You are on page 1of 12

International Journal of Mechanical Sciences 140 (2018) 1–12

Contents lists available at ScienceDirect

International Journal of Mechanical Sciences


journal homepage: www.elsevier.com/locate/ijmecsci

Modeling of plastic deformation induced by thermo-mechanical stresses


considering tool flank wear in high-speed machining Ti-6Al-4V
Xiaoliang Liang a,b, Zhanqiang Liu a,b,∗, Bing Wang a,b, Xin Hou a,b
a
Key Laboratory of High Efficiency and Clean Mechanical Manufacture of MOE, School of Mechanical Engineering, Shandong University, Jinan 250061, PR China
b
Key National Demonstration Center for Experimental Mechanical Engineering Education, Shandong University, Jinan 250061, PR China

a r t i c l e i n f o a b s t r a c t

Keywords: Metallurgical characteristics in the machined surface layer can be modified due to suffering the localized thermo-
Ti-6Al-4V mechanical stresses during high-speed machining. Plastic deformation is generated by the localized thermo-
Plastic deformation mechanical stresses. The depth of plastic deformation will then influence the functional performances and service
Surface integrity
life of the machined components. However, the available tool life is diminished due to rapid tool wear in machin-
Thermo-mechanical stresses
ing titanium alloy. The tool wear induces additional thermo-mechanical stresses on the tool-workpiece interface,
Tool flank wear
which results in deeper plastic deformation. This article proposed a prediction model of the plastic deformation
depth induced by the coupled thermo-mechanical stress considering tool flank wear. The proposed model can
effectively predict the depth of plastic deformation at different tool flank wear stages. This prediction model is
verified with the high-speed turning experiments of Ti-6Al-4V. It is demonstrated that there is a better consis-
tency between the measured and predicted results with the error interval of 11.2% to 15.4%. The results indicated
that the tool flank wear should be limited in an appropriate value from the perspective of the depth of plastic
deformation. This work can be used to guarantee the machined surface integrity during machining Ti-6Al-4V.
© 2018 Elsevier Ltd. All rights reserved.

1. Introduction Several previous research efforts have been devoted to predicting


plastic deformation depth when machining with fresh tools. Thom-
1.1. Literature review sen [9] utilized the shear plane to study the formation mechanism of
plastic deformation. The shear plane was assumed to suffer the shear
In recent years, titanium alloy Ti-6Al-4V has been widely employed stresses and normal stresses. Barash et al. [10] established a math-
in aeronautic and medical manufacturing fields [1,2]. The superior ma- ematical model based on the slip-line field to predict plastic defor-
terial properties of Ti-6Al-4V including high strength-to-weight ratio, mation. The developed model supposed that the plastic deformation
excellent strength at high temperature, as well as relatively good cor- depth was directly related to the shear angle and depth of cut. Park
rosion resistance make it attractive for the applications in gas turbine et al. [11] utilized a numerical method in order to predict the dis-
engine components [3,4]. Medical components such as stents, dental tribution of plastic deformation and plastic strain. Their model sup-
implants, and other devices are also widely utilized titanium alloy Ti- posed that the plastic deformation and strain had the special liner
6Al-4V due to its excellent biological compatibility [5,6]. relationship. The results indicated that the plastic deformation depth
When the critical manufactured components in the aerospace and was depended on the material properties, the shear stress, and the
medical industries are machined with the objectives to improve the func- distribution of plastic strain in the machined subsurface. Based on
tional properties and service life, surface integrity characteristics are the plastic strains distribution, Yang and Liu [12] analyzed the vari-
considered as key factors for evaluating the finishing machined surface ation of the plastic deformation depth near the machined subsurface.
quality, including surface topography, plastic deformation, phase trans- They proposed that the prediction model was involved the Flamant–
formations, microhardness, and residual stress [7]. However, the plastic Boussinesq problem. Their prediction results were verified using the pe-
deformation generation is inevitable in the machined subsurface due to ripheral milling Ti-6Al-4V experiments. Davim et al. [13] utilized the
suffering localized thermo-mechanical stresses. The change of plastic de- finite element (FE) simulation to study the plastic strains and strain
formation depth directly affects the microstructure, microhardness, and rates in the machined surface. There was not little difference between
residual stress in the machined surface [8].


Corresponding author at: Key National Demonstration Center for Experimental Mechanical Engineering Education, Shandong University, Jinan 250061, PR China.
E-mail address: melius@sdu.edu.cn (Z. Liu).

https://doi.org/10.1016/j.ijmecsci.2018.02.031
Received 30 November 2017; Received in revised form 1 February 2018; Accepted 14 February 2018
Available online 14 February 2018
0020-7403/© 2018 Elsevier Ltd. All rights reserved.
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

