iThenticate

Document Viewer
Similarity Index
25%

The effect of tool edge geometry on tool perfor...

By: Yanuar Burhanuddin

As of: Sep 21, 2020 7:12:22 AM
4,328 words - 50 matches - 38 sources

sources:

paper text:

Advanced Materials Research Vols. 264-265 (2011) pp 1211-1221 Online available since 2011/Jun/30 at www.scientific.net © (2011) Trans Tech Publications, Switzerland doi:10.4028/www.scientific.net/AMR.264-265.1211
THE EFFECT OF TOOL EDGE GEOMETRY ON TOOL PERFORMANCE AND SURFACE INTEGRITY
IN TURNING TI-6AL-4V ALLOYS Yanuar Burhanuddin, Che Hassan Che Haron, Jaharah A. Ghani Department of Mechanical and Materials Engineering Faculty of Engineering and Built Environment Universiti Kebangsaan Malaysia Email : yanuarb@eng.ukm.my, chase@eng.ukm.my, jaharah@eng.ukm.my KEYWORDS: surface integrity, Ti–6Al–4V, titanium alloys, turning ABSTRACT This paper focuses on the influence of cutting tool edge preparation,
cutting speed and feed rate on the tool performance and
workpiece’s surface integrity in dry turning of Ti-6Al-4V alloy using PCBN inserts. The parameters evaluated are tool life, wear rate, wear mechanisms, surface roughness and subsurface microstructure alterations.
The rate of wear growth of the insert was assessed by progressive flank wear using optical microscope by taking photographs after certain length of cut. The wear mechanism at the end of tool life was investigated in detail using scanning electron microscope (SEM) and EDAX analysis. The
results show, by increasing the cutting speed and feed rate resulted in tool life reduction. Cutting with honed edge insert at cutting speed of 180 m/min has shown very little wear, even after 20 min of cutting. The honed insert proved less sensitive to increases in feed rate than the chamfered insert. In general the honed insert showed a significant improvement in tool life. All inserts failed due to attrition wear and adhesion. No flank notch wear was observed, but some crater wear was occurred behind the cutting edge. Microstructure alteration was not encountered when machining using the different edge geometry. In these trials the subsurface micro structural
deformation caused by machining consists of deformed grain boundaries in the direction of cutting and elongation of grains. Chip smearing and debris on the surface
was also found. The hardness in the vicinity of machined surface (at distance <0.15 mm) was lower than the hardness in the deeper machined surface (at distance > 0.15 mm). The layer softening was occurred rather than the hard white layer. The machined surface hardness was not affecting by different type of edge geometry.
1. INTRODUCTION Titanium alloys are used extensively in the aerospace industry
for structural components and as compressor blades, discs, casings, etc., in the cooler parts of gas turbine engines. They have also found use in such diverse areas as the energy and chemical processing industries, offshore and marine applications, the automotive industry, medical implants and sporting equipment. Titanium alloys have excellent strength-to-weight ratios and good elevated-temperature properties. Consequently when operating temperatures exceed 1308 oC, titanium can be used as an alternative to aluminium or, at higher temperatures still, titanium can be used as a lightweight alternative to nickel-based alloys or steel (Ezugwu et al. 2003, Boyer 1996). Titanium also exhibits a higher
resistance to corrosion than either aluminium alloys or low-alloy steels; thus it is usually preferred to these materials in corrosive aerospace applications
(Ezugwu 2003). Ti–6Al–4V is a general- purpose grade
alloy. It is by far the most important and widely used titanium alloy, accounting for about 60 per cent of total titanium
production (Boyer 1996). It is unfortunate that the inherent properties that make titanium such an attractive engineering material also ensure that it is classed as ‘difficult to machine’, regardless of the parameters being used to measure machinability. The main tool failure criteria reported are rake and flank face wear
All rights reserved. No part of contents of this paper may be reproduced or transmitted in any form or by any means without the written permission of TTP, www.ttp.net. (ID: 114.79.
