You are on page 1of 14

Surface & Coatings Technology 200 (2006) 5663 5676 www.elsevier.

com/locate/surfcoat

Wear characteristics and performance of multi-layer CVD-coated alloyed carbide tool in dry end milling of titanium alloy
M. Nouari a,*, A. Ginting b
a b

Laboratoire Materiaux Endommagement Fiabilite and Ingenierie des Procedes Ecole Nationale Superieure dArts et Metiers, CER Bordeaux. Esplanade des Arts et Metiers, 33405 Talence Cedex, France Department of Mechanical Engineering, Faculty of Engineering, University of Sumatera Utara Jalan Almamater, 20155 Medan, North Sumatera, Indonesia Received 2 November 2004; accepted in revised form 14 July 2005 Available online 8 September 2005

Abstract This paper deals with a study on the performance of alloyed carbide tools during dry machining of the titanium alloy Ti-6242S, commonly used for aeroengines. Experimental tests were conducted using two kinds of alloyed carbide tool inserts: the uncoated carbide tool (tool A) and the multi-layer CVD-coated tool (tool B). Tool failure modes and wear mechanisms for both tools were examined at various cutting conditions. The localized flank wear (VB) was found to be the predominant tool wear for both tools A and B. At VB close to 0.3 mm, a brittle fracture (cracking, flaking and chipping) was noted, a plastic deformation was significantly observed and a phenomenon of coating delamination was noticed. Adhesion wear (attrition and galling) and diffusion wear have been the wear mechanisms of tools A and B. Coating delamination was the initial wear mode of tool B; it occurred after a few seconds of cutting time. The performance of the uncoated and the multi-layer CVD-coated alloyed carbide tools was analyzed in terms of tool life and surface finish. Thanks to the Finite Element Analysis (FEA) on the tribological parameters and on the chip segmentation, a new proposal on the physical mechanisms of coating delamination was carried out. D 2005 Elsevier B.V. All rights reserved.
Keywords: Wear mechanisms; Localized flank wear VB3; Finite Element Analysis (FEA); Coating delamination; Tool life; Surface finish

1. Introduction In some cases, machining without using cutting fluids (dry machining) can successfully be implemented in the industrial machining applications. Dry machining has two main impacts: an ecological impact and an economic one. Indeed, the absence of cutting fluids preserves the environment and reduces the production costs of 16% to 20% [1]. That is why several European countries incite the metal cutting industry to use dry machining. Cutting fluids are used in metal cutting process as a lubricant and a coolant. Their absence in machining means a high friction and a high cutting temperature at the tool
* Corresponding author. Tel.: +33 5 56 84 54 48; fax: +33 5 56 84 53 66. E-mail address: mohammed.nouari@lamef.bordeaux.ensam.fr (M. Nouari). 0257-8972/$ - see front matter D 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.surfcoat.2005.07.063

workpiece interface. This drastically affects the tool wear and the tool life. Therefore, dry machining represents a great challenge to manufacturing engineers because of the high temperatures generated especially when machining aerospace materials as titanium and nickel based superalloys. For dry machining applications, the cutting tool can be designed in three different ways by: using new tool materials, adapting new tool geometries, and applying different coating materials [1 3]. The last solution is the most frequently chosen. Coating on carbide tools was initially developed using the Chemical Vapour Deposition (CVD) technique in the early sixties, and its application has constantly increased [4 7]. Coating materials such as TiN, TiC and Al2O3 act as a protective layer against abrasion, corrosion and oxidation wear on the tool surface. Nowadays, most of the multi-layer coating materials contain a combination of TiN, TiC,

5664

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676 Table 2 Physical properties of titanium alloy Ti-6242S Tensile strength (MPa) Yield strength (MPa) Creep stress (MPa) Hardness (HRc) Density (kg/m3) Linear thermal expansion (10 6/-C) Thermal conductivity (W/mK) ! 895 830 240 36 4540 9.9 8 12