because of omission of thermo-mechanical stress induced by tool flank


Nomenclature wear.
In addition, tool flank wear occurs inevitably in machining of Ti-6Al-
VB tool flank wear width (mm); 4V, some investigations have been carried out to study the tool wear
𝜓 tool clearance angle (°); effects on the plastic deformation. Che-Haron et al. [14] investigated
𝛾 effective tool rake angle (°); the plastic deformation depth of machined surface with the different
t uncut chip thickness (mm); ranges of prolonged cutting time. They reported that the plastic flow of
w width of cut (mm); microstructure in machined surface did not exist at the initial tool wear
v moving velocity of heat source (mm/s); stage during the cutting process. However, the severe plastic deforma-
Fs shear force (N); tion was produced when the cutting tool reached the failure criteria.
Fp normal force (N); Liang and Liu [15] have studied the tool flank wear effects on the depth
𝜑 shear angle (°); of plastic deformation in dry turning Ti-6Al-4V. There was not appar-
Fcs cutting force component (N); ent plastic deformation produced under machining with fresh tools. The
Fts thrust force component (N); depth of severe plastic deformation and distorted grains could be 48 μm
Fcw cutting force in cutting direction due to tool flank wear in the subsurface material under the condition of worn tool with the
(N); flank wear width 0.3 mm. Yang et al. [16] performed face milling exper-
Ftw cutting force in thrust direction due to tool flank wear iment of titanium alloy. Their results found that the tool flank wear pro-
(N); moted an increase of plastic deformation depth. There was a less depth
Tshear workpiece temperature rise due to primary heat source of plastic deformation produced under machining with fresh tools, while
(°C); severe plastic deformation was generated in the subsurface material at
Trubbing workpiece temperature rise due to rubbing heat source the tool flank wear width reaching 0.35 mm.
(°C);
TM total workpiece temperature rise (°C); 1.2. Challenges and problems description
qshear shear plane heat source intensity (W/mm2 );
qrubbing rubbing heat intensity (W/mm2 ); Ti-6Al-4V is one of the typical difficult-to-machine titanium alloys
L primary heat source length (mm); from the viewpoint of machinability. High chemical activity causes
a thermal diffusivity of workpiece material (mm2 /s); strong affinity with tool and coating materials at high cutting temper-
K0 modified Bessel function of the second kind of order ature, which triggers serious adhesion wear and diffusion wear [17].
zero; Due to the small Young’s modulus of Ti-6Al-4V, the spring back is gen-
k partition of cutting heat conducting into the workpiece; erated at the machined surface. Low thermal conductivity of Ti-6Al-
𝜆t thermal conductivity of tool material (W/(mm°C)); 4V results in higher cutting heat accumulation near the machined sur-
𝜆w thermal conductivity of workpiece material face [18]. The rebound and higher cutting temperature lead to more
(W/(mm °C)); tool flank wear, which exacerbates the generation thermo-mechanical
ct specific heat capacity of tool material (J/(kg °C)); stresses due to frictions between tool flank and machined workpiece
cw specific heat capacity of workpiece material (J/(kg °C)); surfaces. The thermo-mechanical stresses easily cause surface damage,
𝜌t density of tool material (kg/mm3 ); microstructural and mechanical properties alterations along the depth
𝜌w density of workpiece material (kg/mm3 ); gradient of the machined surface [19].
𝜎 s-mech mechanical stress due to fresh cutting edge (MPa); The high-speed machining Ti-6Al-4V diminishes the available tool
𝜎 w-mech mechanical stress due to tool flank wear (MPa); life rapidly, which deteriorates the machined surface [20]. Seen from
𝜎 therm thermal stress (MPa); Fig. 1, the orthogonal cutting process is illuminated considering the tool
p(s) normal force distribution (N); flank wear effects (VB). The tool geometries and contact pattern of the
q(s) tangential force distribution (N); tool-workpiece interface are changed because of tool flank wear effects,
ff maximum magnitude of forces in feed direction due to which produce the practical tool clearance angle 0° and larger contact
fresh tool (N); area [21]. Under the condition of worn tools, the mechanical loads and
fv maximum magnitude of forces in speed direction due to rubbing heat source induced by tool flank wear become significant fac-
fresh tool (N); tors affecting the machined surface. The additional thermo-mechanical
c Half contact length (mm); stresses related to tool flank wear at the tool-workpiece interface result
Rt cutting tool edge radius (mm); in deeper plastic deformation.
Rw radius of workpiece (mm);
Ew elastic modulus of workpiece material (MPa); 1.3. Research objective
Et elastic modulus of tool material (MPa);
𝜎0 maximum normal stress at cutting edge (MPa) This article proposes a model to predict the plastic deformation depth
𝜏0 maximum shear stress at cutting edge (MPa) due to the coupled thermo-mechanical stress associated with tool flank
𝜎s yield strength of workpiece material (MPa) wear. The additional thermo-mechanical stresses induced by tool flank
μ friction coefficient at tool-workpiece interface wear at the tool-workpiece interface are taken into account. As shown in
VB∗ critical tool flank wear width (mm) Fig. 2, the outline of this research is presented. The plastic deformation
G green’s functions under the condition of plane strain as- depth is predicted based on the proposed model. The prediction accu-
sumption racy of this model is verified through high-speed turning experiments of
𝜐 poisson’s rate of the workpiece material Ti-6Al-4V. The research results help to guarantee the required surface
quality of machined components.

2. Plastic deformation prediction model considering tool flank


the finite element simulated and the experimental results. However, wear
the previous researches only studied the depth of plastic deformation
in the use of the fresh sharp tool. The results with fresh tools cannot The machined surface suffers severe localized thermo-mechanical
be applied to analyze the plastic deformation in the machined surface stresses. The critical condition of plastic deformation generation de-

2
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

Fig. 1. Schematic diagram of orthogonal cutting process considering tool flank wear effect, (a) Fresh tool; (b) Worn tool.

Fig. 2. Research outline of predicting depth plastic deformation induced by tool flank wear.

Fig. 3. Schematic of the formation mechanism of plastic deformation consider-


ing tool flank wear effect.
Fig. 4. Cutting force components considering tool flank wear effect, (a) Fresh
tool; (b) Worn tool.
pends on the total stresses which reach the yield strength of Ti-6Al-4V
[22]. Fig. 3 depicts the schematic of generation mechanism of plastic
deformation considering tool flank wear effect. Seen from Fig. 3, the cut- where 𝜑 is the shear angle, Fcs and Fts denote the cutting force compo-
ting temperature and cutting force are affected by tool flank wear states. nent and thrust force component by using the fresh tool, respectively.
The undesirable plastic deformation will be generated in machined sub- The summation of cutting forces induced by the combined effects of
surface because of the coupled effect of thermo-mechanical stresses. sharp tool and tool flank wear are given in Eq. (2).

𝑅𝑐𝑤 = 𝐹 𝑐𝑠 + 𝐹 𝑐𝑤
2.1. Mechanical loads 𝑅𝑡𝑤 = 𝐹 𝑡𝑠 + 𝐹 𝑡𝑤 (2)

The mechanical loads in the orthogonal cutting process are analyzed. where Fcw and Ftw are cutting forces in cutting direction and thrust di-
As shown in Fig. 4, the cutting forces are predicted according to the re- rections considering tool flank wear, respectively.
search of Smithey et al. [23]. As shown in Fig. 4a, the resultant force by
using the fresh tool is calculated at the tool cutting edge. The resultant 2.2. Thermal loads
cutting force is then converted to the shear force Fs and normal force Fp
utilizing geometry relationship in cutting deformation zone, as given in The workpiece temperature rise induced by the primary cutting de-
Eq. (1). formation heat source can be calculated according to the research of Ko-
manduri and Hou [24]. In Fig. 5, this method introduces an imaginary
𝐹 𝑠 = 𝐹 cscos𝜑 − 𝐹 𝑡𝑠 sin 𝜑 heat source and assumes its heat intensity consistent with the primary
𝐹 𝑝 = 𝐹 𝑐𝑠 sin 𝜑 + 𝐹 𝑡𝑠 cos 𝜑 (1) heat source. The workpiece surface is assumed as an insulated boundary

3
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

Fig. 5. Workpiece temperature rise induced by primary heat source.


