You are on page 1of 9

Rev. Tc. Ing. Univ. Zulia. Vol.

37, N 1, 49 - 57, 2014

Optimization of the Cutting Parameters of Ti6Al4V Alloy


by Using Hybrid Approach
A. Marasi
Mechanical and Aerospace Engineering Department, M.A University of Technology, Esfahan, Iran.
mar.abdolali@hotmail.com
Abstract - The prediction of optimal machining conditions for minimum cutting forces plays a very important role in
machining stability, tool life and residual stresses. In this study a series of simulations were run in order to evaluate the
effect of cutting variables and optimize the cutting condition with lower cutting force when orthogonal cutting of titanium
alloy Ti-6Al-4V. A predictive Lagrangian-FEM model based Finite Element Method (FEM) involves Johnson-Cook
material model and fracture criterion was used to simulate chip shapes and cutting forces. The simulation runs plan designed
and carried out based on the central composite design (CCD). A mathematical linear model in terms of cutting speed and
feed rate was developed for cutting force using response surface methodology (RSM). Finally, some numbers of simulation
runs was performed to verify the fitted model which showed its adequacy within 95% prediction interval.
Keywords - FEA model, Surface response mythology, hybrid model, cutting force, central composite design

1 INTRODUCTION
Besides high cutting temperatures and chemical reactivity of titanium alloys in the machining, cutting tools
also suffer from strong high mechanical pressure and high dynamic loads which result in plastic deformation
and/or rapid tool wear. Therefore, various techniques include empirical, mechanistic, analytical, numerical are
applied to model the cutting forces. One of the modeling approaches is empirical modeling which is widely
used in the absence of other meaningful models. Empirical models often utilize statistical methods and they are
only valid for the ranges of the experiments conducted. This technique is based on designing experiments for
varying process inputs and measuring process outputs such as cutting forces, surface roughness and tool-life.
For this reason, this method utilizes heavy experimentation at different cutting conditions and it is very costly.
However, With the enhancement of the digital computer industry, the finite element method (FEM) is
widely used to understand the chip formation mechanism in the multiphysics scale (Usui and Shirakashi 1982;
zel and Altan 2000; zel and Zeren 2007). In early decades, FEM couldn't be proven to be fully capable of
analyzing practical machining processes. One reason for that is the tendency of many researchers to write their
own FEM codes that can handle a specific case which is, mostly, orthogonal cutting (Strenkowski and Moon
1990; Sasahara, Obikawa et al. 1994; Shih 1996; Shinozuka, Obikawa et al. 1996). Even the researchers who
used commercial FEM codes couldnt analyze the practical three-dimensional cutting operations such as ball
end milling and drilling because those codes still need further development to handle such complicated
operations (Hong, Ding et al. 1999; Zhang 1999). Another reason for the limitations of FEM to handle practical
metal cutting operations is the lack of material property data (flow stress) at the high strains, strain rates and
temperatures encountered in metal cutting. Many researchers used flow stress data that made their FEM results
questionable. Stevenson et al. (Stevenson, Wright et al. 1983) modeled the flow stress in the primary
deformation zone in FEM as a function of strain rate at room temperature. Iwata et al. (Iwata, Osakada et al.
1984) assumed the flow stress to be a function of strain alone in their metal cutting FEM simulations.
Nevertheless, in recent years many successful works are reported in analyzing certain work materials such
as titanium alloys by using FEM. In the case of flow stress modeling, Shatla et al. (Shatla, Kerk et al. 2001)
used Oxleys parallel-sided shear zone theory (Oxley 1989) to model orthogonal milling. They modified the
Johnson- Cook material model by inverse analysis and demonstrated applications in numerous materials. Ozel
and Zeren (zel and Zeren 2006) also utilized Oxleys analysis to calculate flow stress at high strain rates. By
combining data from orthogonal machining with SHPB experiments, a JohnsonCook material model was
obtained for AISI 1045 steel, Al 6082-T6 aluminium, and titanium alloy Ti6Al4V. More recently, Shrot and
Baker (Shrot and Bker 2012) used experimental measurements of chip morphology and cutting force in a
Levenberg Marquardt search algorithm to identify the Johnson Cook parameters for high speed machining.
Moreover, the strain softening was identified by carrying out torsion tests at high temperature on pure
aluminium (Kassner, Wang et al. 2002) and on different AlMgSi alloys (Pettersen and Nes 2003). Kassner et al.
(Kassner, Wang et al. 2002) confirm that for pure aluminium, the peak stress is reached at strains less than 0.5.
Increasing the strain further leads to a gradual material softening before a relatively constant level is reached.
This type of flow stressstrain curves has also been obtained for Ti6Al4V titanium alloy (Miller, Bieler et al.
1999). The thermal softening phenomenon has been studied by Nasr et al. (Nasr, Ng et al. 2007) and later by
Ulutan et al. (Ulutan, Sima et al. 2011). Ozel and Ulutan (zel and Ulutan 2012) predicted the residual stresses
49