Ti(C,N) and Al2O3; to improve the tool life, they are deposited in different sequences. In machining of steels, it is proved that the CVD-coated carbide provides a better result comparing to the uncoated one [8]. In the case of aeronautic titanium alloys, Jawaid et al. [9] have reported that the CVD-coated carbide (W Ta/Nb C Co + TiCN/Al2O3) shows a good performance when face milling Ti 6Al 4V under wet cutting conditions. Tool lives of 7 and 30 min were recorded when cutting speeds of 55 and 100 m/min were used. The other considered cutting conditions are: a feed rate f z of 0.1 mm/tooth, an axial depth of cut a a of 2 mm and a radial depth of cut a r of 5.8 mm. Moreover, Fitzsimmons and Sarin [10] have shown that turning the aeronautic titanium alloy Ti6Al 4V with a CVD-coated carbide tool can successfully be done. Nevertheless, the recent review on the machining of titanium alloys conducted by Ezugwu et al. [11] indicates that there are not enough results on the performance of the multilayer CVD-coated carbide tools when machining this kind of aeronautic material under dry cutting conditions. The present paper deals with a study on the dry machining characteristics of the multilayer CVD-coated carbide in end milling of the aeroengine material alloy Ti6242S. The results that are reported in this paper follow the previous work of Ginting and Nouari [12]. This study is about the wear mechanisms of the uncoated alloyed carbide W Ti/Ta/Nb C Co (tool A) and the coated alloyed carbide (tool B) using the CVD multi-layer deposition technique. The coating consists of TiN, TiCN and TiC materials. Many researchers agree that the alloyed carbide tools are not suitable for machining titanium alloys [13 16]. However, a previous study [12] confirmed that the uncoated alloyed carbide tools (round insert type) in the off centre ball end mill configuration shows a good performance when dry end milling titanium alloy Ti-6242S. A tool life of 11.3 min and a surface finish R a of 0.61 Am were obtained. The corresponding cutting conditions are: cutting speed V c = 150 m/min, feed rate f z = 0.15 mm/tooth, axial depth of cut a a = 2 mm and radial depth of cut a r = 8.8 mm. Note that the round

insert has 6 indexes, and each insert has a total tool life of 67.8 min.

2. Materials and methods 2.1. Materials The a h titanium alloy Ti-6242S is selected as a workpiece material. The chemical composition and physical properties are given in Tables 1 and 2, respectively. The cutting experiments were conducted using 2 types of alloyed carbide tool inserts: coated and uncoated with an insert diameter D i of 12 mm (grade P of ISO codes), the chemical composition and physical characteristics of the substrate material is given in Table 3. The CVD multi-layer coating is described in Table 4. During end milling tests, each insert was rigidly mounted on a tool holder to provide the off centre ball end mill configuration with a nominal diameter D n of 16 mm (see Fig. 1). The tool geometry is defined by the following parameters: cutting rake angle go = 6-, axial rake angle gp = 6- and radial rake angle gf = 2-. The end milling process and the ball end mill tool configuration are selected here since they are widely used in machining complex profiles (contouring) of aerospace components as turbine blades [17]. 2.2. Methods Machining tests were carried out on a CNC vertical milling machine with a power of 9 kW. This machine has a variable spindle speed in the range of 60 to 10,000 rpm.
Table 3 Nominal composition, mechanical and physical properties of the uncoated carbide tool (tool A) and the substrate of CVD-coated tool (tool B) Tool substrate characteristic Tool(s) A and B Composition 69.8%WC 9.50%Co 20.7% (Ti/Ta/Nb) 12 1485 600 11.4 45 6.1 510 2.2

Table 1 Chemical composition of the aerospace titanium alloy Ti-6242S (wt.%) Elements Al Zr Mo Sn Fe O2 Si C N2 H2 Y Others Ti Minimum (%) 5.50 3.60 1.80 1.80 0.100 to 100% Maximum (%) 6.50 4.40 2.20 2.20 0.25 0.15 0.10 0.05 0.05 0.015 0.005 0.300

Grain size (Am) Hardness 25 -C (HV10) Hot hardness 800 -C (kg/mm2) Density (g/cm3) Thermal conductivity (W/mK) Thermal expansion (10 6/K) Modulus of elasticity (GPa) Traverse rupture (GPa)

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676 Table 4 Properties of CVD coatings Coating material characteristics Method of deposition Coating design (layer from inner to outer) Total thickness (Am) Coating material(s) Hardness 25 -C (HV10) Thermal conductivity 727 -C (W/mK) Thermal expansion (10 6/K) Melting point (-C) Density (g/cm3) Tool B Chemical Vapour Deposition (CVD) 9 layers TiN/TiC/TiN/TiC/TiN/TiC/TiN/TiCN/TiN ;10 TiN, TiC, TiCN 2200, 2110, 2300 25, 28, 31 9.35, 8.85, 8.65 2950, 3000, 3070 3.44, 3.65, 4.18

5665

analyzed for different tests using the multiple regression method. As shown in Fig. 1, to have a constant cutting speed at the leading edge (contact point A), the spindle speed N is calculated taking into account the tool effective diameter (D e): 2s 3 2  2  Di Di D e 24 aa d 5 2 2 Vc pDe Vc ! q D 2 2 Di i 2p 2 aa d 2

The tool rejection criteria for the cutting experiments were the localized flank wear VB3 ! 0.3 mm and the excessive chipping, flaking and/or fracturing of the cutting edge [18]. In order to understand the tool wear behaviour under dry cutting configuration, the factorial design 23 with 4 centre points and 3 replications (36 data for each tool) was used. The different cutting conditions are: cutting speeds V c in the range of 100 to 125 m/min, feed rates f z of 0.15 to 0.20 mm/tooth, axial depths of cut a a of 2 to 2.5 mm. For all experiments, the radial depth of cut a r was kept constant to 8.8 mm due to the common condition of contour machining [19]. To obtain the extended Taylor tool life model (wear law), the tool life data were recorded and

a a and D i are the axial depth of cut and the tool insert diameter, respectively (see Fig. 1). V c is the cutting speed (in m/min) and N the spindle speed (in rpm).