Fig. 7. Mechanical stresses distribution induced by fresh tool cutting edge.

heat source along the machined surface and moves along with the cut-
ting speed. The imaginary heat source is located at the same position as
rubbing heat source, and also both have same heat intensity.
The workpiece temperature rise at the specific point M (x,y) induced
by rubbing heat source is calculated by Eq. (5).
𝑉𝐵 { [ √ ]}
1 (𝑥−𝑠 )𝑣 𝑣
𝑇 𝑟𝑢𝑏𝑏𝑖𝑛𝑔 = 𝑘𝑞 𝑟𝑢𝑏𝑏𝑖𝑛𝑔 𝑒− 2𝑎 𝐾0 (𝑥 − 𝑠)2 + 𝑦2 𝑑𝑠 (5)
𝜋𝜆𝑤 ∫0 2𝑎
where Trubbing is the workpiece temperature rise considering the rubbing
heat source, s means the integration variable, k is the partition of cut-
ting heat conducting into the workpiece, qrubbing means the rubbing heat
intensity, k and qrubbing are calculated by Eq. (6) [26].

𝜆𝑤𝜌𝑤𝑐𝑤
𝑘= √ √
𝜆𝑤𝜌𝑤𝑐𝑤 + 𝜆𝑡𝜌𝑡𝑐𝑡
Fig. 6. Workpiece temperature rise induced by rubbing heat source. 𝐹 𝑐𝑤𝑣
𝑞𝑟𝑢𝑏𝑏𝑖𝑛𝑔 = (6)
𝑤𝑉 𝐵

condition. The primary shear heat source and imaginary heat source are where 𝜆t is the thermal conductivity of tool material, ct , cw , 𝜌t , and 𝜌w
oblique band heat source moving with the cutting speed. The workpiece mean the specific heat capacity and density of tool and workpiece ma-
temperature rise at the specific point M (x,y) induced by the primary terials, respectively.
heat source is given by Eq. (3). The total workpiece temperature rise TM at the specific point M (x,y)
because of the combined effect of the primary and rubbing heat source
{ [ √ ]
𝑞𝑠ℎ𝑒𝑎𝑟 𝐿 − (𝑥+𝑙 cos 𝜙)𝑣 𝑣 can be as expressed in Eq. (7).
𝑇 𝑠ℎ𝑒𝑎𝑟 = 𝑒 2𝑎 𝐾0 (𝑥 + 𝑙 cos 𝜙)2 + (𝑦 + 𝑙 sin 𝜙)2
2𝜋𝜆𝑤 ∫0 2𝑎
[ √ ]} 𝑇 𝑀 = 𝑇 𝑠ℎ𝑒𝑎𝑟 + 𝑇 𝑟𝑢𝑏𝑏𝑖𝑛𝑔 (7)
𝑣
+ 𝐾0 (𝑥 + 𝑙 cos 𝜙) + (2𝑡 − 𝑙 sin 𝜙 + 𝑦)2 𝑑𝑙
2
(3)
2𝑎
2.3. Mechanical stresses distribution
where Tshear is the workpiece temperature rise induced by primary heat
source, qshear is shear plane heat source intensity, 𝜆w is the thermal con- The total mechanical stresses are the effect results of cutting forces
ductivity of workpiece material, L means the primary heat source length, of the fresh cutting tool and flank wear. The knowledge of contact me-
l is the integration variable, v means the moving velocity of heat source, chanics is utilized to calculate the mechanical stresses distribution [27].
a is the thermal diffusivity of workpiece material, K0 represents the mod- As shown in Fig. 7, the mechanical stresses distribution induced by the
ified Bessel function of the second kind of order zero, t is the uncut chip fresh tool cutting edge can be expressed as Eq. (8) [28,29].
thickness. For orthogonal cutting conditions in this research, L and qshear
2𝑦 𝑐 𝑝(𝑠)(𝑥 − 𝑠)2 2
𝑐
𝑞(𝑠)(𝑥 − 𝑠)3
can be calculated by Eq. (4). 𝜎𝑥𝑠−𝑚𝑒𝑐ℎ = − [ ] 𝑑𝑠 − 𝑑𝑠
𝜋 ∫−𝑐 (𝑥 − 𝑠)2 + 𝑦2 2 𝜋 ∫−𝑐 (𝑥 − 𝑠)2 + 𝑦2 ]2
[
𝑡
𝐿= 𝑐 𝑐
sin 𝜑 2𝑦3 𝑝(𝑠) 2𝑦2 𝑞(𝑠)(𝑥 − 𝑠)
𝜎𝑦𝑠−𝑚𝑒𝑐ℎ = − [ ] 𝑑𝑠 − [ ] 𝑑𝑠
𝐹 𝑠(𝑣 cos 𝛾∕ cos(𝜑 − 𝛾) 𝜋 ∫−𝑐 (𝑥 − 𝑠)2 + 𝑦2 𝜋 ∫−𝑐 (𝑥 − 𝑠)2 + 𝑦2
𝑞𝑠ℎ𝑒𝑎𝑟 = (4)
𝑤𝐿 2𝑦2 𝑐
𝑝(𝑠)(𝑥 − 𝑠) 2𝑦 𝑐 𝑞(𝑠)(𝑥 − 𝑠)2
𝑠−𝑚𝑒𝑐ℎ
𝜏𝑥𝑦 =− [ ] 𝑑𝑠 − [ ] 𝑑𝑠 (8)
where w means the width of cut. 𝜋 ∫−𝑐 (𝑥 − 𝑠) + 𝑦2
2 𝜋 ∫−𝑐 (𝑥 − 𝑠)2 + 𝑦2
As illustrated in Fig. 6, the workpiece temperature rise considering
rubbing heat source can be evaluated according to the work of Huang where 𝜎𝑥𝑠−𝑚𝑒𝑐ℎ , 𝜎𝑦𝑠−𝑚𝑒𝑐ℎ , 𝜏𝑥𝑦
𝑠−𝑚𝑒𝑐ℎ represent the mechanical stress due to

and Liang [25]. The rubbing heat source can be regarded as the band the fresh tool cutting edge in the direction of x, y, and xy, respectively.

4
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

where 𝜎𝑥𝑤−𝑚𝑒𝑐ℎ , 𝜎𝑦𝑤−𝑚𝑒𝑐ℎ , 𝜏𝑥𝑦


𝑤−𝑚𝑒𝑐ℎ represent the mechanical stresses con-

sidering tool flank wear in directions of x, y, and xy, respectively. 𝜎 w (s)


and 𝜏 w (s) are the distribution of normal stress and shear stress, as ex-
pressed by Eqs. (13) and (14).
When the tool flank wear width VB is smaller than the critical tool
flank wear width VB∗ ,
( )4
𝜎𝑤(𝑠) = 𝜎0 1− 𝑠 𝑓 𝑜𝑟 0<𝑠<𝑉𝐵
𝑉𝐵
( √ )
⎧𝜏0 𝑓 𝑜𝑟 0 < 𝑠 < 𝑉 𝐵 1− 𝜎0 𝜏0