Rev. Tc. Ing. Univ. Zulia. Vol. 37, N 1, 49 - 57, 2014

using a fully 3D FE model for machining Ti6Al4V titanium and IN100 nickel-based alloys. Calamaz et al.
(Calamaz, Limido et al. 2009) used the Smooth Particle Hydrodynamics (SPH) method in LSDYNA software
to obtain serrated chip formation for orthogonal cutting of titanium alloy Ti6Al4V, and compared the results to
experimental data. Also, the formation of shear localized chips in the cutting of titanium alloy Ti6Al4V has been
simulated by using elastic-viscoplastic FEM with a modified material model using commercial FEM programs
(Sima and zel 2010).
In some other works, it has been tried to used FEM fundamentals to predict industry relevant outcome of
titanium alloys machining. For example, Calamaz et al. (Calamaz, Limido et al. 2009) used Smooth Particle
Hydrodynamics (SPH) to model tool-wear in dry orthogonal cutting of a titanium alloy Ti6Al4V, and compared
the results with experimental data. Uhlmann et al. (Uhlmann, Gerstenberger et al. 2009) proposed a novel
technique based on the finite pointset technique, and compared the results with FEM simulation for the same
cutting conditions. In a similar approach, Al-Athel and Gadala (Al-Athel and Gadala 2011) compared the
Volume of Solid (VOS) technique with the ALE technique. Illoul and Lorong (Illoul and Lorong 2011) used the
Constrained Natural Element Method (CNEM) to simulate orthogonal and 3D oblique cutting processes.
Sasahara et al. (Sasahara, Obikawa et al. 2004) developed a thermo-elasticplastic finite element model to
investigate the effect of cutting parameters, tool rake angle and tool nose radius on residual stresses. They were
later able to also incorporate the effect of tool-edge configurations (Sasahara 2005) and link this to fatigue life.
Ozel et al. (zel, Sima et al. 2010) investigated the effect of multi-layered coated tool inserts in machining of
Ti6Al4V titanium alloy, while Attanasio et al. (Attanasio, Ceretti et al. 2010) observed the evolution of the
crater area when uncoated tools were used for machining mild steels.
The objective of this study is to combine the both FEA model and response surface methodology to model
the effects of cutting parameters on the main cutting force Ti6Al4V when orthogonal cutting. This approach
may tends to decrease the expense of modeling compare to the conventional methods.
2 Finite element aspects and material modelling
2.1 Numerical model of orthogonal machining
The commercial FEA software DEFORM-2DTM, a Lagrangian implicit code, was used to simulate the
orthogonal cutting process of Ti6Al4V Titanium alloy (Deform 2006). FEM model of the orthogonal cutting
process was developed and was composed of the workpiece and the tool. The workpiece was initially meshed
with 8000 iso-parametric quadrilateral elements, while the tool modeled as rigid, was meshed and subdivided
into 1000 elements. A plane-strain coupled thermo-mechanical analysis was performed using orthogonal
assumption.
2.2 Material modeling
To model the thermo-visco plastic behaviour of titanium alloy Ti6Al4V, the JohnsonCook constitutive
equation was employed which can be represented by the following equation:
= +

1 +

0
0

(1)

where is the flow stress; the plastic strain; the strain rate ( 1 ); 0 the reference plastic strain rate ( 1 ); T
the workpiece temperature; (1600 C) the melting temperature of the workpiece material; 0 (20 C) is the
room temperature. Coefficient A (MPa) is the yield strength; B (MPa) the hardening modulus; C the strain rate
sensitivity coefficient; n the hardening coefficient; m is the thermal softening coefficient.
In particular, the material Johnson-Cook constants (Table 1) (Lee and Lin 1998) were found under a constant
strain rate of 2000 1 within the temperature range 7001100 C and a maximum true plastic strain of 0.3. The
room temperature was also taken into account in order to determine the basic mechanical properties.