0.165-21/11/12,03:23:27) resulting
from two wear mechanisms: dissolution–diffusion and attrition, the former producing smoothly worn surfaces and the latter irregularly worn surfaces
(Ezugwu & Wang 1997). One of the main properties of titanium alloys that lead to problems during machining is
low thermal conductivity (about 7 W/m K). The high
temperatures generated during cutting cannot be dissipated into the bulk of the workpiece and as a consequence they are localized at the tool tip. Indeed, Narutaki and Murakoshi (1983) have shown that the average tool interface temperature reaches about 1000 oC at a surface cutting speed of just 60 m/min, with approximately 1400 oC being recorded at a speed of 210 m/min. Hartung (1982) found that
most potential tool materials either rapidly dissolve in or chemically react with titanium
under these conditions, thus leading to the presence of a built-up layer and premature failure due to diffusion and/or attrition. This problem is further aggravated as a result of the thinner chips obtained during titanium machining, which produces a chip–tool contact area around one-half that of steel (Siekmann 1955). As a result, cutting forces are concentrated
at the very edge of the tool where the maximum cutting temperature is
generated, leading to rapid wear and plastic deformation of tools at high cutting speeds (von Turkovich & Durham 1982).
To attain higher cutting speeds, it is necessary to suppress the cutting temperature as much as possible while dissipating heat quickly
(Narutaki & Murakoshi 1983). As a solution to this problem, Hong & Zhao (1999) found that tool life could be improved by up to three times when liquid nitrogen was applied as a coolant instead of conventional cutting fluid when turning Ti–6Al–4V. The improvement was attributed to a reduction in tool–workpiece interface temperature and the subsequent maintenance of tool hardness and reduction in chemical reactivity of the titanium alloy. However, expensive modifications to the machine tool are required to utilize liquid nitrogen. Similar improvements in tool life were obtained by Ezugwu et al. (2005) when machining Ti–6Al–4V
using high-pressure coolant, which improved coolant penetration to the cutting interface and reduced cutting temperature
and diffusion wear rate. However, the coolant use will harm the operator and imposed on the high coolant treatment cost. Therefore, applying high temperature resistant cutting tools (e.g. CBN and PCD insert) when machining in dry condition is one of the alternatives to solve the problem. 2. EXPERIMENTAL SETUP 2.1 Tool The experimental work detailed in this paper was undertaken
to evaluate the performance of two different inserts when turning Ti–6Al–4V.
The first insert was a chamfered insert (ISO classification CNGA 120408S1020). The second was an insert with an edge honed (ISO classification CNGA 120408E). TABLE 1 Chemical compositions
of Ti-6Al-4V C Fe N O H Al V
Sn - 0.216 0.0024 0.160 0.0033 6.28 4.27 -
TABLE 2 Physical properties of Ti-6Al-4V Characteristics Ti-6Al-4V
Melting point (K) Density (x103 kg/m3) Young’s modulus (MPa) Modulus of rigidity (MPa)
Poisson’s ratio Specific heat (KJ /(kg.K) Thermal conductivity (W/(m.K)
1813 – 1923 4.42 113,190 44,100 0.3 – 0.33 0.56 7.54 2.2 Workpiece materials and equipment The
workpiece material used was a Ti–6Al–4V bar with 150 mm diameter and 300 mm long. The chemical composition
of Ti–6Al–4V was shown in Table 1. All machining
trials were conducted using a Colchester T2 CNC Lathe. Samples of the machined workpiece were cut from the bar using a hand saw.
These samples were used for microstructural and microhardness evaluation. Samples were hot mounted in bakelite, ground using SiC paper and then polished using diamond grit. After polishing they were etched by immersing in Kroll’s reagent for about 10 s. Microhardness measurements were carried out using a Vickers indenter at a load of 10 g for 15 s. One of the problems associated with microhardness measurements concerns, was its sensitivity to hard particles just below the workpiece surface. To overcome this, series of measurements were taken to obtain an average value. Subsurface microstructural analysis was conducted using a Leica optical microscope up to a maximum of 50 times magnification. Surface roughness measurements
were carried out using Perthometer Mahr. Measurements were conducted number of times at various positions on the bar. Tool wear was measured with a toolmaker’s microscope. 2.3 Experimental procedure The machining trials were carried out according to
ISO 3685 for tool-life testing with single point cutting tools.
Testing was carried out on a previously turned workpiece material, cuts for between 30 s and 1 minute were taken depending upon the level of wear experienced. Tests were stopped when either the average flank wear or maximum notch wear reached 0. 30 mm or 0.5 mm respectively.