𝜏𝑤(𝑠) = ⎨ ( √ ) (13)
⎪𝜇𝜎𝑤(𝑠) 𝑓 𝑜𝑟 𝜏0
𝑉 𝐵 1− 𝜎0 <𝑠<𝑉𝐵

When the tool flank wear width VB is larger than the critical tool
flank wear width VB∗ ,
{
𝜎0 𝑓 𝑜𝑟 0 < 𝑠 < 𝑉 𝐵 −𝑉 𝐵 ∗
𝜎𝑤(𝑠) = ( )4
𝜎0 1− 𝑉𝑠𝐵 𝑓 𝑜𝑟 𝑉 𝐵 −𝑉 𝐵 ∗ < 𝑠 < 𝑉 𝐵

⎧ √
Fig. 8. Mechanical stress induced by tool flank wear. 𝜏0
⎪𝜏0 𝑓 𝑜𝑟 0 < 𝑠 < 𝑉 𝐵 −𝑉 𝐵 ∗ 𝜎0
𝜏𝑤(𝑠) = ⎨ √ (14)
𝜏0
p(s) and q(s) are the normal force and tangential force distribution as ⎪𝜇𝜎𝑤(𝑠) 𝑓 𝑜𝑟 𝑉 𝐵 −𝑉 𝐵 ∗ 𝜎0 <𝑠<𝑉𝐵

given in Eq. (9).
( ) where 𝜎 0 means the maximum normal stress at the cutting edge, 𝜏 0
|𝑠| represents the maximum shear stress at the cutting edge. The value of
𝑝(𝑠) = 𝑓 𝑓 1 −
𝑐 𝜎 0 and 𝜏 0 can be calculated by the work of Smithey et al. [23]. μ means
( )
|𝑠| the friction coefficient at the tool-workpiece.
𝑞(𝑠) = 𝑓 𝑣 1 − (9)
𝑐
2.4. Thermal stresses distribution
where ff and fv mean the maximum magnitudes of cutting forces in the
directions of feed rate and cutting speed as a result of fresh tool cutting As results of the heat sources distribution and cutting heat conduc-
edge. The values of ff and fv are same as the cutting force component tion during high-speed machining, the cutting temperature distribution
Fcs and thrust force component Fts by using the fresh tool. c means the is non-uniform in the machined subsurface. The thermal expansion will
half contact length, as given by Eq. (10). generate thermally-induced strain and deformation in the machined sub-
√ surface material, the thermal stress is then produced. According to the
4𝐹 𝑡𝑠𝑅
𝑐= (10) research of Saif et al. [32,33], the thermal stress distribution can be
𝜋𝐸
evaluated by Eq. (15).
∞(
where Fts is the total normal force due to fresh tool cutting edge, R and E ∞
𝛼𝐸𝑤 𝜕𝑇 𝑀(𝑥, 𝑦) ′ ′
are the equivalent radius and elastic modulus, as expressed by Eq. (11). 𝜎𝑥𝑡ℎ𝑒𝑟𝑚 = − 𝐺𝑥ℎ (𝑥 , 𝑦 )
1 − 2𝜐 ∫0 ∫−∞ 𝜕𝑥
1 1 1 )
= + 𝜕𝑇 𝑀(𝑥, 𝑦) ′ ′
𝑅 𝑅𝑡 𝑅𝑤 + 𝐺𝑥𝑣 (𝑥 , 𝑦 ) 𝑑 𝑥′ 𝑑 𝑦′
𝜕𝑦
1 1 1
= + (11)
𝐸 𝐸𝑡 𝐸𝑤 2𝑦 ∞ 𝑝(𝑡)(𝑡 − 𝑥)2 𝛼𝐸 𝑤𝑇 𝑀(𝑥, 𝑦)
+ 𝑑𝑡 −
𝜋 ∫−∞ ((𝑡 − 𝑥)2 + 𝑦2 )2 1 − 2𝜐
where Rt and Rw mean the radius of tool edge and workpiece, respec-
∞ ∞(
tively. Ew and Et are the elastic modulus of workpiece and tool materials, 𝛼𝐸𝑤 𝜕𝑇 𝑀(𝑥, 𝑦) ′ ′
𝜎𝑦𝑡ℎ𝑒𝑟𝑚 =− 𝐺𝑦ℎ (𝑥 , 𝑦 )
respectively. 1 − 2𝜐 ∫0 ∫−∞ 𝜕𝑥
Due to the coupled effect of thermo-mechanical loads at the tool- )
𝜕𝑇 𝑀(𝑥, 𝑦) ′ ′
workpiece interface, severe adhesion tool wear easily appears in high- + 𝐺𝑦𝑣 (𝑥 , 𝑦 ) 𝑑 𝑥′ 𝑑 𝑦′
𝜕𝑦
speed machining process. However, the friction behaviors of tool- ∞
workpiece interface are different from general sliding friction. Seen from 2𝑦3 𝑝(𝑡) 𝛼𝐸 𝑤𝑇 𝑀(𝑥, 𝑦)
+ 𝑑𝑡 −
Fig. 8, the tool flank wear width can be divided into the sticking and slid- ∫
𝜋 −∞ ((𝑡 − 𝑥)2 + 𝑦2 )2 1 − 2𝜐
ing regions [30,31]. The friction in the sticking region is generated due ∞ ∞(
𝑡ℎ𝑒𝑟𝑚 𝛼𝐸𝑤 𝜕𝑇 𝑀(𝑥, 𝑦) ′ ′
to the material internal shear effect. The shear stress of sticking region is 𝜏𝑥𝑦 =− 𝐺𝑥𝑦ℎ (𝑥 , 𝑦 )
1 − 2𝜐 ∫0 ∫−∞ 𝜕𝑥
related to the shear strength of workpiece material. However, the sliding )
𝜕𝑇 𝑀(𝑥, 𝑦) ′ ′
region obeys friction law determined by the normal stress and friction + 𝐺𝑥𝑦𝑣 (𝑥 , 𝑦 ) 𝑑 𝑥′ 𝑑 𝑦′
𝜕𝑦
coefficient. The mechanical stresses distribution at the specific point M

