tool can be above 8000 C when machining titanium alloy, even at moderate cutting condition (Hartung, Kramer et al. 1982; Narutaki, Murakoshi et al. 1983; Kitagawa, Kubo et al. 1997; Hong and Ding 2001; Zhang, Li et al. 2010). The high temperature and the close contact between the tool and the workpiece provided an ideal environment for diffusion of tool material atoms across the tool-workpiece interface (Dearnly PA 1986; Jawaid, Che-Haron et al. 1999).

Fig. 1. Adhesion of workpiece material on the flank face of tool under CCNG (00 C) condition (Su, He et al. 2006). Venugopal et al (Venugopal, Paul et al.1 2007; Venugopal, Paul et al.2 2007) have found that in machining of titanium alloys crater wear occurred predominantly by adhesion dissolutiondiffusion wear as had been reported earlier, but with cryogenic cooling reduction of such wear mechanism over the domain of the process parameters employed was observed and they have also reported that Edge depression of the cutting tool insert, a major problem in machining of titanium alloys has also been significantly decreased due to effective control of cutting zone temperature by cryogenic cooling. The appearance of crater surface is smooth but the extent of crater wear (crater depth) is rather high indicating adhesiondissolutiondiffusion nature of crater wear. The flank wear region however was not smooth, rather abrasive wear marks were visible. The cutting edges underwent micro and macro fracturing along with plastic deformation. Such wear of cutting edge modified the effective tool geometry due to edge depression of the cutting edge. Jianxin (Jianxin, Yousheng et al. 2008) have studied the diffusion wear of titanium machining under dry machining. Fig. 2 shows the width of flank wear of WC/Co carbide tools in dry turning of Ti6Al4V at different cutting speeds. It can be seen that the flank wear of the carbide tools is greatly increased when using high cutting speeds. The rake face of the WC/Co carbide tool was worn more at the high cutting speed of 120 m/min as can be seen in Fig. 3.

Fig. 2. Flank wear of the WC/Co carbide tools in dry machining of Ti6Al4V alloy at different cutting speeds (Jianxin, Yousheng et al. 2008)

83-5

Fig. 3. SEM micrographs of the worn rake face of the WC/Co carbide tool after cutting (a) V=200m/min (a) 3min, (b) 6min, (c) 9min, and (d) 12 min (test conditions: V = 80m/min, f = 0.1mm/r, ap = 0.5 mm) (Jianxin, Yousheng et al. 2008) G. A. Ibrahim et al (G. A. Ibrahim 2009) have studied the wear mechanisms of CVD coated carbide tools in turning of Ti-6Al-4V with multilayer layer (TiN-Al2O3-TiCN-TiN) coated tool. Figure 3 shows the flank wear, crater wear and chip welded on the cutting edge near the nose radius, when machining titanium alloy at cutting speed of 55 m/min, feed rate of 0.25 mm/rev and depth of cut of 0.10 mm. The flank wear that occurred near to the nose radius was due to low depth of cut. The low depth of cut caused a small contact area between cutting tool and workpiece material. Turning at low cutting speed, will generate low temperature.

Fig. 4. SEM micrograph shows the flank wear, crater wear and chip stick on cutting edge of CVD coated insert at v = 55 m/min, f = 0.25 mm/rev., d = 0.10 mm and Cutting Time = 24,53 min (G. A. Ibrahim 2009). Ginting and Nouari (Nouari and Ginting 2006) also confirmed that the dominant tool wear of uncoated and coated tool was localized flank wear. Abrasive and adhesive wear which occurred at the cutting speed of 95 m/min, feed rate of 0.35 mm/rev and depth of cut of 0. 20 mm are shown in Figure 4. The wear increases with cutting speed owing to increase in the slip distance and cutting temperature during machining. Adhesion or welding of titanium alloy onto the flank and rake faces were also observed. Adhesion of titanium alloy can be seen clearly, which clearly proves a strong bond (no evidence of any gaps) at the workpiece-tool interface. Figure 5 shows the different types of wear that has occurred while machining Titanium with a CVD coated tool. According to Konig (1979), the adhesion wear took place after the coating had worn out or coating delamination had occurred. The adhered or welded titanium will be swept away by the chip, and

83-6

deposited on to the workpiece continuously, thus leading to the initiation of chipping, flaking and finally the breakage of the carbide tool at the cutting edge.

Fig. 5. SEM Micrograph showing different types of wear while machining Titanium on cutting edge of a CVD Coated Insert at V= 95 m/min, f = 0.35 mm/rev., d = 0.20 mm and Cutting time = 20, 15 min (G. A. Ibrahim 2009). Flank wear, notching and built-up edge are the common types of tool wear when machining titanium. Edge notching appears as a localized abrasive wear on both the flank and rake face, along the line corresponding with the depth-of-cut parameter. This wear is caused partially by the presence of a hardened layer that typically is formed by previous casting, forging, heat treating, or prior machining operations. Chemical reaction between the cutting tool material and the workpiece also could lead to a notching-wear mechanism. This occurs when machining temperatures exceed 800C, and induce diffusion between the tool and the workpiece.