All inserts were held in a
tool holder corresponding to the PCLNR configuration. Tests were carried out to investigate the effect of insert preparation, cutting speed, feed rate and tool edge condition on tool life/wear and workpiece surface integrity. Samples from the workpiece were taken
from machined surface when cutting using a worn tool to about 0.30
mm average flank wear; by doing this, the effects of both cutting parameters and the level of tool wear could be examined. The detailed cutting parameters and
test matrix are given in Table 3. Both inserts tested were high content PCBN (grade KD081) with a -5o rake angle, 5o clearance angle, 85o side cutting edge angle and a 0.8 mm nose radius. TABLE 3 Level designation of different cutting and tool parameters Level V F D (mm) Edge Geo. (m/min) (mm/rev) -1 (Low) 180 0.05 0.1 Chamfered 1 (High) 280 0.25 0.5 Honed
3. RESULTS AND DISCUSSION 3.1 Tool life and wear
rate The flank wear progression curves for two different edge geometries of PCBN cutting tools when turning Ti-6Al-4V alloy is shown in Figure 1. For the test
at cutting speed of 180 m/min, feed rate of 0.05 mm/rev, depth cut of
0.1 mm and honed shape geometry the flank wear less than 0.2 mm even though the tool life has reached 20 minutes. The cutting test is stopped if one of the criteria has been fulfilled such as the cutting time elapsed more than 20 minutes, even though with the flank wear is less than 0.2 mm. The statistical ANOVA confirms that
cutting speed had the largest influence on tool life and followed by feed and depth of cut
as commonly reported by other researchers (Ezugwu 2003 et al., Che-Haron 2001, Nabhani 2001, Kahles et al. 1985). The curve seemed horizontal after the cutting elapsed time of 1.66 minutes
at cutting speed of 180 m/min, feed rate of 0.05 mm/rev, depth cut of
0.5 mm and honed edge. For both cutting speeds, an increase in feed rate resulted in reducing tool life for both inserts, therefore the chamfered insert proved far more sensitive to the parameters change. 3.2 Wear mechanism The wear mechanism for two different edge of PCBN cutting tool and two different CBN content are similar. Tools wear occurred mostly by adhesion and diffusion. The adherent layer was seen
welded to the tool surface on both the flank and the crater faces due to the high pressures and temperatures developed during cutting. This demonstrates the high chemical affinity of titanium which is similarly reported by other
researchers (Ezugwu 2003 et al., Ezugwu & Wang 1997, Yang & Liu 1999, Bhaumik et al 1995). However the attrition and adhesion occurred more often in both chamfered edge and honed edge (Figure 4). These mechanisms occurred especially at low feed rate and depth of cut. The dissolution-diffusion and catastrophic failure were occurred at high
feed rate and depth of cut (Figures 5, 6). At the higher cutting speed,
it may have been expected that the honed tool edge geometry would plastically deformed. Furthermore, the behaviors of chip formed at an accelerated rate were due to localized temperatures and stresses at the tool tip were not observed but surprisingly the honed inserts showed catastrophic fracture at
0.25 mm/rev feed rate and 0.5 mm depth of cut.
Figure 4 also show the Ti-6Al-4V smeared the tool rake face. The splashed titanium occurred when the burnt chips stuck both on the tool and workpiece during machining. The thin chip like a paper is easily ignited. The burnt chip during machining is unwanted because endangered the operator. The SEM micrograph and EDAX analysis (Figure 3) proved the splash occurrence with the existing of titanium and aluminum element. No evidence of notch wear
was observed on the flank face of the inserts as shown in
Figures 4, 5 and 6. 3.3 Surface Roughness The machined surface roughness values using two different edge PCBN tools are quite higher than using two different contents PCBN tools. At the
cutting speed of 280 m/min, feed of 0.25 mm/rev, depth of cut of 0. 5 mm and chamfered tool had a