(x,y) induced by tool flank wear can be expressed as Eq. (12). 2𝑦2 𝑝(𝑡)(𝑡 − 𝑥)
+ 𝑑𝑡 (15)
𝜋 ∫−∞ ((𝑡 − 𝑥)2 + 𝑦2 )2
𝑉𝐵 𝑉𝐵
2𝑦 𝜎𝑤(𝑠)(𝑥 − 𝑠)2 2 𝜏𝑤(𝑠)(𝑥 − 𝑠)3
𝜎𝑥𝑤−𝑚𝑒𝑐ℎ = − [ ]2 𝑑𝑠 − 𝜋 ∫ [ ]2 𝑑𝑠 where 𝜎𝑥𝑡ℎ𝑒𝑟𝑚 , 𝜎𝑦𝑡ℎ𝑒𝑟𝑚 , 𝜏𝑥𝑦
𝑡ℎ𝑒𝑟𝑚 represent the thermal stresses in the direc-
𝜋 ∫0 (𝑥 − 𝑠) + 𝑦2
2 0 (𝑥 − 𝑠)2 + 𝑦2
tion of x, y, and xy, respectively. 𝜐 is the Poisson’s rate of the workpiece
𝑉𝐵 𝑉𝐵
2𝑦3 𝜎𝑤(𝑠) 2𝑦2 𝜏𝑤(𝑠)(𝑥 − 𝑠) material, Gxh , Gxv , Gyh , Gyv , Gyxh , Gxyv represent the Green’s functions
𝜎𝑦𝑤−𝑚𝑒𝑐ℎ = − [ ] 𝑑𝑠 − [ ]2 𝑑𝑠
𝜋 ∫0 (𝑥 − 𝑠)2 + 𝑦2
2 𝜋 ∫0 (𝑥 − 𝑠)2 + 𝑦2 under the condition of plane strain assumption. p(t) means the equiva-
𝑉𝐵
lent surface traction as x = y = 0. The value of p(t) can be calculated by
𝑤−𝑚𝑒𝑐ℎ 2𝑦2 𝜎𝑤(𝑠)(𝑥 − 𝑠) 2𝑦 𝑉 𝐵 𝜏𝑤(𝑠)(𝑥 − 𝑠)3 Eq. (16)
𝜏𝑥𝑦 = [ ] 𝑑𝑠 − [ ]2 𝑑𝑠
𝜋 0∫ 2 𝜋 ∫0
(𝑥 − 𝑠)2 + 𝑦2 (𝑥 − 𝑠)2 + 𝑦2 𝛼𝐸𝑤𝑇 𝑀(𝑥 = 𝑦 = 0)
𝑝(𝑡) = (16)
(12) 1 − 2𝜐

5
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

Table 1
The primary parameters of physical properties of Ti-6Al-4V.

Density (kg/m3 ) Specific heat capacity (J/(kg °C)) Thermal diffusivity (m2 /s) Thermal conductivity (W/(m °C)) Melting point (°C)

4500 611 2.76E-6 7.6 1650

Table 2
The primary parameters of mechanical properties of Ti-6Al-4V.

Elastic modulus (GPa) Tensile strength (MPa) Shear strength (MPa) Poisson’s ratio Hardness (HV0.05 )

112 902 485 0.34 305

Fig. 10. Schematic of orthogonal cutting experiment setup.

plastic deformation generation can be defined as that the equivalent


stress reaches the yield strength exactly, as expressed in Eq. (19).
√ √
2 ( )
(𝜎𝑥 − 𝜎𝑦)2 + (𝜎𝑦 − 𝜎𝑧)2 + (𝜎𝑧 − 𝜎𝑥)2 + 6 𝜏𝑥𝑦
2 + 𝜏2 + 𝜏2
𝑦𝑧 𝑧𝑥 = 𝜎𝑠
2
(19)

where 𝜎 s means the yield strength of workpiece material.


It is worth nothing that the direction of plastic deformation depth
is along the y-axis. Thus the plastic deformation depth is expressed by
Eq. (20).
|
𝑑max = 𝑦|𝜎𝑒𝑞𝑢𝑖=𝜎𝑠 (20)
Fig. 9. Depth of plastic deformation calculation considering tool flank wear. |
As shown in Fig. 9, the global frame of the proposed method for the
2.5. Total stresses distribution purpose of predicting the plastic deformation depth is illustrated. Firstly,
the thermo-mechanical loads are evaluated by Eqs. (1)–(7) with the in-
The total stresses are superposed by the mechanical stresses and ther- putted relevant parameters. Secondly, the thermo-mechanical stresses
mal stresses as expressed by Eq. (17). distribution can be calculated based on the plane strain assumption, as
expressed by Eqs. (8)–(18). Finally, the depth of plastic deformation is
𝜎𝑥 = 𝜎𝑥𝑠−𝑚𝑒𝑐ℎ + 𝜎𝑥𝑤−𝑚𝑒𝑐ℎ + 𝜎𝑥𝑡ℎ𝑒𝑟𝑚 calculated by Eqs. (19)–(20), which is determined by the von-Mises yield
𝜎𝑦 = 𝜎𝑦𝑠−𝑚𝑒𝑐ℎ + 𝜎𝑦𝑤−𝑚𝑒𝑐ℎ + 𝜎𝑦𝑡ℎ𝑒𝑟𝑚 criterion. In this work, the total algorithm has been implemented with
the programming software Matlab R2012b.
𝑠−𝑚𝑒𝑐ℎ 𝑤−𝑚𝑒𝑐ℎ 𝑡ℎ𝑒𝑟𝑚
𝜏𝑥𝑦 = 𝜏𝑥𝑦 + 𝜏𝑥𝑦 + 𝜏𝑥𝑦 (17)

Based on the assumption of plane strain condition, the stresses in the 3. Materials and cutting experiments
third direction z can be calculated by Eq. (18).
3.1. Machining experiments setup
𝜎𝑧 = 𝜐(𝜎𝑥 + 𝜎𝑦 ) − 𝛼𝐸𝑤𝑇 𝑀
𝜏𝑧𝑥 = 𝜏𝑧𝑦 = 0 (18) Ti-6Al-4V disk with the diameter 100 mm was utilized as the work-
piece material in machining experiments. The chemical element com-
2.6. Determination of the plastic deformation depth positions of Ti-6Al-4V are as follows (wt.%): 5.2–6.7% Al, 3.2–4.6%
V, < 0.1% C, < 0.1% Fe, < 0.07% O, < 0.05% N, < 0.01% H and balance
The workpiece material can be considered isotropic and obeys the Ti. The primary parameters of physical and mechanical properties of
von-Mises yield criterion. In such case, the critical condition for initial Ti-6Al-4V disk are presented in Tables 1 and 2.

6
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

Fig. 11. Measurement of cutting edge radius Rt .

Table 3
Physical and mechanical properties of cutting tools material.

Density (kg/m3 ) Specific heat capacity (J/(kg °C)) Thermal conductivity (W/(m °C)) Elastic modulus (GPa) Poisson’s ratio

15,700 520 84 705 0.24

As illustrated in Fig. 10, the machining experiments were carried out


on PUMA 200 M turning lathe. The tool inserts (Kennametal NG3125R
K313) with effective rake angle 𝛾 = 0° and inclination angle 𝜆 = 0° were
used in the orthogonal turning experiments. As shown in Table 3, the
primary parameters of physical and mechanical properties of the cutting
tools material used in the experiment were presented. The plane strain
assumption condition was guaranteed due to the width of tool cutting
edge (3.125 mm) being wider than the width of cut (2 mm). The whole
experiments were performed in dry conditions without using the exter-
nal coolant fluids. The cutting forces were analyzed through three com-
ponent piezoelectric dynamometer Kistler 9129AA. The constant cutting
parameters were selected at the cutting speed 200 m/min and the feed
rate 0.1 mm/rev.