5. Summary
Titanium-based alloys are classified as difficult to machine materials as they cause high tool wear. The low thermal conductivity of titanium-based alloys causes higher rate of heat transfer to the tool that leads to rapid tool deterioration. The current research into the chemistry of the materials transferred to the tool flank land indicates that the presence of compounds with lower thermal conductivity at the tool surface leads to lower tool wear. From the literature it is suggested that coating materials with thermal conductivity lower than that of the workpiece material can be used to improve tool life for machining titanium-based alloys. Based on the previous research results there is a need to develop suitable tool coatings to improve the machinability of titanium based alloys. In addition to thermal conductivity, the solubility and reactivity of coating materials with titanium is also an important factor in tool wear/life. Further it is suggested that multi-layer thermal barrier coatings with improved surface properties, in terms of reactivity with titanium may be newly developed that could be used effectively in the industry for machining titanium-based alloys. Also it is desired that while designing such multilayered coatings, the thermal expansion coefficients of the constituents should be appropriately taken into consideration so as to minimize thermal shock and therby leading to better adhesion of the coating materials onto the substrate. Detailed review of tool wear while machining of titanium and its alloys has been presented in this paper. Different types of tool wear have also been explained which was encountered by many researchers while machining of titanium alloys. The coatings on inserts have been deposited by two main process routes - chemical vapour deposition (CVD) and physical vapour deposition (PVD), each of which has its own advantages,limitations, and have received wide application for the deposition of wear resistance coatings on the cutting inserts. From the literature it has been reported that machining of titanium alloy with PVD coatings during continuous and intermittent cutting resulted in high efficiency of these coatings both for the turning and also for milling operations. The type of wear observed in such machining

83-7

processes is crater wear, delamination of coating and diffusion on rake face and abrasion wear on the flank face in both CVD and PVD coatings. Hence a proposal of a new coated tool with multi layer coating for machining of titanium and its alloys to resist high wear resistance and at the same time having a low thermal conductivity has been made in the current paper. In short, this paper gives a complete idea about the different wear mechanisms which exist for major tool wear like crater and flank wear. At the end a possible way to control such wear in cutting tools has been suggested.

References A.R. Machado, J. W. (1990). "Machining of titanium and its alloysa review." Proc. Inst. Mech. Eng 204: 5360. Abele, E. a. F., B (2008). "High Speed Milling of Titanium alloys." Advances in Production Engineering & Management 3(3): 131-140. Bermingham, M. J., J. Kirsch, et al. (2011). "New observations on tool life, cutting forces and chip morphology in cryogenic machining Ti-6Al-4V." International Journal of Machine Tools and Manufacture 51(6): 500-511. Blinkov, I., V. Anikin, et al. (2011). "Acquisition and properties of wear-resistant PVD/CVD coatings on a hard-alloy tool." Russian Journal of Non-Ferrous Metals 52(1): 109-114. Bouzakis, K. D., S. Hadjiyiannis, et al. (2003). "The influence of the coating thickness on its strength properties and on the milling performance of PVD coated inserts." Surface and Coatings Technology 174-175(0): 393-401. Bouzakis, K. D., S. Hadjiyiannis, et al. (2004). "The effect of coating thickness, mechanical strength and hardness properties on the milling performance of PVD coated cemented carbides inserts." Surface and Coatings Technology 177-178(0): 657-664. Bouzakis, K. D., N. Michailidis, et al. (2001). "Failure mechanisms of physically vapour deposited coated hardmetal cutting inserts in turning." Wear 248(1-2): 29-37. Bouzakis, K. D., N. Michailidis, et al. (2000). "Interpretation of PVD Coated Inserts Wear Phenomena in Turning." CIRP Annals - Manufacturing Technology 49(1): 65-68. Che-Haron, C. H., A. Ginting, et al. (2007). "Performance of alloyed uncoated and CVD-coated carbide tools in dry milling of titanium alloy Ti-6242S." Journal of Materials Processing Technology 185(13): 77-82. Dearnly PA, G. A. (1986). "Evaluation of principal wear mechanisms of cemented carbides and ceramics used for machining titanium alloy IMI 318." Mater Sci Technol 2: 47-58. Ezugwu, E. O., J. Bonney, et al. (2003). "An overview of the machinability of aeroengine alloys." Journal of Materials Processing Technology 134(2): 233-253. Ezugwu, E. O. and Z. M. Wang (1997). "Titanium alloys and their machinabilitya review." Journal of Materials Processing Technology 68(3): 262-274. G. A. Ibrahim, C. H. C. H. a. J. A. G. (2009). "Progression and wear mechanism of CVD Carbide tools in turning Ti-6Al-4V ELI." International Journal of Mechanical and Materials Engineering 4(1): 35-41. Guimu, Z., Y. Chao, et al. (2003). "Experimental study on the milling of thin parts of titanium alloy (TC4)." Journal of Materials Processing Technology 138(13): 489-493. Hartung, P. D., B. M. Kramer, et al. (1982). "Tool Wear in Titanium Machining." CIRP Annals Manufacturing Technology 31(1): 75-80. Hong, H., A. T. Riga, et al. (1993). "Machinability of steels and titanium alloys under lubrication." Wear 162164, Part A(0): 34-39.