3. Wear characteristics 3.1. Flank wear progression and wear rate In accordance with the results previously presented in [12], the current investigation also confirms that the dominant tool wear of the uncoated tool A and the coated tool B is the localized flank wear VB3. As shown in Fig.

Fig. 1. The off centre ball end mill. a a is the axial depth-of-cut; D e the effective diameter; D n the nominal diameter; D i the insert diameter and d the off centre radius.

5666

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

2a and b, the localization of the flank wear takes place at the leading cutting edge or at the contact point A of Fig. 1. This is due to the fact that for the off centre ball endmilling configuration, the leading cutting edge and especially the contact point A is subject to extreme thermal and mechanical loads during the chip formation. In this work, end-milling tests were conducted using a down cutting configuration. It means that the greatest chip thickness equal to the feed rate value f z is produced at the entry position when the tool comes into contact with the machined part. Moreover, according to Eq. (2), the maximum rotation during the interrupted cutting process is located at point A due to the value of the effective diameter D e. Figs. 3 and 4 show the evolution of the flank wear VB vs. the cutting time for tools A and B, respectively. An increase in the flank wear level can easily be observed in terms of the cutting conditions V c, f z and a a. It can also be noticed that the flank wear progression of the coated tool B tends to be linear just after the initial wear (see the first point of each plot in Figs. 3 and 4). This behaviour can clearly be observed on the different curves of Fig. 4b. Moreover, Figs. 3 and 4 show that flank wear evolutions are similar for both tools. This fact is confirmed by Fig. 5 in which the evolution of the average flank wear rate vs. the cutting time is illustrated. The average values of the flank wear rate were calculated using the cutting time that is necessary to obtain VB3 near to 0.3 mm for each cutting condition. 3.2. Mechanical wear: flank wear and adhesion wear 3.2.1. Uncoated alloyed carbide (tool A) A microanalysis using Scanning Electron Microscopy (SEM) was performed on the uncoated tool A. Fig. 6 shows some micrographs of the leading cutting edge where the flank wear VB3 of 0.3 mm is located. A brittle fracture as cracking, flaking and chipping can be seen. In Fig. 6a, the worn tool profile when VB3 = 0.3 mm can easily be compared with the original one represented by the white dashed line. Because of flaking and chipping, some parts of the rake face and the flank face have been removed during the chip flow. Unfortunately, comparing with the results found in [12], the cracking mechanism is not clearly observed in the present work. It is probably due to the different tool rejection criteria taken into account in the two studies (in [12], VB3 = 0.9 mm and here VB3 = 0.3 mm). However, the tool life criteria VB3 of 0.3 mm provides new information on the tool wear behaviour. Fig. 6b illustrates the tool cutting edge that was plastically deformed. In fact, for brittle material such as tungsten carbide it is not easy to observe the plastic deformation phenomenon since the yield transition point is not clear between the elastic and plastic states of the brittle material. The SEM examination of the leading cutting edge (Fig. 7) proves that adhesion wear (attrition and galling)

is the wear mechanism of flaking and chipping for tool A. Adhesion that occurs during the machining of titanium alloys results from its high chemical reactivity to many tool materials [20]. Fig. 7a and b demonstrate that the Ti-6242S chip is bonded without any gap with the rake and the flank faces of tool A. The arrows in Fig. 7a and b show some grains in the adhesive layer; grains move away from their original positions in the tool material and diffuses into the chip material through the tool chip interface. The results of EDAX HPD and ViP analyses prove that these grains contain some constituents of substrate material such as W, C and Co (see Table 5). 3.2.2. CVD-coated alloyed carbide (tool B) Concerning the coated tool B, flaking, cracking, chipping, and plastic deformation can be observed at the leading cutting edge (see Figs. 8 11). In addition, a phenomenon of coating delamination is clearly distinguished. As for tool A, the results of the chemical analyses (EDAX HPD and ViP) confirm that adhesion is the wear mechanism of tool B, see Table 6. Adhesion wear is caused by the mechanical removal of the tool material when the adhesive junctions are broken. When VB3 attains 0.3 mm, the plastic deformation of the leading cutting edge of tool B occurs, Fig. 8a. This deformation significantly changes the tool shape; the black dashed line in Fig. 8a represents the original tool profile. SEM measurements gave 180 Am as the value of the tool profile deviation (measured perpendicularly from the dashed line to the top of the deformed area). Moreover, the high magnification of the affected zone (see the area delimited by the white circle in Fig. 8a) highlights cracking at the rake face of tool B. Other cracks observed on this tool are presented in Fig. 10a; this one shows the cracking phenomenon of the coating layer at the leading cutting edge when VB3 = 0.3 mm. In the case of tool B, micro-cracks take place in the coating layer and they are immediately followed by the removal of the coating material; this phenomenon is known as the coating delamination phenomenon. The evidence of the coating delamination is shown in Fig. 10a and b. Furthermore, Fig. 11 reveals that micro-cracks in the coating layer propagate not only in the horizontal direction but also in the vertical one through the interface between the tool substrate and the coating materials. From a physical point of view, the question then is: when and how is the coating delamination phenomenon made? To answer this question, supplementary dry machining tests were performed. In order to understand the initial tool wear, for each cutting test only one passage of the tool on the workpiece surface was authorized (only one machining pass); the total cutting length considered in the experiments was about 250 mm. The worn tool was sectioned and examined at the area where the initial flank wear was located. The results of these investigations

