You are on page 1of 14

Journal of Manufacturing Processes 58 (2020) 1304–1317

Contents lists available at ScienceDirect

Journal of Manufacturing Processes


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

Three-dimensional turning force prediction based on hybrid finite element


and predictive machining theory considering edge radius and nose radius
Tao Zhou a, Lin He a, b, *, Zhongfei Zou a, Feilong Du a, c, Jinxing Wu a, Pengfei Tian a
a
College of Mechanical Engineering, Guizhou University, Guiyang, China
b
Liupanshui Normal College, Liupanshui, China
c
Key Laboratory of Advanced Manufacturing Technology of Ministry of Education, Guizhou University, Guiyang, China

A R T I C L E I N F O A B S T R A C T

Keywords: Cutting force provides a basis for predicting tool wear and fracture, surface quality of component, as well as
Three-dimensional oblique cutting vibration and power demand of machine tools during the process of turning. Based on the finite element and
Cutting force prediction mechanical analysis model, a three-dimensional (3-D) cutting force prediction model with consideration given to
Hybrid FEM-oxley method
the edge radius and the nose radius of cutting tool is proposed in this paper. Firstly, combined with cutting force
Johnson-cook material constitutive equation
Edge radius
experiment, 3-D cutting simulation is performed in Deform to determine the appropriate Johnson-Cook (J-C)
Nose radius material constitutive model, and the two-dimensional (2-D) orthogonal cutting model is simulated by applying
the determined J-C model. By observing the strain rate distribution in the shear zone, the shear angle is obtained,
while the nonlinear regression model of shear angle, cutting speed and cutting thickness is conducted. Then,
taking into account the impact of nose radius on the cutting process, 3-D oblique cutting is converted into
equivalent orthogonal cutting using the chip flow angle and the equivalent cutting edge. The method used to
determine the strain rate coefficient of the second deformation zone is redefined in the modified Oxley model,
and the chip forming force is calculated. Based on the Waldorf slip line field model, the plowing force generated
by the edge radius is predicted. Finally, the cutting force of 3-D oblique turning is obtained by means of coor­
dinate transformation. The proposed model is validated by the excellent consistency between the theoretical
predictions and the results obtained after referencing literature data and performing experiment of cutting 304
stainless steel and Inconel 718.

1. Introduction and cutting parameters can be established by multi-factor orthogonal


regression test. However, a large amount of data is needed to test
As a sort of primary processing technology, turning machining has different processing scenarios, which is time-consuming and costly. The
been widely applied in aerospace and nuclear power equipment due to analytical model is more computationally efficient in predicting the
its excellent versatility and high efficiency. Cutting forces could have a processing variables and quantifying the correlation between the input
significant impact on the optimization of the cutting process, especially variables and the response values [1], which makes the cutting mech­
in the fine turning of some critical components. The modeling of cutting anism more explicit. In the numerical model, the finite element simu­
forces is considered as the basis for the prediction of tool wear and lation method (FEM) is commonly applied. The effective FEM is capable
fracture, part surface quality, machine vibration and power re­ to address the limitations on the empirical model [2] or the analytical
quirements. Therefore, in order to ensure the accuracy of prediction for model [3].
cutting force, there is a necessity to construct a reliable cutting force With the material properties, tool geometry and machining process
model. parameters as input, the finite element model can calculate the physical
The calculation of cutting force is classed into three major categories, parameters such as cutting force and cutting temperature during the
which are empirical model, analytical model and numerical model. The cutting process. Maurel-Pantel et al. [4] applied the finite element
empirical model is premised on the principle of general statistics and commercial software LS-Dyna to simulate the machining of 304 stainless
regression analysis. The exponential relationship between cutting forces steel by shoulder milling. The cutting force was predicted by taking the

* Corresponding author at: College of Mechanical Engineering, Guizhou University, Guiyang, China.
E-mail addresses: helin6568@163.com, lhe@gzu.edu.cn (L. He).

https://doi.org/10.1016/j.jmapro.2020.09.034
Received 6 October 2019; Received in revised form 22 August 2020; Accepted 13 September 2020
Available online 25 September 2020
1526-6125/© 2020 The Society of Manufacturing Engineers. Published by Elsevier Ltd. All rights reserved.
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