50

Rev. Tc. Ing. Univ. Zulia. Vol. 37, N 1, 49 - 57, 2014

Table 1 The material Johnson-Cook constants (Lee and Lin 1998)

(MPa)

(MPa)

782.7

498.4

0.028

0.28

105

2.3 Fracture criterion and friction tool/chip interface


In this research, Cockroft and Lathams criterion (Cockroft and Latham 1968) is employed to predict the effect
of tensile stress on the chip segmentation during orthogonal cutting. Cockroft and Latham criterion is expressed
as:

1 =

(2)

where is the effective strain; 1 the maximum principal stress; D is a material constant. Cockroft and
Lathams criterion says that when the integral of the largest tensile principal stress component over the plastic
strain path in Eq. (2) reaches the value of D, usually called damage value, fracture occurs or chip segmentation
starts.
In particular, the critical damage value is calculated for each element under deformation at each time-step
by the program (Deform 2006). Once the damage value in a element reaches the critical one, a crack is initiated
in two steps: (i) this element is deleted with all the parameters related to it, including element connectivity
definition, the strain and stress values; (ii) the rough boundary produced by element deletion is smoothed by
cutting out the considered rough angle and adding new points. Moreover, the utilized software in this research
(Deform 2006) permits to apply the fracture criterion by two different techniques, namely remeshingrezoning
methodology and element deletion feature.
As far as friction modeling is concerned, a simple model based on the constant shear hypothesis, =0 ,
was implemented in the FE-code, being the shear stress, the friction factor and 0 the shear yield stress
obtained as: 0 = 0 / 3. This assumption is based on recent investigation (Filice, Micari et al. 2007), where it
was demonstrated as both cutting forces and chip morphology can well predicted setting the correct friction
coefficient, independently of which friction law is taken into account. On the contrary, friction model becomes
strategically important when other variables are considered, such as shear angle, normal pressure and thermal
effects (Astakhov and Outeiro 2005; Sartkulvanich, Altan et al. 2005), even if either normal pressure or
temperature were found with accuracy using shear model too (Filice, Micari et al. 2007).
2.4 FEA model verification
In this study, the simulated cutting conditions to verify FEA model was identical to those of the
experiments and simulations performed by Umbrello (Umbrello 2008) which consists of dry orthogonal cutting
of Ti6Al4V alloy at cutting speeds of 60 and 120 m/min, feed rate of 0.12 mm and depth of cut of 2.5mm.
Moreover, the D and values verified by him for Ti-6Al-4V alloy were used which are equal to 245 and 0.7
respectively. The cutting tool geometry for WC ISO-P20 consisted of a normal rake angle and a normal flank
angle of 15 and 6, respectively. The tool cutting edge angle was 90, the tool cutting edge inclination angle
was 0 and the tool cutting edge radius of 0.030mm. Accordingly, the FEA model was verified with average
error of 8%, which shows the reliability of the model in order to use it for further investigation of cutting
parameter effects on cutting force.

3 Methodology
51

Rev. Tc. Ing. Univ. Zulia. Vol. 37, N 1, 49 - 57, 2014

3.1 Response surface methodology


Response surface methodology is a collection of mathematical and statistical techniques, which are useful
for the modeling and analyzing the engineering problems and developing, improving, and optimizing processes.
It also has important applications in the design, development, and formulation of new products, as well as in the
improvement of existing product designs, and it is an effective tool for constructing optimization models
(Myers, Montgomery et al. 2009).
RSM consists of the experimental strategy for exploring the space of the process or input factors, empirical
statistical modeling to develop an appropriate approximating relationship between the yield and the process
variables, and optimization methods for finding the levels or values of the process variables that produce
desirable values of the response outputs (Myers, Montgomery et al. 2009). Response surface method designs
also help in quantifying the relationships between one or more measured responses and the vital input factors.
The first step of RSM is to define the limits of the experimental domain to be explored. These limits are made as
wide as possible to obtain a clear response from the model (Kilickap, Huseyinoglu et al. 2011). The cutting
speed, feed rate are the orthogonal cutting variables, selected for this investigation.
In the next step, the planning to accomplish the experiments by means of RSM using a central composite design
(CCD). In many engineering fields, there is a relationship between an output variable of interest (y) and a set off
controllable variables (1 , 2 , , ). The relationship between the cutting control parameters and the responses
is given as:
y= f (1 , 2 , , )+