(a)
chip flaking
M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

VB1 aa< 2 mm VB3

(b)
chip flaking

VB1 aa< 2 mm VB3

Fig. 2. Flank wear obtained under a cutting speed V c of 100 m/min, a feed rate f z of 0.15 mm/tooth, an axial depth of cut a a of 2 mm, and a radial depth of cut a r of 8.8 mm. (a) Tool A, (b) tool B.

5667

5668

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

(a)

0.4
Wear progression of Tool A at cutting speed of 100 m/min

thermal and mechanical loads and cannot resist to the wear during the interrupted cutting of the milling process. Nevertheless, the cutting forces obtained when machining

0.3 Flank wear (mm)

(a) 0.4
0.2 0.3 0.1
0.15 mm - 2.0 mm 0.20 mm - 2.0 mm 0.15 mm - 2.5 mm 0.20 mm - 2.5 mm
Wear progression of Tool B at cutting speed of 100 m/min

Flank wear (mm) 20

0.2

0 0 5

(b)

10 15 Cutting time (min)

0.1

0.4
Wear progression of Tool A at cutting speed of 110 and 115 m/min

0.15 mm - 2.0 mm 0.20 mm - 2.0 mm 0.15 mm - 2.5 mm 0.20 mm - 2.5 mm

0 0 5 10 Cutting time (min) 15

0.3 Flank wear (mm)

(b)
0.2

0.4
Wear progression of Tool B at cutting speed of 110 and 115 m/min

0.3
110 m/min - 0.165 mm - 2.25 mm 110 m/min - 0.185 mm - 2.25 mm 115 m/min - 0.165 mm - 2.25 mm 115 m/min - 0.185 mm - 2.25 mm

Flank wear (mm)

0.1

0.2

0 0 2.5 5 Cutting time (min) 7.5

(c)

0.1

0.4
Wear progression of Tool A at cutting speed of 125 m/min

110 m/min - 0.165 mm - 2.25 mm 110 m/min - 0.185 mm - 2.25 mm 115 m/min - 0.165 mm - 2.25 mm 115 m/min - 0.185 mm - 2.25 mm

0 0.3 Flank wear (mm) 0 2.5 5 Cutting time (min) 7.5

(c)
0.2

0.4
Wear progression of Tool B at cutting speed of 125 m/min

0.1

Flank wear (mm)

0.15 mm - 2.0 mm 0.20 mm - 2.0 mm 0.15 mm - 2.5 mm 0.20 mm - 2.5 mm

0.3

0 0 2.5 5 7.5 Cutting time (min)


Fig. 3. Flank wear progression of tool A vs. cutting time. (a, c) Curves for different feed rates f z and different axial depths of cut a a. (b) Curves for different cutting speeds V c, different feed rates and different axial depths of cut.

0.2

0.1

0.15 mm - 2.0 mm 0.20 mm - 2.0 mm 0.15 mm - 2.5 mm 0.20 mm - 2.5 mm

0 0 2.5 5 7.5 Cutting time (min)


Fig. 4. Flank wear progression of Tool B vs. cutting time. (a, c) Curves for different feed rates f z and different axial depths of cut a a. (b) Curves for different cutting speeds V c, different feed rates and different axial depths of cut.

presented in Fig. 12 show that the coating delamination phenomenon is the initial wear mode of tool B. Initially, it may be supposed that the coating delamination is due to mechanical wear. The coating material is subject to great

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

5669

1000
Average flank wear rate of Tool A Average flank wear rate of Tool B Trend line of wear rate for Tool A

(a)
bakelite

Average flank wear rate (m/min)

Trend line of wear rate for Tool B


chip on rake face

100

(b)
10 1 10 Cutting time (min)
Fig. 5. Evolution of the average flank wear rate for both tools A and B. The cutting conditions are similar to those of Figs. 3 and 4.
1

bakelite

100
chip on rake face

flaking due to attrition and or galling

chip on flank face

Fig. 7. Adhesion wear mechanism (attrition/galling) on tool A. (a) Rake face (V c = 100 m/min, f z = 0.15 mm/tooth, a a = 2 mm, a r = 8.8 mm). (b) Flank face and rake face (V c = 125 m/min, f z = 0.2 mm/tooth, a a = 2.5 mm, a r = 8.8 mm).

titanium alloys and steels are roughly similar [21]. When machining steels, coating materials such as TiN, TiCN and TiC have a benefic effect on friction and the tool life is generally extended [4,7,8]. So, it can be concluded that coating delamination is mainly due to chemical mechanisms. In the following section and according to the experimental results presented above, a proposal describing chemical wear mechanisms is made and several simulations using Finite Element Analysis were performed to support the analysis.