83-8

Hong, S. Y. and Y. Ding (2001). "Cooling approaches and cutting temperatures in cryogenic machining of Ti-6Al-4V." International Journal of Machine Tools and Manufacture 41(10): 1417-1437. Hong, S. Y., Y. Ding, et al. (2001). "Friction and cutting forces in cryogenic machining of Ti 6Al4V." International Journal of Machine Tools and Manufacture 41(15): 2271-2285. Hong, S. Y., I. Markus, et al. (2001). "New cooling approach and tool life improvement in cryogenic machining of titanium alloy Ti-6Al-4V." International Journal of Machine Tools and Manufacture 41(15): 2245-2260. Jaffery, S. and P. Mativenga (2011). "Wear mechanisms analysis for turning Ti-6Al-4V towards the development of suitable tool coatings." The International Journal of Advanced Manufacturing Technology: 1-15. Jawaid, A., C. H. Che-Haron, et al. (1999). "Tool wear characteristics in turning of titanium alloy Ti-6246." Journal of Materials Processing Technology 9293(0): 329-334. Jawaid, A., S. Sharif, et al. (2000). "Evaluation of wear mechanisms of coated carbide tools when face milling titanium alloy." Journal of Materials Processing Technology 99(1-3): 266-274. Jianxin, D., L. Yousheng, et al. (2008). "Diffusion wear in dry cutting of Ti6Al4V with WC/Co carbide tools." Wear 265(11-12): 1776-1783. Kitagawa, T., A. Kubo, et al. (1997). "Temperature and wear of cutting tools in high-speed machining of Inconel 718 and Ti6Al6V2Sn." Wear 202(2): 142-148. Klocke, F., N. Michailidis, et al. (2010). "Investigation of coated tools cutting performance in milling Ti6Al4V and its correlation to the temperature dependent impact resistance of the film." Production Engineering 4(5): 509-514. Komanduri, R. and W. R. Reed Jr (1983). "Evaluation of carbide grades and a new cutting geometry for machining titanium alloys." Wear 92(1): 113-123. Komanduri, R. and B. F. Von Turkovich (1981). "New observations on the mechanism of chip formation when machining titanium alloys." Wear 69(2): 179-188. Li, A., J. Zhao, et al. (2011). "Progressive tool failure in high-speed dry milling of Ti-6Al-4V alloy with coated carbide tools." The International Journal of Advanced Manufacturing Technology: 1-14. Li, S. J., R. Yang, et al. (2004). "Wear characteristics of TiNbTaZr and Ti6Al4V alloys for biomedical applications." Wear 257(9-10): 869-876. Lpez de lacalle, L. N., J. Prez, et al. (2000). "Advanced cutting conditions for the milling of aeronautical alloys." Journal of Materials Processing Technology 100(13): 1-11. Narutaki, N., A. Murakoshi, et al. (1983). "Study on Machining of Titanium Alloys." CIRP Annals - Manufacturing Technology 32(1): 65-69. Nouari, M. and A. Ginting (2006). "Wear characteristics and performance of multi-layer CVDcoated alloyed carbide tool in dry end milling of titanium alloy." Surface and Coatings Technology 200(18-19): 5663-5676. Ribeiro, M. V., M. R. V. Moreira, et al. (2003). "Optimization of titanium alloy (6Al-4V) machining." Journal of Materials Processing Technology 143-144: 458-463. Su, Y., N. He, et al. (2006). "An experimental investigation of effects of cooling/lubrication conditions on tool wear in high-speed end milling of Ti-6Al-4V." Wear 261(7-8): 760766. Trent, E. M. (2000). "Metal Cutting." 4th Edition, Butterworth.

83-9

Venugopal, K. A., S. Paul, et al. (2007)1. "Growth of tool wear in turning of Ti-6Al-4V alloy under cryogenic cooling." Wear 262(9-10): 1071-1078. Venugopal, K. A., S. Paul, et al. (2007)2. "Tool wear in cryogenic turning of Ti-6Al-4V alloy." Cryogenics 47(1): 12-18. Yang, X. and C. Richard Liu (1999). "MACHINING TITANIUM AND ITS ALLOYS." Machining Science and Technology 3(1): 107-139. Zhang, S., J. Li, et al. (2010). "Tool wear and cutting forces variation in high-speed end-milling Ti-6Al-4V alloy." The International Journal of Advanced Manufacturing Technology 46(1): 69-78.

83-10

You might also like