(3)

where, represents the noise or error observed in the response (y). If we denote the expected response be E(y)=
f (1 , 2 , , )= and then the surface represented by;
= f (1 , 2 , , )

(4)

is called a response surface. The variable 1 , 2 , , in Eq. 4 are called natural variables, because they are
expressed in natural units of measurement. In most RSM problems, the form of the relationship between the
independent variables and the response is unknown, it is approximated. Thus, the first step in RSM is to find an
appropriate approximation for the true functional relationship between response and the set of independent
variables. Usually, a low order polynomial in some region of the independent variables is employed. If the
response is well modeled by a linear function of the independent variables, then the approximating function is
the first order model;
y = 0 +1 1 +2 2 + + +

(5)

If there is curvature in the system, then a polynomial of higher degree must be used, such as the second order
model;
y = 0 +

=1 +

2
=1

(6)

3.2 Central composite design


The necessary data for building the response models are generally collected by the experimental design. In
this study, the collections of experimental data were adopted using central composite design (CCD). The
factorial portion of CCD is a full factorial design with all combinations of the factors at two levels (high, +1 and
low, 1) and composed of the four star points and four central points (coded level 0) which is the midpoint
between the high and low levels. Table 2 shows the levels of two machining parameters and their ranges. The
experimental plans were carried out using the stipulated conditions involving 12 runs as shown in Table 3. The
design was generated and analyzed using Design Expert statistical package.
52

Rev. Tc. Ing. Univ. Zulia. Vol. 37, N 1, 49 - 57, 2014

Table 2 Design scheme of machining parameters and their levels


Parameters

Unit

-1

+1

Cutting speed (Vc)

m/min

47

60

90

120

132

Feed rate (f )

mm/rev

0.10

0.12

0.18

0.24

0.26

Table 3 Design layout and FEA simulation results

Run
number
1
2
3
4
5
6
7
8
9
10
11
12

Cutting
(Vc)
60
120
60
120
47
132
90
90
90
90
90
90

speed Feed rate Cutting


(f)
(F)
0.12
710.14
0.12
679.77
0.24
1378.29
0.24
1278.49
0.18
1073.43
0.18
982.64
0.10
599.87
0.26
1341.59
0.18
978.22
0.18
1005.63
0.18
969.09
0.18
981.88

force

4. Results and discussion


4.1 ANOVA
The statistical significances of the fitted linear model for the cutting force (F) were evaluated by the F-test
of ANOVA, and shown in Table 4.When the values of Prob. > F in Table 4 for the term of models are less
than 0.05 (i.e. = 0.05, or 95% confidence), this indicates that the obtained models are considered to be
statistically significant, which is desirable, as it demonstrates that the terms in the model have a significant
effect on the responses. The other important coefficient, 2 , which is called determination coefficients in the
resulting ANOVA table, is defined as the ratio of the explained variation to the total variation and is a measure
of the degree of fit. When 2 approaches to unity, the better the response model fits the actual data. It exists that
the less is the difference between the predicted and actual values. Furthermore, the value of adequate precision
(AP) in this model, which compares the range of the predicted value at the design point to the average prediction
error, is well above 4 (Noordin, Venkatesh et al. 2004). The value of ratio is greater than 4 which present the
adequate model discrimination. These obtained models present higher values of the determination coefficients
(2 ) and adequate precision (AP) at the same time. These values were obtained as follows: 2 = 0.98 and AP =
43.54 for the model. Consequently, this obtained models can be regard as significant effects for fitting and
predicting the experimental results and meantime it also displays that the test of lack-of-fit is insignificant for
the effects.
Table 4: The ANOVA table for the fitted cutting force models
53