J-C equation as the material model, while the accuracy of the model was which is normally assumed to be sharp at the edge of the tool, the three
validated by referencing the experimental data. Based on Deform-3D, parameters of the iterative process play a crucial role in the prediction
Sahho et al. [3] simulated the orthogonal cutting Ti-6Al-4 V and accuracy of cutting force, including shear angle, Strain rate coefficient of
extracted the cutting force coefficient of cemented carbide tools from the the first deformation zone and the second plastic deformation zone. In
orthogonal cutting data for predicting the micro milling force. Accord­ the course of calculations [7], if the actual parameters are out of the
ing to the result, the change in cutting force is consistent with that in the feasible range, it is difficult to find a precise shear angle using this
experimental value. Though finite element commercial software is method. In addition, the method is based on the assumption that the
capable of reducing the number of experiments and obtaining a sub­ normal stress distribution of the chip-tool interface is uniform and the
stantial amount of processing data with less manpower and material resultant force intersects the tool face at a distance from point B. which
resources required, it is computationally intensive and time consuming compromises the accuracy of shear angle prediction to some extent. If
in the context of complicated 3D models. Such relatively accurate the shear angle is known, the iteration speed of the Oxley model will
physical parameters as shear angle, deformation zone temperature, increase and the accuracy of prediction will be improved. Therefore, the
strain and strain rate can be obtained within a short space of time using a two-dimensional finite element simulation is proposed in this paper to
simplified 2-D finite element. These physical parameters play a critical obtain the shear angle. Besides, the Oxley model is introduced to obtain
role in the development and verification of the theoretical model, such the cutting force, and a hybrid FEM-Oxley model is developed to obtain
as the evolution of stress distribution [5] and grain size [6] involved in the orthogonal cutting force.
the shear zone, tool-chip friction zone and machined surface. The finite A large proportion of the turning tools used in the industry are
element observation of the shear angle to validate the shear angle designed with an edge radius, a nose radius and a complex cutting angle,
theoretical model has provided a feasible approach to machining for instance, rake angle, clearance angle, main cutting edge angle,
modeling. For example, Wang [7] applied the FEM to obtain the shear inclination angle, and so on. According to previous studies, complex
angle and the strain rate coefficient of the first deformation region for tools can reduce cutting force and cutting energy, improve surface fin­
the high-speed cutting of the Ti-6Al-4 V, before cutting force was ish, as well as achieve higher strength and wear resistance [17,18].
calculated. Zhu et al. [2] constructed an analytical model of cutting Sheikh-Ahmad et al. [19] proposed a method to predict the cutting force
force for ultrasonic vibration assisted cutting. Based on ABQUS, an of a tool with complex geometry. It combines mechanical modeling
analytical model of shear angle taking account of servo motion char­ method and neural network data. Allowing for the impact of edge
acteristics was applied to the cutting force analysis model, based on radius, Deng et al. [20] constructed the cutting force model of
which an excellent consistency was achieved between the simulation micro-machined particle-reinforced metal matrix composites (PMMCs),
and experiment of cutting force. which led to the conclusion that the particle-reinforced and edge radius
Compared with the 3-D finite element model with the capability to could have a significant impact on the micro-milling process of PMMCS.
simulate the oblique cutting force, the analysis model is usually char­ Considering the size effect, Wang et al. [21] suggested a turning
acterized by a fast calculation and clear mechanism. Merchant [8] was analytical force model for round inserts that discretized cutting edge.
the first to propose a single shear plane model. Oxley [9] suggested the Nevertheless, this is an extraordinary insert with no universality shown
prominent parallel plane shear zone model. There are various unknown in ordinary machining. Based on the model proposed by Molinari et al.
coefficients in Oxley’s predictive machining theory, but there is no ac­ [22], Fu et al. [23] decomposed the actual cutting edge of ordinary
curate experimental data to reflect the material deformation behavior turning tool into a number of infinitesimal cutting units, to study the
during the cutting process, which makes it difficult to generalize. global and local chip flow angles, predict the cutting force and analysis
Therefore, Adibi-sedeh [10] and Lalwani [11] extended Oxley’s pro­ the model for its effectiveness. Ordinary turning tools show complex
cessing prediction theory by replacing Oxley’s material model with geometries, and the cutting edges involved are usually comprised of the
other commonly used material models. Huang and Liang [12] con­ main cutting edge, the end cutting edge and the connected tool nose
structed a model to investigate the impact of tool thermal performance radius, which could affect the cutting force and improve the machined
on cutting force. The heat transfer simulation of the heat sources in the surface. S.V.A et al. [24] pointed out that the geometry of the tool has
first and second deformation zones was performed to correct the pre­ the most impact on the feed force. Therefore, it is essential to factor the
dictive machining theory proposed by Oxley. In order to generalize the impact of nose radius into the prediction of the cutting force.
modeling method, the performance of the workpiece material was In summary, at present, there are still few studies on predicting
expressed as a function of strain, strain rate and temperature, while the cutting force using hybrid method, especially considering the cutting
modified J-C equation was applied to the modified Oxley method, which force modeling of tool edge radius and nose radius. Therefore, this study
revealed that the modified J-C equation is better at reflecting the ma­ aims to develop a hybrid prediction model of 3-D oblique turning forces
terial deformation behavior of the cutting process. Pan et al. [13] pro­ for difficult-to-machine materials. As shown in Fig. 1, the cutting force
posed to consider the effect of grain growth and material softening in the as obtained by Deform-3D is compared against the experimental cutting
JC equation and incorporate it into the Oxley model for the iterative force, which provides a relatively accurate J-C material constitutive
calculation of cutting force. Compared with the traditional Oxley model, model for the subsequent 2-D finite element shear angle prediction and
the predicted cutting force is closer to the experimental cutting force. Oxley model. Then the shear angle regression model is constructed by
Not only are the standard Hopkinson pressure bar experiments restricted simulating orthogonal cutting based on Deform-2D. In combination with
to relatively low strain and strain rates, they also fail to reflect strain and the chip flow angle, the oblique cutting is converted into the equivalent
strain rates in the initial shear zone (PSZ). In addition, the premise of orthogonal cutting model. The cutting force is obtained iteratively in the
predictive processing theory for the approximation of shear strain and Oxley model, and the plowing force generated by the edge radius is
shear strain rate is known as the characteristics of PSZ thickness or chip predicted using Waldorf sliding line field model. Finally, coordinate
morphology, while these characteristics are often difficult to measure. transformation is performed to determine the cutting force of 3-D obli­
As a result, when the research on machining theory involves material que turning. A number of cutting tests on 304 stainless steel and Inconel
constitutive model, a majority of them continue to apply the classical J-C 718 are conducted and compared with the experimental results to
model [14], and more researchers have attempted improving the validate the model.
feasibility of other aspects to optimize the model. For example, Pang
et al. [15] applied the improved Oxley predictive machining theory to 2. Determination of material constitutive model based on FEM
conduct analysis of the cutting mechanics during the end milling. Li
et al. [16] suggested an unequal shear region model and achieved The material constitutive model is crucial to not only the subsequent
consistency with the experimental results. For the Oxley cutting theory, 2-D finite element shear angle observation but also the calculation of the