surface roughness value of 3.5 µm. The maximum surface roughness value occurred at the end of machining run. The severe failure caused the irregular surface at the nose of tool. The irregular and rough tool nose surface resulted the rough surface and the abruptly raising the roughness value. The variation of machined surface roughness at 180 m/min cutting speed was almost similar to the machined surface roughness at 280 m/min cutting speed. The surface roughness was more influenced by feed rather than cutting speed. This is in accordance with the theory which stated that the surface roughness was a function of squared feed (Shaw 1986, Stepehenson & Agapiou 1996, Kalpakjian 2001). While the surface roughness from the different edge geometry view
can be seen in Figure 7. The machined surface roughness was influenced by the
feed and depth of cut rather than edge geometry. There were no considerable roughness differences at the different feed when using the honed edge. 3.4 Microstructure alterations and surface hardness Microstructure alteration was not encountered when machining using the different edge geometry. In these trials, however,
the subsurface microstructural deformation caused by machining consisted of deformed grain boundaries in the direction of cutting and elongation of grains (Fig 8). These types of defect are typically reported during the machining of titanium alloys
(Zlatin & Field 1973, Narutaki & Murakoshi 1983, Che-Haron 2001, Kahles et al 1985, Ezugwu & Wang 1997). The deformed grain boundary in the workpiece is very similar to other work using two different CBN content tools (Burhanuddin et al. 2008) as shown in Figure 8. Although plastic deformation is a phenomenon usually associated with abusive machining conditions, in these trials no discernible pattern could be found. The depths of deformation of about 0.15 mm were found (Fig. 8). It is suggested that a heavily deformed material layer could be formed at high temperatures generated during cutting without substantially altering the adjacent microstructure. However, the deformed grain boundary was caused by the excessive compression of cutting tool to the workpiece and exceeds the elastic limit. Figure 8 shows the deformed grain boundary. The deformed grain was not uniform beneath the machined surface. It shows the work piece surface was not fully circular so that the excessive compression or pressure occurred at certain areas. Chip smearing and debris on the surface were also found (Fig. 8). Chips were pulled back into the cutting zone and left rewelded to the surface as debris. It is important to note that no evidence of microcracks running in the bulk of the workpiece was observed in any of the operating parameters. Based on the Vickers hardness test to the machined surface samples there are some inferences should be considered. First, the hardness in the vicinity of machined surface ( <0.15 mm) less than the hardness in the deeper machined surface (> 0.15 mm). It showed that the softening layer was occurred rather than the hard white layer. A softer surface was obtained due to the ‘overageing’ of the titanium as a result of the very high cutting temperatures developed. Similar results have been reported for ground Ti–6Al–4V samples (Zlatin & Field 1973). Secondly, the effect of edge geometry difference was not affected the machined surface hardness as shown in Table 3. The hardness of 4B sample number (cutting condition: v = 180
m/min, f = 0. 05 mm/rev, d = 0. 1 mm and honed edge) is relatively equal to the
7B sample number (cutting condition:
v = 180 m/min, f = 0. 25 mm/rev, d = 0.5 mm
dan chamfered edge). Although a slight microhardness variation was found between the surfaces generated by the two tool types, the results only indicate a general trend and no conclusions can be deduced. The microhardness variation is not significant in any of the conditions examined. 4. CONCLUSIONS
An increase in cutting speed and feed rate resulted in tool life