3.2. Measurement methods

The metallographic examination was measured by scanning electron


microscope (SEM). The X-Ray diffractometer (XRD) was utilized to an- Fig. 12. Measurement method of shear angle 𝜑.
alyze the phase compositions of Ti-6Al-4V. The confocal laser scanning
microscope (CLSM) was used to quantitatively measure the tool flank
wear profiles and cutting edge radius. Seen from Fig. 11, according to
the cutting edge measured profiles, the cutting edge radius Rt was ap-
proximately 15 μm through multiple fitting results. The measurement
method of shear angle 𝜑 was presented in Fig. 12. This method can calcu-
late the shear angle 𝜑 based on the obtained chip shear band and cutting
tool angles. The shear angle 𝜑 was defined as 45° according to average
values of multiple measurement results. The experimental specimens
were sectioned by electric discharge machining (EDM) and mounted in
polyester resin. The specimens were polished, and then chemical etching
was performed with the corrosive liquid containing 5 ml HNO3 + 3 ml
HF + 100 ml H2 O. As shown in Fig. 13, the split Hopkinson pressure bar
(SHPB) technique was used to analyze the dynamic mechanical proper-
ties of titanium alloy Ti-6Al-4V. The test specimens were prepared by
wire-EDM, the top and bottom surface were polished to guarantee test
requirements. The chemical element compositions in tool wear regions
were examined by the energy-dispersive spectrometer (EDS). Fig. 13. Split Hopkinson pressure bar setup.

7
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

Fig. 14. Typical microstructure and phase compositions of workpiece material Ti-6Al-4V, (a) Microstructure; (b) Phase compositions.

could be considered negligible at the beginning of machining process


because of the very small wear value. In addition, Fig. 16 indicates that
the non-uniform wear land with the practical clearance angle 0° format-
ted at the tool flank face.
The wear topographies of tool flank face at VB reaching 0.3 mm are
illustrated in Fig. 17. As the results of higher thermo-mechanical loads
and strong affinity, the workpiece material Ti-6Al-4V easily adheres to
the tool flank face. As shown in Figs. 17b and 17c, the severe adhe-
sive layers are clearly visible at the tool-workpiece interface. Based on
the energy-dispersive spectrometer (EDS) analysis, the chemical element
compositions of adhesive layers are shown in Fig. 17d, which proved to
be the workpiece material Ti-6Al-4V.
Seen from Fig. 18, the evolutions of cutting force component (Fx)
and thrust force component (Fy) are presented at the different tool
flank wear states. In present work, the average values of cutting force
components were taken as the responses. It is remarkable that both
cutting force component and thrust force component have the growth
trend with the development of tool flank wear. Under the condition of
worn tool at VB = 0.3 mm, Fx and Fy increased by 54% and 246%, re-
Fig. 15. True stress-true strain responses of titanium alloy Ti-6Al-4V.
spectively, compared with the fresh tool. The rapid growth of thrust
force component at the VB = 0.3 mm indicates that more severe adhe-
4. Results and discussions sion wear occurs because of the severe tool wear. In this paper, the
critical tool flank wear width VB∗ was defined as 0.25 mm to ensure the
4.1. Material properties cutting forces Fcw and Ftw consistency with the experimental results. As
shown in Fig. 18, the cutting force components are predicted accord-
Fig. 14 presents the typical metallurgical characteristics and phase ing to the sticking-sliding friction model. Analysis results indicate that
composition of workpiece material Ti-6Al-4V. Seen from Fig. 14a, Ti- the predicted cutting forces are well consistent with the measured cut-
6Al-4V is one of 𝛼 + 𝛽 dual-phase titanium alloys. The white strip zones ting forces. Based on the measurements of cutting force components, the
are 𝛽-phases while the gray slice zones represent 𝛼-phases. The initial value of friction coefficient μ can be calculated by Eq. (21) [34].
XRD pattern of titanium alloy Ti-6Al-4V is shown in Fig. 14b. The vol- 𝐹 x sin 𝛾 + 𝐹 y cos 𝛾
𝜇= (21)
ume fraction of 𝛼-phase and 𝛽-phase are estimated from the integrated 𝐹 x cos 𝛾 − 𝐹 y sin 𝛾
peak area. The volume fraction of 𝛼-phase and 𝛽-phase are 91.7% and Here 𝛾 is the tool rake angle in orthogonal cutting, 𝛾 = 0°. Fx and
8.3%, respectively. Fy mean cutting force component and the thrust force component by
Fig. 15 presents the true stress-true strain responses of Ti-6Al-4V at using the fresh tool, respectively. Based on the experimental results, the
various strain rates under the condition of room temperature. The true friction coefficient μ was defined as 0.65.
stress-true strain responses descript the mechanical properties of Ti-6Al-
4V under high strain as well as high strain rates. As shown in Fig. 15, the 4.3. Model verification of plastic deformation depth
yield strength of Ti-6Al-4V is enhanced when the strain rates increase.
Similarly, the yield strength of Ti-6Al-4V is enhanced on account of the As illustrated in Fig. 19, the microstructures of machined surface at
machining process, especially in the case of higher strain rates resulting the six kinds of different tool flank wear condition are shown. Along
from high-speed machining and tool flank wear. In present work, the the direction of grain deformation, some asymptote lines are utilized to
yield strength of Ti-6Al-4V is defined as 1100 MPa. analyze the degree of the plastic deformation. Seen from Fig. 19, the
microscopic characteristics at the plastic deformation regions are differ-
4.2. Tool wear and cutting forces ent from those in the bulk material. Several specific variations emerge
in the plastic deformation regions, such as grains distortions and elon-
Tool failure criterion is set according to ISO 3685 standard, and the gation, grain boundary deformation, and slip lines along the direction
average tool flank wear width (VB) is measured to be 0.3 mm. Seen from of the cutting speed.
Fig. 16, the six different tool flank wear stages are selected to for quanti- As shown in Fig 19a, the plastic deformation is hard to capture by
tative analysis the experimental machining results. The tool flank wear employing the fresh tool since it mainly appears in extreme narrow re-

8
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

Fig. 16. Tool flank wear states, (a) Fresh tool; (b) VB = 0.06 mm; (c) VB = 0.12 mm; (d) VB = 0.18 mm; (e) VB = 0.24 mm; (f) VB = 0.3 mm. The red curves represent
the profiles of tool flank wear as measured from the cross sections of cutting tools.