1305
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

Fig. 1. Flow chart of the proposed approach.

flow stress in the shear zone and the tool-chip friction zone. At the
Table 2
present time, finite element simulation technology has been applied in
Thermal and mechanical properties of 304 stainless steel.
various researches for the observation and verification of material pro­
cessing and its results. In this part, combined with the cutting force Density (kg/m3) 7930
experiment, the J-C material model with a higher accuracy is chosen to Thermal conductivity (w/m/K) 16
substitute the default flow stress model of Deform. Specific heat (J/kg/K) 500
6
When the irreversible cutting deformation process is simulated, J-C Thermal expansion Coefficient (l/K) 16×10−
Young’modulus (Gpa) 193
material model [25] has been widely applied in the simulation. The Poisson’s ratio 0.28
material constitutive model of J-C is defined as follows:
[ ][ ( )m ]
ε̇ T − Tw
σ = (A + Bεn ) 1 + Cln 1− (1) τindicates the shear stress,μdenotes the friction coefficient,
ε̇0 Tm − Tw
andτ0 represents the shear yield stress.
Where σ indicates the equivalent flow stress, while ε,ε̇ and ε̇0 Due to the low thermal conductivity of difficult materials, serrated
represent equivalent strain, equivalent strain rate and reference equiv­ chips will appear in the process of machining. Default Normalized
alent strain rate, respectively, Tdenotes instantaneous temperature of Cockroft and Latham [27] fracture criteria was adopted in this study, as
the material, Tw represents the ambient temperature, Tm refers to the shown in Eq. (3).
melting temperature. The model parameters of 304 stainless steel are