Table 5 EDAX ViP (variable pressure) analysis on the area located by the arrow number 1 in Fig. 7b Element Tungsten Titanium Tantalum Niobium Carbon Cobalt Symbol W Ti Ta Nb C Co Wt. (%) 40.01 3.14 7.25 15.09 5.59 1.15

Fig. 6. Tool A (V c = 100 m/min, f z = 0.15 mm/tooth, a a = 2 mm, a r = 8.8 mm). (a) Brittle fracture: flaking, chipping. The white dashed line is the original tool profile: plastic deformation. (b) High magnification of the brittle fracture area.

5670

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

(a)

(b)
bakelite

chip

temperature is about 950 -C and its maximum value is located about 50 Am from the tool cutting edge. This is in a good agreement with the results found in literature. Many authors confirmed that the cutting temperature is greater than 800 -C and its maximum occurs close to the cutting edge when machining titanium alloy [14 16,22,23]. The cutting temperature strongly depends on the cutting speed; Kitagawa et al. [23] have reported that the cutting temperature increases from 700 -C to 1100 -C when the cutting speed increases from 60 m/min to 200 m/min. In addition, Dearnley and Grearson [16] have declared that a cutting temperature superior to 800 -C activates the diffusion process when machining titanium alloys. They confirmed that the diffusion wear mechanism occurs during turning of Ti-64 under carbide tools. Their results demonstrate that the worn surface of the tool rake face is very smooth, in other words there is no obvious mechanical wear. In the present study the smooth surface could not be observed because the chips and the adhesive layer covered completely the worn tool surface. Wang and Zhang [24] reported that the Co and C elements migrate from the tool into the chip during machining. The migration of C atoms produces the

(a)
bakelite
cracking

Fig. 8. Adhesion wear mechanism (attrition/galling) on tool B. (a) Plastic deformation, the dashed line is the original tool profile. (b) High magnification of the area located by the white circle showing cracks. The cutting conditions are: cutting speed V c = 125 m/min, feed rate f z = 0.2 mm/ tooth, axial depth of cut a a = 2.5 mm and radial depth of cut a r = 8.8 mm.

chip on rake face

3.3. Chemical wear: diffusion and coating delamination 3.3.1. Diffusion During dry machining, the tool chip contact occurs under extreme conditions such as an intense friction and a high cutting temperature. This supports the activation of the diffusion process and atoms move from the tool towards the chip through the tool chip interface, and vice versa. The cutting temperature is a crucial parameter controlling the diffusion rate of the tool and the chip constituents. It is very hard to obtain information about this parameter only via experimental tests; thus, the finite elements code Thirdwave Advantedge was used. The distribution of temperature at the tool chip interface was calculated; the influence of material parameters and cutting conditions on the diffusion and coating delamination processes is analysed. The numerical simulation result shows that when machining under a cutting speed V c of 100 m/min, a feed rate f z of 0.15 mm/tooth, an axial depth of cut a a of 2 mm, and a radial depth of cut a r of 8.8 mm, the cutting

(b)
bakelite
2

chip on flank face


1

Fig. 9. Adhesion wear mechanism (attrition/galling) on tool B: (a) Rake face. (b) Flank face and rake face. The cutting conditions are V c = 100 m/ min, f z = 0.15 mm/tooth, a a = 2 mm and a r = 8.8 mm.

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

5671

(a)
chip

Table 6 EDAX ViP (variable pressure) analysis on the area located by the arrow number 1 in Fig. 9b Element Tungsten Titanium Tantalum Niobium Carbon Cobalt Symbol W Ti Ta Nb C Co Wt. (%) 64.18 13.24 1.72 19.29 5.15 0.50

coated surface

(b)

area located by arrow number 2 of Figs. 7b and 9b for tool A and tool B, respectively. The results presented in Table 7 for tool A and Table 8 for tool B confirmed that the tool chip interface layer for both tools contains substrate elements such as W, Ti, Ta, Nb, C, and Co. The results indicate that all elements of the tool substrate were found at the tool chip interface. This is in good agreement with the proposal of Wang and Zhang [24] which stipulates that the tool chip interface is a layer rich in carbon atoms. C chemically reacts with the Ti
(a)

Fig. 10. Tool B. (a) Cracking of the coating layer at the leading cutting edge. (b) High magnification of the area located by the black circle. Cutting conditions: 100 m/min, 0.15 m/min, 2 mm and 8.8 mm.

formation of a brittle layer (layer rich in carbon atoms) which easily breaks under mechanical load at the tool chip interface. Diffusion was highlighted by the EDAX analysis. The last one was done at the tool chip interface layer in the

(b)

end of VB3

bakelite

coating layer plucking

coating delamination

Tool B substrate

Fig. 11. Coating delamination of tool B when VB3 = 0.3 mm. The cutting conditions are: 125 m/min, 0.20 mm/tooth, 2.5 mm, 8.8 mm.