Fig. 17. Wear topographies of the tool flank face at VB = 0.3 mm. (a) Lens: X 150; (b) Lens: X 500; (c) Lens: X 500; (d) EDS analysis.

gions of the subsurface material. Based on the obtained the microstruc- Fig. 20 presents the comparative analysis of the measured and pre-
ture images, the depth of plastic deformation in subsurface material dicted plastic deformation values at different tool flank wear states. Av-
presents increasing trend at the tool flank wear states varying 0–0.3 mm. erage values of the depth of plastic deformation are selected as the re-
It indicates that the tool flank wear states have a major contribution to sponses. The error analysis between measured and predicted values is
the changes in the depth of plastic deformation. In addition, the intensity calculated, as given by Eq. (22).
of plastic deformation gradually reduces along the top machined sur- |𝑑(𝑝) − 𝑑(𝑚)|
face to the bulk material. This result is mainly because the plastic strain %(𝑒𝑟𝑟𝑜𝑟) = × 100 (22)
𝑑(𝑚)
obeys the first order exponential distribution, and the plastic strain has where %(error) is the percentage of error between the measured and
the maximum value at the machined surface and reduces rapidly along predicted the depth of plastic deformation values, d(p) represents the
the direction of perpendicular to the machined surface [12, 22]. predicted values based on the proposed model, and d(m) are the mea-
sured values.

9
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

Fig. 18. Evolutions of cutting force components under different tool flank wear states.

Fig. 19. SEM images of microstructure under different tool flank wear states. (a) Fresh tool; (b) VB = 0.06 mm; (c) VB = 0.12 mm; (d) VB = 0.18 mm; (d) VB = 0.24 mm;
(d) VB = 0.3 mm.

10
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

flank wear, which is fitted with a higher R-squared value of R2 = 0.9964.


The plastic deformation depth shows a slow increasing as the VB ranges
0–0.18 mm. However, a rapid growth is presented with the VB exceed-
ing 0.24 mm, especially VB = 0.3 mm.
Based on the experimental and predicted values, the evolution mech-
anisms of plastic deformation considering the tool wear effects are dis-
cussed in following section. The cutting forces increase slowly with
the VB ranging 0–0.18 mm. The additional lower thermo-mechanical
stresses induced by tool flank wear between the tool and workpiece sur-
faces result in the smaller changes in the depth of plastic deformation.
However, the cutting force components increase rapidly due to the de-
terioration of tool wear at the VB = 0.3 mm. Thus material Ti-6Al-4V
suffers the higher local thermo-mechanical stresses at the machined sur-
face. More slip systems in the subsurface material are activated under
machining with worn tools. Thus the plastic deformation depth presents
a rapid growth trend at the VB = 0.3 mm.
The changes of plastic deformation depth closely related to the ma-
terial properties of Ti-6Al-4V. Due to the low thermal conductivity,
Fig. 20. Comparative analysis of the measured and predicted the depth of plas-
the cutting heat is difficult to conduct into the bulk material. The re-
tic deformation.
sults cause severe cutting heat accumulation in the subsurface material,
which also changes of the ability of plastic flow of Ti-6Al-4V under the
condition of tool flank wear increasing. The coupling effects of thermo-
mechanical stresses are more significant in high-speed machining with
worn tools. The additional mechanical stresses result in the deeper plas-
tic deformation due to the greater ranges of softened machined surface
material [35].

4.5. Research values and expected application

This work has significant scientific and engineering values in


aerospace and medical titanium alloy Ti-6Al-4V components produc-
tion. On one hand, the tool flank wear easily induces severe plastic
deformation in the substrate material. Suitable tool flank wear values
should be limited in the viewpoint of getting better surface integrity.
The proposed prediction model can provide an instruction for optimiz-
ing the cutting parameters and tool flank wear states to enhance the
Fig. 21. Relationships among the cutting force, plastic deformation, and tool surface quality of the manufactured components. On another hand, the
flank wear. results can help get deeper insights into understanding the evolution
mechanisms of plastic deformation during the high-speed machining Ti-
6Al-4V under the condition of worn tools.
The comparative analysis indicates that the maximum error of plastic
deformation depth is 15.4% measured at the VB = 0 mm, and the mini-
mum error is 11.2% measured at the VB = 0.18 mm. The results indicate
that there is a better consistency between the measured and predicted 5. Conclusions
plastic deformation depth values in this research. In practical engineer-
ing process, these errors can be acceptable considering the measured and This article proposed a method to predict plastic deformation depth
predicted uncertainties. These errors maybe attributed to excluding the in the subsurface material by using the coupled thermo-mechanical
existence of variables in the predicted model. For examples, the changes stresses considering tool flank wear. Some important conclusions can
of tool cutting edges, the tool crater wear, and the material properties at be summarized as follows:
different temperatures are omitted during the prediction program. The
complex contact pattern between the tool-workpiece interfaces may be (1) In the proposed method, the additional thermo-mechanical stresses
another important factor. induced by tool flank wear at the tool-workpiece interface were
taken into account. The validity of prediction method was verified
4.4. Relationships among the cutting force, plastic deformation, and tool through high-speed turning experiments of Ti-6Al-4V. There was a
flank wear better consistency between the measured and predicted plastic de-
formation values within the error interval of 11.2% to 15.4%.
Seen from Fig. 21, the relationships among the cutting forces, the (2) Experimental results indicated that tool flank wear states had a ma-
plastic deformation, and the tool flank wear are presented. Experimen- jor contribution to the changes in the depth of plastic deformation.
tal results demonstrate that the cutting force components and plastic The evolution mechanisms of plastic deformation were discussed
deformation depth have similar trends as the tool flank wear increases. with the analyses of the relationship among the tool flank wear
The regression analysis can be carried out between the plastic defor- states, the localized thermo-mechanical stresses and the material
mation depth and tool flank wear values. R-squared statistic (R2 ) is the properties of Ti-6Al-4V.
index of the fitting degree, and the value closer to R2 = 1 indicates that (3) The proposed depth of plastic deformation prediction method has
the reliability of fitting line is higher. As shown in Fig. 21, the exponen- an instruction for guaranteeing the surface integrity of machined
tial relationship exists between the depth of plastic deformation and tool titanium alloy Ti-6Al-4V components.