obtained by Zhou [26], as shown in Table 1. The material properties are εf
(σ ∗ /σ )dε = C (3)
indicated in Table 2. 0
Friction has impact on both cutting force and cutting temperature. In
this paper, the simplified shear friction conditions are used in the finite Whereσ∗ indicates the maximum main tensile stress, εf represents the
element software Deform, as shown in Eq. (2). fracture strain, and C denotes the material constant. In order to avoid the
separation criterion and prevent severe mesh deformation near the edge
τ = μτ0 (2)
radius, the Lagrange mesh repartition technique is applied [28].
During the process of cutting simulation, as shown in Fig. 1, the tool
Table 1
is modeled as a rigid part, while the workpiece is regarded as an elas­
material constants for 304 stainless steel [26]. toplastic part. The tool is restricted to moving along a straight line in the
feed direction, while the workpiece rotates around the axis x, Table 3
A B C n m ε̇0 Tm
shows the tool geometry and working angle. When the cutting force is
452Mpa 694Mpa 0.0067 0.311 0.996 0.001 1673K stabilized, the cutting force corresponding to different friction

1306
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

Table 3
Tool geometry and working angle.
tool types Main cutting edge angle Kr(◦ ) Rake angle α(◦ ) Clearance angle γ(◦ ) Inclination angle λs(◦ ) edge radius re(μm) nose radius r(mm)

CNMG120404 95 8 7 − 5 30 0.41

coefficients under the same cutting conditions can be obtained. As


Table 4
shown in Fig.2 (b), when the friction coefficient reaches 0.6, the pre­
The force of finite element simulation using different material models.
dicted value is close to the experimental value. Therefore, friction co­
efficient 0.6 is considered as more effective in predicting cutting force in ap = 1 mm ap = 1 mm ap = 1 mm
V = 90 m/min V = 90 m/min V = 150 m/
Deform-3D. As shown in Fig. 2 (a), the maximum temperature is 618 ◦ C Error
Material model min
for the workpiece. At this time, the maximum temperature of the tool f = 0.05 mm/ vf = 0.1 mm/ f = 0.1 mm/
(%)
reached concentrated in the area close to the edge, that is, 396 ◦ C. rev re rev
When the developed analytical model has yet to be verified, finite Exp 167 272 236 —
element provides a robust simulation platform for testing the J-C model Zhou [1,26] 150 264 231 5.1
for its rationality. Therefore, our choice for a reasonable 304 stainless Paolo [29] 166 293 281 8.7
steel J-C model is to draw comparison between the existing models Mohammad 144 234 238 9
[30]
(Zhou [1,26], Paolo [29] and Mohammad [30]), as shown in Table 4.
When the default friction coefficient (0.6) of the Deform is set, the
predicted value as obtained by the Zhou model is relatively closer to the
observed value.
Three sets of cutting forces predicted using the optimum J-C model
and the Deform default flow stress model are shown in Fig. 3. Appar­
ently, the deviation between the cutting force predicted by JC material
model and the experimental cutting force is shown to be minimal. The
cutting process is accompanied by the change to chip shape, cutting
force and tool-chip contact temperature, which conforms to dynamic
mechanical behavior. By analyzing and fitting the dynamic mechanical
properties of 304 stainless steel, Zhou [26] obtained the material pa­
rameters, as a result of which it is more consistent with the strain
behavior in the cutting process of stainless steel. The J-C material model Fig. 3. Cutting force predicted by different material models and observed force.
is incorporated into the two-dimensional finite element model and the
Oxley model. force model [9]. Considering of the impact of material flow stress with
strain, strain rate and temperature, the J-C [25] flow stress model is
3. Modeling of cutting forces applied. The coordinate system description diagram is illustrated in
Fig. 4. The calculation of the plowing force is performed using the slip
In this section of cutting force modeling, the cutting force can be line field model suggested by Waldorf [31], as shown in Fig. 5.
considered as being comprised of the chip forming force and the plowing
force. The calculation of chip formation force is based on Oxley’s cutting

Fig. 2. (a) Temperature field of tool and workpiece; (b) Predictive value of cutting force.

1307
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

N = Rcosλ (8)

Fs KAB t1 w
R= = (9)
cosθ0 sinϕcosθ0

Whereλindicates friction angle, whilet1 andθ0 represent undeformed chip


thickness and cutting width, respectively.θ0 denotes the angle between
the resultant force R and the shear surface AB.Fc andFt refer to the com­
ponents of chip formation force in cutting and vertical direction,Fin­
dicates the friction force of tool-chip interface, N denotes the normal
cutting force of tool-chip interface, and R stands for the resultant force of
normal force and friction force of tool chip interface.
Fig. 4. Oxley’s shear zone model [9]. According to the Von Mises stress yield criterion, once the shear
angle is determined, the strainεAB and strain rateε̇AB in the shear zone (as
shown in Fig. 4) can be calculated using the following formula:
1 cosα
εAB = √̅̅̅ (10)
2 3 sinϕcos(ϕ − α)