Fig. 12. Coating delamination as the initial wear of tool B. (a) New sectioned cutting edge. (b) Worn cutting edge. The cutting conditions are: 115 m/min, 0.165 mm/tooth, 2.25 mm, 8.8 mm. The total contact length is about 250 mm.

5672

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

Table 7 EDAX ViP (variable pressure) analysis on the area located by the arrow number 2 in Fig. 7b Element Tungsten Titanium Tantalum Niobium Carbon Cobalt Symbol W Ti Ta Nb C Co Wt. (%) 41.11 11.48 6.59 3.88 35.38 0.52

element coming from the workpiece material and from coating TiC. As reported by Dearnley and Grearson [20], the TiC layer is harder than the tool substrate WC and acts as a rubbing surface on the tool flank face. In Figs. 7b and 9b a smooth flank surface beneath the adhesive layer can be seen a well as an irregular rake surface in Figs. 7a and 9a. 3.3.2. Coating delamination Delamination of the coating when machining Fdifficultto-cut-materials_ such as titanium alloys, nickel based superalloys and ceramics, has been reported previously by various researchers [9,25 27]. Some have attributed it to chemical reactions, whereas others have confirmed that it is due to a crack propagation at the substrate interface. The evidence of the coating delamination is shown in Figs. 10 12 for different cutting conditions. The removed coating layers at the tool cutting edge can clearly be seen under the SEM. Fig. 11 demonstrates that the coating delamination occurs when the localized flank wear VB3 reaches 0.3 mm while in Fig. 12 this phenomenon is found to be the initial wear mechanism, it happens after a few seconds of cutting time. According to the experimental observations presented above and to the Finite Element Analysis, a proposal on the physical mechanisms causing the coating delamination is presented below. The effect of the tribological parameters (temperature and pressure) and of the chip segmentation on wear mechanisms when machining titanium alloy are discussed. Due to the availability of the limited number of coating layers provided by the software, it is assumed that the coating of tool B is composed of only two layers TiN and TiC with a total thickness of 10 Am. The TiCN is not available in the software package, thus it can be replaced by TiC since their physical properties are similar, see Table 4. The schematic view presented in Fig. 13 gives the explanation of the different stages of the coating delamination process. At the beginning, when the first chip is formed by the leading cutting edge, the coating layer still exists, this situation is described by the illustration number 1 in Fig. 13. At this stage, the contact occurs between three bodies, i.e. the chip, the coating layer, and the tool substrate. The Finite Element Analysis show that the shear angle / and the contact length l c for tools A and tool B are (/ = 32-, l c = 0.18 mm) and (/ = 35-, l c = 0.20 mm),

respectively, see Fig. 14. Note that the cutting conditions for both tools are similar: V c = 100 m/min, f z = 0.15 mm/ tooth, a a = 2 mm and a r = 8.8 mm. As shown in Fig. 15, the maximum cutting temperature for the coated tool B is about 1180 -C while for the uncoated tool A it is about 950 -C. These maxima occur at a distance of 50 Am from the tool cutting edge. It has been reported previously by various researchers that the contact length and the location of the maximum cutting temperature on the tool surface depend strongly on the chip shape (continuous or segmented chip). Chip segmentation is the consequence of the thermoplastic shear localization at the primary shear zone. Simulations also show that at the first stage (before delamination), a cutting pressure of 1600 MPa for tool A and 2000 MPa for tool B take place on the tool surface, and peaks are located at the distances of 350, 300 Am from the tool cutting edge A and B, respectively. The previous work presented in [12] confirmed that the values of the shear angle, the contact length and the cutting temperature are controlled by the cutting conditions, particularly the cutting speed. The last one has the predominant effect on the chip formation. High cutting speeds generate a low shear angle, a short contact length, a high cutting pressure and a high cutting temperature. The current results show that these parameters are different for tools A and B even if the geometrical parameters and the cutting conditions for both tools are similar. Consequently, the only responsible for changes in the tribological conditions and in the tool chip contact is coating. As said before, the high temperature and the intimate contact between the tool chip interface provide an ideal environment for the diffusion of atoms of tool material through the tool workpiece interface. The cutting temperature of tool B (1180 -C) crosses the temperature-threshold that sets off the diffusion process (850 -C) [20], and the pressure on the cutting edge is more significant (2000 MPa). Consequently, the cutting tool loses its hardness and quickly fails [28]. The illustration number 2 (stage 2) in Fig. 13 shows that under extreme mechanical and thermal loads, the asperities are squeezed and the true area of contact approaches the apparent area of contact. This allows assuming a perfect contact at the tool chip interface (adhesive layer). Since the thermal conductivities of the coating constituents are different, an average value of 28 W/mK was assumed in the finite elements calculation. The results of Fig. 16 confirm a significant
Table 8 EDAX ViP (variable pressure) analysis on the area located by the arrow number 2 in Fig. 9b Element Tungsten Titanium Tantalum Niobium Carbon Symbol W Ti Ta Nb C Wt. (%) 46.92 14.38 5.85 3.17 27.92