reduction. Little wear was observed after 20 min of cutting using honed edge insert at cutting speed of 180 m/min. The honed insert proved less sensitive in increasing the feed rate than the chamfered insert. In general the honed insert showed a significant improvement in tool life compared to chamfered insert. All inserts failed due to attrition and adhesion wear; no flank notch wear was observed but some crater wear was observed behind the cutting edge. Microstructure alteration was not encountered beneath the machined surface when machining using the different edge geometry. However, in these trials, the
subsurface microstructural deformation caused by machining consists of deformed grain boundaries in the direction of cutting and elongation of grains. Chip smearing and debris on the surface
were also found. The hardness in the vicinity of machined surface ( <0.15 mm) less than the hardness in the deeper machined surface (> 0.15 mm). The layer softening was occurred rather than the hard white layer. The edge geometry difference was not very affecting to the increase of machined surface hardness ACKNOWLEDGEMENTS
The authors would like to thank the Malaysian Ministry of Science, Technology and Environment for sponsoring this work under project IRPA 03-02-02- 0062-EA122. REFERENCES
[ 1 ] Ezugwu, E. O., Bonney, J. and Yamane, Y. An overview of the machinability of aeroengine alloys. J. Mater. Processing Technol., 2003, 134, 233–253. [ 2 ] Boyer, R. R. An overview on the use of titanium in the aerospace industry. Mater. Sci. Engng, 1996, A213, 103–114. [ 3 ] Smith, W. F. Structure and Properties of Engineering Alloys, 2nd edition, 1993 (McGraw- Hill, New York). [ 4 ] Boyer, R., Welsch, G. and Collings, E. W. Materials Proper- ties Handbook: Titanium Alloys, 1994 (American Society for Metals, Materials Park, Ohio). [ 5 ] Donachie, M. J. and Donachie, S. J. Superalloys: A Technical Guide, 2002 (American Society for Metals, Materials Park, Ohio). [ 6 ] Zlatin, N. and Field, M. Procedures and precautions in machining titanium alloys. Titanium Sci. Technol., 1973, 1, 489–501. [ 7 ] Narutaki, N. and Murakoshi, A. Study on machining of titanium alloys. Ann. CIRP, 1983, 32(1), 65–69. [ 8 ] Che-Haron, C. H. 2001. Tool life and surface integrity in turning titanium alloy. J. Mater. Processing Technol. 118: 231–237. [ 9 ] Hartung, P. D. 1981. Tool wears in titanium machining. Master Thesis, Massachuset Int. of Technology. [ 10 ] Nabhani, F. Machining of aerospace titanium alloys. Robotics Computer Integrated Mfg, 2001, 17, 99–106. [ 11 ] Nabhani, F. 2001. Wear mechanisms of ultra-hard cutting tool materials. J. Mater. Processing Technol. 17: 99–106 [ 12 ] Min, W. and Youzhen, Z. 1988. Diffusion wears in milling titanium alloys. Mater. Sci. Technol. 4: 548–553. [ 13 ] Kahles, J. F., Field, M., Eylon, D. and Froes, F. H. 1985. Machining of titanium alloys. J. Metals 37: 27–35. [ 14 ] Bhaumik, S. K., Divakar, C. and Singh, A. K. 1995. Machining Ti–6Al–4V alloy with a wBN–cBN composite tool. Mater. Des. 16(4): 221–226. [ 15 ] Komanduri, R. and Reed Jr., W. R. Evaluation of carbide grades and a new cutting geometry for machining titanium alloys. Wear 92: 113–123. [ 16 ] Wang, Z. M. and Ezugwu, E. O. 1997. Performance of PVD- coated carbide tool when machining Ti–6Al–4V. Tribology Trans. 40: 81–86. [ 17 ] Burhanuddin, Y., Che Haron, C.H., Ghani, J.A., Ariffin, A. K., Ibrahim, G.A., Yasir, A. & El-Maghribi, N.H. (2008). The Effects of CBN Cutting Tool Grades on the Tool Life and Wear Mechanism When Dry Turning of Titanium Alloy. Asian Int’l Journal of Sci. and Tech. in Production and Manufacturing 1(2): 105-110. [ 18 ] Machado, A. R., Wallbank, J., Pashby, I. R. and Ezugwu, E. O. 1998. Tool performance and chip control when machining Ti6Al4V and Inconel 901 using high pressure coolant supply. Machining Sci. Technol. 2: 1–12. [ 19 ] Komanduri, R. Some clarifications on the mechanics of chip formation when machining titanium alloys. Wear, 1982, 76, 15–34. [ 20 ] Hong, S.Y. & Zhao, Z. 1999. Thermal aspects, material considerations and cooling strategies in cryogenic machining. Clean Products and Processes 1: 107–116 [ 21 ] Machado, A. R. and Wallbank, J. 1990. Machining of titanium and its alloys—a review. Proc. Instn Mech. Engrs, Part B: J. Engineering Manufacture. 