1 Vcosα
ε̇AB = √̅̅̅ (11)
3 Δs1 cos(ϕ − α)

WhereΔs1 indicates the thickness of the first deformation zone. Through


the cutting experiment on steel, it was discovered by Shaw [33]
thatΔs1 is about 0.025, which reduces the complexity of the model, as
recognized by many scholars. Subsequently, Grzesik [34] demonstrated
that the thickness valueΔs1 changes with the thickness of the cut and the
shear angle, and the relationship is:Δs1 = t1 /(10sinϕ). Considering of
the impact of cutting parameters on the machining process, the formula
Fig. 5. Schematic diagram illustrating the model of slip-line field for plow­ of Grzesik is used in this paper.
ing [31]. According to the velocity vector diagram shown in Fig. 4, the chip
velocityVc and the flow velocityVs of the shear surface can be calculated
3.1. Shear surface analysis as follows:
sinϕ
The Oxley model input consists of two elements, which are cutting Vc = V (12)
cos(ϕ − α)
parameters (cutting speed, depth of cut and tool rake angle) and
workpiece properties (heat capacity, thermal conductivity, and consti­ cosα
tutive model). In the process of calculation, the core algorithm of the Vs = V (13)
cos(ϕ − α)
Oxley model iteratively calculates other variables by applying different
shear angle valuesϕ, the Strain rate coefficient in the first deformation The deformation conditions in the cutting process indicate such
zoneCOxley and the strain rate coefficient in the second plastic deforma­ characteristics as high strain rate, large strain and rapid heating up.
tion zone δ. According to Oxley’s cutting theory, the calculation will be Therefore, the flow stress of shear plane AB is usually expressed by the
terminated when the output satisfies three equilibrium conditions [32]: constitutive model of Eq. (1). In the presence of plane strain, the flow
first, stress balance at the tool-chip interface, where the tool-chip stress of shear plane is expressed as
interface tangential stress (τint ) is equal to the strength of the chip root /√̅̅̅
KAB = σ 3 (14)
material (Kchip ); second, stress balance at the tool edge, that is, the
normal stress at the tool-chip interface calculated by resultant force (σ N ) After the flow stress of the shear plane is obtained, the remaining
is equal to the normal stress calculated using stress boundary condition cutting components are calculated by applying Eq. (4)–(9), the friction
at tool edge (σ N ); and third, the principle of minimum cutting force. angle isλ = θ0 + α − ϕ, the friction coefficient isμ = tanλ, andθ0 in Fig. 4

Firstly, an analysis is conducted of the stress distribution for shear satisfies the following relationship:
plane AB. As shown in Fig. 4, the region between boundary CD and EF is (π )
the chip formation region. Oxley proposed a parallel shear theory for tanθ0 = 1 + 2 − ϕ − Cn (15)
4
this region, and predicted chip formation forces based on the assumption
that CD and EF boundaries are parallel and equidistant from AB. Once The calculation method devised by Wang [7] for modifiedCn is shown
the shear angle is determined, the cutting force can be obtained by the in formula (16), and the influence of material strain is taken into account
equilibrium equation of the shear plane and the tool-chip interface. In and combined with Johnson-Cook flow stress model,
this case, the chip thickness and other cutting force components can be BεnAB
determined by applying equation [8] as follows: Cn = COxley n (16)
A + BεnAB
t2 = t1 cos(ϕ − α)/sinϕ (4) COxley represents the Strain rate coefficient in the first deformation
zone, and is dependent on the machining shape and the workpiece
Fc = Rcos(λ − α) (5)
material, with a typical value ranging from 2 to 10 [35].
The average temperature of the shear plane AB [35] can be calcu­
Ft = Rsin(λ − α) (6)
lated by:
F = Rsinλ (7) (1 − β)Fs Vs
TAB = Tw + η (17)
ρwk Vt1 wCp