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

5673

Fig. 13. A schematic view of the coating delamination mechanisms.

reduction in the temperature level through the coating material from 1050 -C to 850 -C; the last one being at the tool substrate surface, see the curve for tool B in Fig. 16. However, the temperature of tool B remains above the temperature of tool A and cannot still be reduced due to the low thermal conductivity of the titanium alloy chip 8 12 W/mK compared to the one of coating material (25 31 W/mK) and tool substrate (45 110 W/mK). The heat flux q flows through the coating layer and penetrates the tool substrate. Diffusion occurs then at the tool substrate surface in which the Co binder elements are not chemically stable; thus, the adhesive surface between the tool substrate surface and the coating layer is gradually removed. At the same time, a similar phenomenon occurs between the coating material and titanium alloy chips. In illustration number 3 (stage 3), the appearance of microcracks in the coating layer can be seen. The evidence of cracking is shown by the micrograph presented in Fig. 10. Micro-cracks randomly propagate and the adhesive surface between the coating layer surface and the tool substrate surface becomes unstable. Finally in stage 4, the coating layer is completely removed on the tool surface as observed in Figs. 11 and 12.

4. Cutting performance: tool life and surface finish The tool life data for both tools A and B recorded during machining tests are illustrated in Fig. 17. A total of 36 data for each tool are analyzed by using a multiple regression method. Taylor laws giving the tool life in terms of cutting conditions (speed, feed rate, etc.) are identified for both tools as follow: T Vc3:3 fz2:97 a2:8 1:767 106 a T Vc3:58 fz1:25 a2:04 68:312 106 : a 3 4

Eqs. (3) and (4) correspond to the uncoated tool A and CVD coated tool B, respectively. Theses two wear laws and Fig. 17 indicate that the performances of both tools are approximately similar. However, when the comparison is made in terms of cutting speed at a tool life of 1 min (calculated using the previous equations), the cutting speed that can be tolerated by tool B (V c = 155 m/min) is higher than the one of tool A (V c = 131 m/min). It means that the performance of tool B is slightly better. The surface roughness was measured by using stylus profilometer on the machined surface of the Ti-6242S

5674

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

(a)
Third Wave AdvantEdge
1.9 1.8 1.7 1.6 1.5
Temperature (C) 1198.48 1118.96 1039.43 959.911 880.389 800.867 721.345 641.822 562.3 482.778 403.256 323.734 244.212 164.689 85.1672

Y (mm)

1.4 1.3 1.2 1.1 1 0.9 4 4.5 5

35

lc 0.20 mm

X (mm)

(b)
Third Wave AdvantEdge
Temperature (C) 1145.89 1070.77 995.661 920.548 845.434 770.321 695.208 620.095 544.982 469.869 394.756 319.643 244.529 169.416 94.3031

1.9 1.8 1.7 1.6

Y (mm)

1.5 1.4 1.3 1.2 1.1 1 0.9 4 4.5 5

32

lc 0.18 mm

X (mm)
Fig. 14. Chip formation by numerical simulations. (a) Tool A. (b) Tool B. The cutting conditions are: 100 m/min, 0.15 mm/tooth, 2 mm, and 8.8 mm.

workpiece produced by both tools. A value of the roughness R a ranging from 0.30 to 0.75 Am is obtained. A high cutting speed, a low feed rate and a low depth of cut are generally the best combination of cutting conditions giving the best surface finish.

5. Conclusions In the present study, the performance and the wear mechanisms of uncoated and CVD coated carbide tools have been investigated when dry end milling titanium

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

5675

Rake Face Data Extraction

1100

Tool B

Tool chip interface temperature (C)

900

Tool A

700

500

300

100 0.5 1 1.5

Distance from cutting edge (mm)


Fig. 15. Evolution of the cutting temperatures on the tool surface. The cutting conditions are similar to those of Fig. 14.

alloy Ti-6242S. The results show that the localized flank wear VB3 on the tool leading cutting edge is the dominant wear determining the tool life of uncoated and multilayer CVD-coated tools. VB3 is due to the tool geometry and to the cutting conditions. The tool cutting edge is subject to extreme thermal and mechanical loads for every engagement of the tool in the machined part during end milling process. The localized flank wear of 0.3 mm is frequently followed by a brittle fracture
TiN TiC 1050 1000 950 Tool substrate

(cracking, flaking and chipping), a plastic deformation and a coating delamination. Adhesion wear is the wear mechanism for both tools A and B; it is caused by the mechanical removal of the tool material parts when the adhesive junctions are broken. Chemical analyses using EDAX HPD and ViP techniques prove the presence of diffusion process between tool substrate, coating and workpiece materials. Diffusion wear mechanism takes place at the interface because of the high

Temperature (C)

900 850 800 750 700 650

Tool B

Tool A
600 5.09 5.1 5.11 5.12 5.13 5.14 5.15 5.16

Distance (mm)
Fig. 16. Evolution of temperature through the coating and the substrate materials. The coatings TiN and TiC are about 5 Am each one on the tool substrate B. For both tools, the cutting conditions are similar to those of Fig. 14.