Rev. Tc. Ing. Univ. Zulia. Vol. 37, N 1, 49 - 57, 2014

Source

Sum of Degree of Mean


squares
freedom
square

F value

Model
678732.4
2
339366.2
239.99
Lack of fit
11998.82
6
1999.804
8.24
Residual
12726.45
9
1414.05
Pure error
727.62
3
242.5418
Cor total
691458.9
11
Standard deviation
Mean
Coefficient of variation
Predicted residual error of sum of square (PRESS)
2
2 adjusted
Predicted 2
Adequate Precision (AP)

Prob>F
< 0.0001
0.0558

37.60
998.25
3.76
26965.62
0.98
0.97
0.961
43.54

The backward elimination process eliminates the insignificant terms so as to adjust the fitted linear models.
These insignificant model terms can be removed and it also displays that the test of lack-of-fit is insignificant.
Through the backward elimination process, the final linear models of cutting force equation in terms of actual
factors are presented as follows:
F= 4824.61 f - 1.07Vc + 226.78

(7)

4.2 Effects of machining parameters on the cutting force


Tables 5 shows that the significant and standard errors of each coefficient for the fitted models, which were
also determined by values of F-value and Prob.> F. As can be seen both cutting force and feed rate can be
regards as significant terms due to their Prob. > F value being less than 0.05. However, feed rate has the
highest contribution to the cutting force model. This can be also seen from the response surface of the fitted
model shown in Figure 1. Therefore, feed rate increase has proportional effect on the cutting force due the
increase of unreformed chip thickness and tool load consequently.
Table 5 Results of the analysis of variance for the cutting force ( F )

Source

Sum of Degree of Mean square


squares freedom

F value

Prob>F

cutting speed

8356.225 1

8356.225

5.909427

0.0379

feed rate

670376.2 1

670376.2

474.0824

< 0.0001

Residual

12726.45 9

1414.05

Total

691458.9 11

54

Rev. Tc. Ing. Univ. Zulia. Vol. 37, N 1, 49 - 57, 2014

Fig.1 The response surface of cutting force (F) according to change of cutting speed and feed rate

It can also be noticed from Figure 1 that increase of cutting speed results in reduction of cutting force
owning to the increase of shear plane angle and reduction of chip thickness accordingly. This tends to reduce the
friction force between the chip-tool contact on the rake face. Figure 2 shows the chip morphology of deformed
chip at the highest and lowest level of cutting speed in which the maximum chip thickness is 0.30 and 0.26 mm
for cutting speed of 47 and 132 m/min respectively. In general, it can be concluded that lowest cutting force will
be obtained at highest cutting speed and lowest feed rate.

Figure 2: Chip morphology of deformed chips at feed rate of 0.18 mm/rev and cutting speed of a) 47 m/min
and b) 132 m/min
4.3 Confirmation Runs
The five confirmation runs are simulated in order to compare with predicted values by fitted model and
verify the adequacy of the model (Eqs.7). The results of the confirmation runs and their comparisons with the
predicted values for the cutting force (F) are listed in Table 6. The results of Table 7 reveal that percentage error
are small. The percentage error range between the simulated and the predicted value of cutting force and lies
within 0.50% to 5.7%. All the simulated values of confirmation runs are within the 95% prediction interval. It
obviously demonstrates that the obtained Eqs. 7 is excellently accurate mathematical model.
5 Conclusion
In this study a hybrid approach of FEA model and response surface methodology was utilized in order to
model the effect of cutting parameter on main cutting force of Ti6Al4V alloy. From the results of this study, the
following conclusions can be drawn and recommendations can be made:
The adjusted parameters was good enough for FEA model to simulate the cutting mechanism of Ti6Al4V
alloy with reasonable adequacy.
The fitted model can predict the main cutting force in terms of cutting speed and feed rate with 0.98%
adequacy in the range of selected parameters.
Feed rate changes had the highest contribution to the fitted model while cutting speed effect is just
significant.
Cutting force increases with feed rate and decreases with cutting speed increment, where lowest cutting
force was obtained at highest cutting speed and lowest feed rate level.
55

Rev. Tc. Ing. Univ. Zulia. Vol. 37, N 1, 49 - 57, 2014

Table 6: The results of confirmation runs

Run
No.