204(B1): 53–60. [ 22 ]Donachie, M. J. Titanium, A Technical Guide, 1988 (American Society for Metals, Metals Park, Ohio). [ 23 ] Yang, X. and Liu, C. R. 1999. Machining titanium and its alloys. Machining Sci. Technol. 3(1): 107–139. [ 24 ] ISO 3685 Specification for Tool Life with Single Point Turn- ing Tools, 1993 (International Standardization Organization, Geneva, Switzerland). [ 25 ] Ezugwu, E. O. and Wang, Z. M. 1997. Titanium alloys and their machinability—a review. J. Mater. Processing Technol. 68: 262–274. [ 26 ]Yeon Kim, K., Young Lee, W. and Chol Sin, H. A finite-element analysis of machining with the tool edge considered. J. Mater. Processing Technol. 86: 45–55. [ 27 ] Kalpakjian, S. 1995. Manufacturing Engineering and Technology, 3rd edition. Addison- Wesley: Reading, Massachusetts. [ 28 ] Kishawy, H. A. and Elbestawi, M. A. 1999. Effects of process parameters on material side flow during hard turning. Int. J. Mach. Tools Mfg 39:1017–1030. [ 29 ] Thiele, J. D. and Melkote, S. N. 1999. Effect of edge geometry and workpiece hardness on surface generation in the finish hard turning of AISI 52100 steel. J. Mater. Processing Technol. 94:216–226. [ 30 ] Zhou, J.M., Walter, H., Anderson, M. & Stahl, J.E. 2003. Effect chamfer angle on wear of PCBN cutting tool. Int. J. Mach. & Manufacture 43: 301-305. [ 31 ]Klocke, F., Konig, W. & Gerschwiller, K. 1996. Advanced machining of titanium and nickel- based alloys. in E. Kuljanic (Ed.) Advanced Manufacturing Systems and Technology pp. 7-21 [ 32 ] Wang, Z.G., Rahman, M. & Wong, Y.S.. 2005. Tool wear characteristics of binderless CBN tools used in high-speed milling of titanium alloys Wear 258:752–758 0.6 180 m/min-0.05 mm/rev;0.1mm;Honed 280
m/min;0. 05 mm/rev; 0.1 mm; Chamfered 180 m/min;0. 25 mm/rev; 0.1 mm;
Chamfered 0.5 280 m/min;0.25 mm/rev;0.1 mm;Honed Flank Wear VB (mm) 180
m/min;0.05 mm/rev;0.5 mm;
Chamfered 0.4 280
m/min;0. 05 mm/rev; 0 .5 mm; Honed 180 m/min;0. 25 mm/rev; 0.5 mm; Honed 280 m/min;0. 25 mm/rev;0.5 mm;
Chamfered 0.3 0.2 0.1 0 0 200 400 600 800 1000 1200 Time (s) FIGURE 1 Tool wear progression curves 0.01400 0.01200 Wear Rate (mm/sec) 0.01000 0.00800 0.00600 0.00400 0.00200 0.00000 Honed 0.1 0.05 180.0 Honed 0.5 0.05 280.0 Honed 0.5 0.25 180.0 Honed 0.1 0.25 280.0 Chamfered 0.1 0.25 180.0 Chamfered 0.1 0.05 280.0 Chamfered 0.5 0.05 180.0 Chamfered 0.5 0.25 280.0 FIGURE 2 Tool wear rate FIGURE 3 SEM micrograph of honed-edge PCBN cutting tools after machining
at cutting speed of 280 m/min, feed of 0. 25 mm/rev and depth of cut of 0. 1 mm.
FIGURE 4 EDAX analysis of honed-edge PCBN cutting tools after machining
at cutting speed of 280 m/min, feed of 0. 25 mm/rev and depth of cut of 0. 1 mm.
FIGURE 5 SEM micrograph of honed-edge PCBN cutting tools after machining at
cutting speed of 280 m/min, feed of 0.05 mm/rev and depth of cut of 0..5 mm.
FIGURE 6 SEM micrograph of honed-edge PCBN cutting tools after machining at
cutting speed of 180 m/min, feed of 0.25 mm/rev and depth of cut of 0.5 mm.
2.00 1.80 Mean
Average Roughness ( µm) 1.60 1.40 1.20 1.00 0.80 0.60 0.40 0.20 0.00
1.84 1.65 1.62 0.99 0.56 0.30 c 0.47 0.29 Honed Honed Honed Honed Chamfered Chamfered Chamfered Chamfered 0.1 0.5 0.5 0.1 0.5 0.1 0.1 0.5 0.05 0.25 0.05 0.25 0.05 0.25 0.05 0.25 180.0 180.0 280.0 280.0 180.0 180.0 280.0 280.0 FIGURE 7 Machined surface roughness variation FIGURE 8 SEM micrograph of deformed grain boundary in the cutting direction when machining using honed edge PCBN tool.
Advances in Materials and Processing Technologies II 10.4028/www.scientific.net/AMR.264-265 The Effect of Tool Edge Geometry on
Tool Performance and Surface Integrity in Turning Ti-6Al-4V Alloys 10.4028/www.scientific.net/AMR.264-265.1211 1212 Advances in Materials and Processing Technologies II Advanced Materials Research Vols. 264-265 1213 1214 Advances in Materials and Processing Technologies II Advanced Materials Research Vols. 264-265 1215 1216 Advances in Materials and Processing Technologies II Advanced Materials Research Vols. 264-265 1217 1218 Advances in Materials and Processing Technologies II Advanced Materials Research Vols. 264-265 1219 1220 Advances in Materials and Processing Technologies II Advanced Materials Research Vols. 264-265 1221