5676

M. Nouari, A. Ginting / Surface & Coatings Technology 200 (2006) 5663 5676

1000

Experimental data for Tool A Model for Tool A Experimental data for Tool B Model for Tool B
Cutting speed (Vc) (m/min)

for the invited professor program and the continuation of joint research. Moreover, thanks to the Department of Education of Indonesia for the grant through TPSDP Batch-III that awarded the Department of Mechanical Engineering, Faculty of Engineering, University of Sumatera Utara.

References
100
[1] P.S. Sreejith, B.K.A. Ngoi, J. Mater. Process. Technol. 101 (2000) 287. [2] M. Nouari, G. List, F. Girot, D. Coupard, Wear 255 (2003) 1359. [3] H. Schulz, J. Dorr, I.J. Rass, M. Schulze, T. Leyendecker, G. Erkens, Surf. Coat. Technol. 146 147 (2001) 480. [4] W. Schintlmeister, W. Wallgram, J. Kanz, K. Gigl, Wear 100 (1989) 153. [5] V.K. Sarin, Surf. Coat. Technol. 73 (1995) 23. [6] M. Fitzsimmons, V.K. Sarin, Surf. Coat. Technol. 76 77 (1995) 250. [7] K.L. Choy, Prog. Mater. Sci. 48 (2003) 57. [8] C.H. Che Haron, A. Ginting, J.H. Goh, J. Mater. Process. Technol. 116 (2001) 49. [9] A. Jawaid, S. Sharif, S. Koksal, J. Mater. Process. Technol. 99 (2000) 266. [10] M. Fitzsimmons, V.K. Sarin, Surf. Coat. Technol. 137 (2001) 158. [11] E.O. Ezugwu, J. Bonney, Y. Yamane, J. Mater. Process. Technol. 134 (2003) 233. [12] A. Ginting, M. Nouari, J. Mach. Tools Manuf. (submitted for publication). [13] Siekmann, Tool Eng. 34 (1955) 78. [14] R.M. Freeman, PhD thesis, University of Birmingham, UK (1974). [15] P.D. Hartung, B.M. Kramer, CIRP Ann. 31 (1) (1982) 75. [16] P.A. Dearnley, A.N. Grearson, Mater. Sci. Technol. 2 (1986) 47. [17] R.R. Boyer, Mater. Sci. Eng., A 213 (1996) 103. [18] ISO 8688-2, Tool Life Testing in MillingPart 2. End Milling, 1989. [19] A. Ginting, Tool-path generation for tapered machining features, M.Eng thesis, Toyohashi University of Technology (TUT), Japan 1999. [20] P.A. Dearnley, A.N. Grearson, Mater. Sci. Technol. 2 (1986) 47. [21] E.M. Trent, Metal Cutting, Butterworth-Heinemann, 1991. [22] Z.M. Wang, E.O. Ezugwu, Tribol. Trans. 40 (1) (1997) 81. [23] T. Kitagawa, A. Kubo, K. Maekawa, Wear 202 (1997) 142. [24] Z.M. Wang, Zhang, J. Mater. Sci. Technol. 4 (1988) 548. [25] S.S. Cho, K. Komvopoulos, J. Tribol. Trans. ASME 119 (1997) 8. [26] A. Sharman, C.R. Dewes, D.K. Aspinwall, J. Mater. Process. Technol. 118 (2001) 29. [27] H.G. Prengel, W.R. Pfouts, A.T. Santhanam, J. Manuf. Eng. (1996) 82. [28] B.M. Kramer, J. Eng. Ind. 107 (1985) 99.

10 1 10 Tool life (T) (min)


Fig. 17. Tool life vs. cutting speed for tools A and B. The cutting conditions are similar to those of Fig. 14.

100

cutting temperature (950 -C 1180 -C) generated during dry end milling of Ti-6242S. For the multi-layer CVD coated tool, the coating delamination is found to be the initial wear mode, it occurs after a few minutes of cutting time and just after the first passage of the cutting tool on the workpiece surface. Thanks to Finite Element Analyses on the tribological parameters, a new proposal on the physical mechanisms of the coating delamination is done. From all the data that were obtained during dry machining tests, it can be concluded that the best performance of tools A and B with a good compromise in terms of tool life and surface finish is given by a cutting speed of 100 110 m/min. The corresponding tool life is about 6 13 min and the surface finish about 0.4 0.5 Am. Because each insert contains 6 indexes, the total tool life is ranging from 36 to 78 min.

Acknowledgments One of the authors (A. Ginting) would like to thank the LAMEFIP laboratory-ENSAM CER Bordeaux, France

You might also like