Cutting force Feed


rate Simulated
Predicted
Error
(m/min)
(mm/rev)
cutting force cutting force (%)
(N)
(N)

100

0.18

1049.04

988.20

-5.7

80

0.14

797.03

816.62

2.4

110

0.14

788.57

784.52

-0.5

105

0.21

1092.91

1127.47

3.1

70

0.21

1174.99

1165.04

-0.8

References
Al-Athel, K. S. and M. S. Gadala (2011). "The use of volume of solid (VOS) approach in simulating metal cutting with
chamfered and blunt tools." International Journal of Mechanical Sciences 53(1): 23-30.
Astakhov, V. P. and J. C. Outeiro (2005). "MODELING OF THE CONTACT STRESS DISTRIBUTION AT THE TOOLCHIP INTERFACE." Machining Science and Technology 9(1): 85-99.
Attanasio, A., E. Ceretti, et al. (2010). "Investigation and FEM-based simulation of tool wear in turning operations with
uncoated carbide tools." Wear 269(56): 344-350.
Calamaz, M., J. Limido, et al. (2009). "Toward a better understanding of tool wear effect through a comparison between
experiments and SPH numerical modelling of machining hard materials." International Journal of Refractory
Metals and Hard Materials 27(3): 595-604.
Cockroft, M. G. and D. J. Latham (1968). "Ductility and workability of metals." Journal of Institute of Metals 96: 3339.
Deform (2006). User manual. Columbus, OH, USA, Scientific Forming Technologies Corporation Ed.
Filice, L., F. Micari, et al. (2007). "A critical analysis on the friction modelling in orthogonal machining." International
Journal of Machine Tools and Manufacture 47(34): 709-714.
Hong, S. Y., Y. Ding, et al. (1999). "Improving low carbon steel chip breakability by cryogenic chip cooling." International
Journal of Machine Tools and Manufacture 39(7): 1065-1085.
Illoul, L. and P. Lorong (2011). "On some aspects of the CNEM implementation in 3D in order to simulate high speed
machining or shearing." Computers & Structures 89(1112): 940-958.
Iwata, K., K. Osakada, et al. (1984). "Process modeling of orthogonal cutting by the rigid-plastic finite element method."
Transactions of the ASME, Journal of Engineering for Industry 106: 132138.
Kassner, M. E., M. Z. Wang, et al. (2002). "Large-strain softening of aluminum in shear at elevated temperature."
Metallurgical and Materials Transactions A 33(10): 3145-3153.
Kilickap, E., M. Huseyinoglu, et al. (2011). "Optimization of drilling parameters on surface roughness in drilling of AISI
1045 using response surface methodology and genetic algorithm." The International Journal of Advanced
Manufacturing Technology 52(1-4): 79-88.
Lee, W.-S. and C.-F. Lin (1998). "High-temperature deformation behaviour of Ti6Al4V alloy evaluated by high strain-rate
compression tests." Journal of Materials Processing Technology 75(13): 127-136.
Miller, R. M., T. R. Bieler, et al. (1999). "Flow softening during ho tworking of Ti6Al4V with a lamellar colony
microstructure." Scripta Materialia 40(12): 13871393.
Myers, R. H., D. C. Montgomery, et al. (2009). Response Surface Methodology: Process and Product Optimization Using
Designed Experiments, Wiley.
Nasr, M. N. A., E. G. Ng, et al. (2007). "Effects of Workpiece Thermal Properties on Machining-Induced Residual Stresses
Thermal Softening and Conductivity." Proceedings of the Institution of Mechanical Engineers Part B: Journal of
Engineering Manufacture 221: 13871400.
Noordin, M. Y., V. C. Venkatesh, et al. (2004). "Application of response surface methodology in describing the performance
of coated carbide tools when turning AISI 1045 steel." Journal of Materials Processing Technology 145(1): 46-58.
Oxley, P. L. B. (1989). The Mechanics of Machining: An Analytical Approach to Assessing Machinability. Chichester,
England, Ellis Horwood Limited.
zel, T. and T. Altan (2000). "Process simulation using finite element method prediction of cutting forces, tool stresses
and temperatures in high-speed flat end milling." International Journal of Machine Tools and Manufacture 40(5):
713-738.
zel, T., M. Sima, et al. (2010). "Investigations on the effects of multi-layered coated inserts in machining Ti6Al4V alloy
with experiments and finite element simulations." CIRP Annals - Manufacturing Technology 59(1): 77-82.
zel, T. and D. Ulutan (2012). "Prediction of machining induced residual stresses in turning of titanium and nickel based
alloys with experiments and finite element simulations." CIRP Annals - Manufacturing Technology 61(1): 547550.
56