11
X. Liang et al. International Journal of Mechanical Sciences 140 (2018) 1–12

Acknowledgments [15] Liang X, Liu Z. Experimental investigations on effects of tool flank wear on sur-
face integrity during orthogonal dry cutting of Ti-6Al-4V. Int J Adv Manuf Technol
2017;93:1617–26.
The authors would like to acknowledge the financial support [16] Houchuan Y, Zhitong C, ZiTong Z. Influence of cutting speed and tool wear on the
from the National Natural Science Foundation of China (51425503, surface integrity of the titanium alloy Ti-1023 during milling. Int J Adv Manuf Tech-
51375272, 51705293), and the Major Science and Technology Program nol 2015;78:1113–26.
[17] Fan Y, Hao Z, Zheng M, Yang S. Wear characteristics of cemented carbide tool in
of High-end CNC Machine Tools and Basic Manufacturing Equipment dry-machining Ti-6Al-4V. Mach Sci Technol 2016;20:249–61.
(2014ZX04012014). This work was also supported by grants from Tais- [18] Pramanik A. Problems and solutions in machining of titanium alloys. Int J Adv Manuf
han Scholar Foundation (TS20130922). Technol 2014;70:919–28.
[19] Meng B, Fu MW, Shi SQ. Deformation behavior and microstructure evolution in ther-
mal-aided mesoforming of titanium dental abutment. Mater Des 2016;89:1283–93.
References [20] Wang B, Liu Z. Evaluation on fracture locus of serrated chip generation with stress
triaxiality in high speed machining of Ti6Al4V. Mater Des 2016;98:68–78.
[1] Ulutan D, Ozel T. Machining induced surface integrity in titanium and nickel alloys: [21] Ducobu F, Arrazola PJ, Rivière-Lorphèvre E, Filippi E. Finite element prediction of
a review. Int J Mach Tools Manuf 2011;51:250–80. the tool wear influence in Ti6Al4V machining. Procedia CIRP 2015;31:124–9.
[2] Thakur A, Gangopadhyay S. State-of-the-art in surface integrity in machining of nick- [22] Zhang P, Liu Z. Plastic deformation and critical condition for orthogonal machining
el-based super alloys. Int J Mach Tools Manuf 2016;100:25–54. two-layered materials with laser cladded Cr-Ni-based stainless steel onto AISI 1045.
[3] Ezugwu EO. Key improvements in the machining of difficult-to-cut aerospace super J Clean Prod 2017;149:1033–44.
alloys. Int J Mach Tools Manuf 2005;45:1353–67. [23] Smithey DW, Kapoor SG, DeVor RE. A new mechanistic model for predicting worn
[4] M’Saoubi R, Axinte D, Soo SL, Nobel C, Attia H, Kappmeyer G, Engin S, Sim W. High tool cutting forces. Mach Sci Technol 2001;5:23–42.
performance cutting of advanced aerospace alloys and composite materials. CIRP [24] Komanduri R, Hou ZB. Thermal modeling of the metal cutting process: part
Ann-Manuf Technol 2015;64:557–80. I—temperature rise distribution due to shear plane heat source. Int J Mech Sci
[5] Qiu C, Fones A, Hamilton HGC, Adkins NJE, Attallah MM. A new approach to de- 2000;42:1715–52.
velop palladium-modified Ti-based alloys for biomedical applications. Mater Des [25] Huang Y, Liang SY. Modelling of the cutting temperature distribution under the tool
2016;109:98–111. flank wear effect. Proc Inst Mech Eng C-J Mech E 2003;217:1195–208.
[6] Niinomi M, Boehlert CJ. Titanium alloys for biomedical applications. Adv Meta Bio- [26] Grzesik W, Nieslony P. A computational approach to evaluate temperature and
mater 2015:179–213. heat partition in machining with multilayer coated tools. Int J Mach Tools Manuf
[7] M’Saoubi R, Outeiro JC, Chandrasekaran H, Dillon OW, Jawahir IS. A review of 2003;43:1311–17.
surface integrity in machining and its impact on functional performance and life of [27] Johnson KL. Contact mechanics. Cambridge University Press; 1987.
machined products. Int J Sustain Manuf 2008;1:203–36. [28] Lazoglu I, Ulutan D, Alaca BE, Engin S, Kaftanoglu B. An enhanced analytical model
[8] Wang Q, Liu Z, Yang D, Mohsan AUH. Metallurgical-based prediction of stress-tem- for residual stress prediction in machining. CIRP Ann-Manuf Technol 2008;57:81–4.
perature induced rapid heating and cooling phase transformations for high speed [29] Yan L, Yang W, Jin H, Wang Z. Analytical modeling of the effect of the tool flank
machining Ti-6Al-4V alloy. Mater Des 2017;119:208–18. wear width on the residual stress distribution. Mach Sci Technol 2012;16:265–86.
[9] Thomsen EG. Application of the mechanics of plastic deformation to metal cutting. [30] Waldorf D, Kapoor SG, DeVor RE. Worn tool forces based on ploughing stresses.
Ann CIRP 1966;14:113. Trans North Am Manuf Res Inst SME 1999;27:165–70.
[10] Barash MM, Schoech WJ. A semi-analytical model of the residual stress zone in [31] El-Zahry RM. On the hydrodynamic characteristics of the secondary shear zone in
orthogonal machining. In: Proceedings of the 5th International M.T.D.R. Conference; metal machining with sticking-sliding friction using the boundary layer theory. Wear
1970. p. 603–13. 1987;115:349–59.
[11] Park YW, Cohen PH, Ruud CO. The development of a mathematical model for pre- [32] Saif MTA, Hui CY, Zehnder AT. Interface shear stresses induced by non-uniform
dicting the depth of plastic deformation in a machined surface. Mater Manuf Process heating of a film on a substrate. Thin Solid Films 1993;224:159–67.
1993;8(6):703–15. [33] Huang K, Yang W. Analytical modeling of residual stress formation in workpiece
[12] Yang D, Liu Z. Surface plastic deformation and surface topography prediction in pe- material due to cutting. Int J Mech Sci 2016;114:21–34.
ripheral milling with variable pitch end mill. Int J Mach Tools Manuf 2015;91:43–53. [34] Braham-Bouchnak T, Germain G, Morel A, Furet B. Influence of high-pressure
[13] Davim JP, Maranhão C. A study of plastic strain and plastic strain rate in machining coolant assistance on the machinability of the titanium alloy Ti555-3. Mach Sci
of steel AISI 1045 using FEM analysis. Mater Des 2009;30:160–5. Technol 2015;19(1):134–51.
[14] Che-Haron CH, Jawaid A. The effect of machining on surface integrity of titanium [35] Liu M, Takagi J, Tsukuda A. Effect of tool nose radius and tool wear on resid-
alloy Ti-6% Al-4% V. J Mater Process Technol 2005;166:188–92. ual stress distribution in hard turning of bearing steel. J Mater Process Technol
2004;150:234–41.

12

You might also like