Rev. Tc. Ing. Univ. Zulia. Vol. 37, N 1, 49 - 57, 2014

zel, T. and E. Zeren (2006). "A Methodology to Determine Work Material Flow Stress and ToolChip Interfacial Friction
Properties by Using Analysis of Machining." ASME Journal of Manufacturing Science & Engineering 128: 119.
zel, T. and E. Zeren (2007). "Finite element modeling the influence of edge roundness on the stress and temperature fields
induced by high-speed machining." The International Journal of Advanced Manufacturing Technology 35(3-4):
255-267.
Pettersen, T. and E. Nes (2003). "On the origin of strain softening during deformation of aluminum in torsion to large
strains." Metallurgical and Materials Transactions A 34(12): 2727-2736.
Sartkulvanich, P., T. Altan, et al. (2005). "EFFECTS OF FLOW STRESS AND FRICTION MODELS IN FINITE
ELEMENT SIMULATION OF ORTHOGONAL CUTTINGA SENSITIVITY ANALYSIS." Machining
Science and Technology 9(1): 1-26.
Sasahara, H. (2005). "The effect on fatigue life of residual stress and surface hardness resulting from different cutting
conditions of 0.45%C steel." International Journal of Machine Tools and Manufacture 45(2): 131-136.
Sasahara, H., T. Obikawa, et al. (1994). "FEM analysis on three dimensional cutting." International Journal of Japanese
Society for Precision Engineering 28(2): 473478.
Sasahara, H., T. Obikawa, et al. (2004). "Prediction model of surface residual stress within a machined surface by combining
two orthogonal plane models." International Journal of Machine Tools and Manufacture 44(78): 815-822.
Shatla, M., C. Kerk, et al. (2001). "Process modeling in machining. Part I: determination of flow stress data." International
Journal of Machine Tools and Manufacture 41(10): 1511-1534.
Shih, A. J. (1996). "Finite element analysis of orthogonal metal cutting mechanics." International Journal of Machine Tools
and Manufacture 36(2): 255-273.
Shinozuka, J., T. Obikawa, et al. (1996). "Chip breaking analysis from the viewpoint of the optimum cutting tool geometry
design." Journal of Materials Processing Technology 62(4): 345-351.
Shrot, A. and M. Bker (2012). "Determination of JohnsonCook parameters from machining simulations." Computational
Materials Science 52(1): 298-304.
Sima, M. and T. zel (2010). "Modified material constitutive models for serrated chip formation simulations and
experimental validation in machining of titanium alloy Ti6Al4V." International Journal of Machine Tools and
Manufacture 50(11): 943-960.
Stevenson, M. G., P. K. Wright, et al. (1983). "Further development in applying the finite element method to the calculation
of temperature distribution in machining and comparison with experiment." Journal of Engineering for Industry
105: 149154.
Strenkowski, J. S. and K. J. Moon (1990). "Finite element prediction of chip geometry and tool/workpiece temperature
distributions in orthogonal metal cutting." ASME Journal of Engineering for Industry 112: 313318.
Uhlmann, E., R. Gerstenberger, et al. (2009). "The Finite-Pointset-Method for the Meshfree Numerical Simulation of Chip
Formation." Proceedings of the 12th CIRP Conference on Modelling of Machining Operations 1: 145151.
Ulutan, D., M. Sima, et al. (2011). "Prediction of Machining Induced Surface Integrity using Elastic-Viscoplastic
Simulations and Temperature-Dependent Flow Softening Material Models in Titanium and Nickel-based alloys."
Advanced Materials Research 223 401410.
Umbrello, D. (2008). "Finite element simulation of conventional and high speed machining of Ti6Al4V alloy." Journal of
Materials Processing Technology 196(13): 79-87.
Usui, E. and T. Shirakashi (1982). "Mechanics of Machining from Descriptive to Predictive Theory, On the Art of Cutting
Metals." ASME PED 7: 1335.
Zhang, L. (1999). "On the separation criteria in the simulation of orthogonal metal cutting using the finite element method."
Journal of Materials Processing Technology 8990(0): 273-278.

57

You might also like