1308
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

Whereηindicates the percentage of total shear energy converted to the tool edge is completely sharp, and the tool edge has a certain edge
enthalpy, and 0.9 [35] is taken in this analysis. The total force applied radius in the actual machining process. As for ordinary cutting, the in­
along the shear plane AB can be calculated asFs = KAB LAB wCp refers to fluence exerted by the fillet action of the edge radius can be discounted.
the specific heat of the workpiece material, andβdenotes the heat dis­ Nevertheless, in respect of precision cutting, when the depth of cut has
tribution coefficient of the shear region. the same order of magnitude as the edge radius, the influence of the edge
radius must be taken into consideration. The force applied by the contact
β = 0.5 − 0.35log10 (Rt tanϕ)for0.04 ≤ Rt tanϕ ≤ 10
(18) between the edge radius and the machined surface is known as the
β = 0.3 − 0.15log10 (Rt tanϕ)forRt tanϕ > 10
plowing force.
WhereRt represents a dimensionless thermal coefficient calculated by As shown in Fig. 5, the slip line field model proposed by Waldorf [31]
Eq. (19). is involved to make prediction of the plowing force due to the edge
radius. The anglesηplow θfan andγplow of the sector are determined by the
ρwk Cp Vt1
Rt = (19) geometric and frictional relationships described by the following
Kwk
formula.
Whereρwk refers to the density of the workpiece material,
Kwk represents the heat transfer coefficient of the workpiece material. Rfan
CB = (27)
The normal stress of point B indicated in Fig. 4 can be obtained by sinηplow
combining the boundary stress of point B and the flow stress model of
Johnson-Cook. ηplow = 0.5⋅cos− 1 (μplow ) (28)
( π )
σ ’N = KAB 1 + − 2α − 2Cn (20) θfan =
π
− ρprow − ϕ (29)
2 4
√̅̅̅
3.2. Tool-chip interface analysis γ plow = ηplow + ϕ − sin− 1 ( 2sin(ρprow )sin(ηplow )) (30)

The sector radius of the sector region indicated by Waldorf [31] is


The plastic region of the tool-chip interface can be assumed to be a
obtained.
rectangular plastic region of constant thicknessΔS2 = δt2 . Therefore, the
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
effective plastic strain rate at the tool-chip interface can be obtained by: √⎛ ⎞2
√ √̅̅̅
√ (π α) 2 R sin( ρ ) [( ) ]2
⎜ ) ⎟
Rfan = sinηplow ⋅√
1 Vc fan prow
ε̇int = √̅̅̅ (21) √⎝re ⋅tan +
4 2
+ ( ⎠ + 2 Rfan ⋅sin(ρprow )
3 δt2 tan π4 + α2
Wheret2 indicates the tool-chip contact length, calculated by Eq. (4). (31)
δdenotes the Strain rate coefficient in the second plastic deformation
re is the tool edge radius,ρprow indicates the angle between the un­
zone [13]. The tool-chip contact lengthhis determined by the moment
balance formula of the shear plane: processed convex portion and the horizontal plane, which is set to 10◦
( ) [37], andμplow denotes the friction factor, which is assumed to be 0.8 [1]
h=
t2 sinθ0
1+
Cn
(22) in this analysis.
cosλsinϕ 3tanθ0 The plough force is capable to be decomposed into the forcepc along
It is assumed that the tool-chip interface stress conforms to uniform the cutting direction and the forcept [31] perpendicular to the machined
distribution so as to obtain stressesτint andσ N of the tool-chip interface and surface. If the shear angleϕand the flow stressKAB of the shear surface are
the B point, which can be calculated by the formula: τint = F/hwandσ N = known, the plough force can be obtained by the formula below.
[ ]
N/hw. The average temperature of the tool-chip interface [36] is cos(2ηplow )cos(ϕ − γplow + ηplow )+
expressed as: Pc = KAB ⋅w⋅CB
(1 + 2θfan + 2γplow + sin(2ηplow ))sin(ϕ − γ plow + ηplow )
(1 − β)Fs Vs (32)
Tint = Tw + + Ψ ΔTM (23)
ρwk Vt1 wCp [ ]
(1 + 2θfan + 2γplow + sin(2ηplow ))cos(ϕ − γplow + ηplow )
Pt = KAB ⋅w⋅CB
WhereΨ indicates the correction factor, taken as 0.6 [35], − cos(2ηplow )cos(ϕ − γ plow + ηplow )
ΔTM denotes the maximum temperature rise in the chip at the tool-chip (33)
interface, andΔTC represents the average temperature rise in the
chip.ΔTM andΔTC bear the relationship as follows. In summary, the total cutting force refers to the superposition of the
[ chip forming force and the plowing force. In the orthogonal cutting, the
( )]
√̅̅̅̅̅
R t
̅ R t
cutting force in the cutting direction and the cutting force perpendicular
to the direction of the machined surface can be expressed as:
0.06− 0.195(δ) T 2 +0.5log10 T 2
h h

ΔTM = ΔTC 10 (24)


FC = Fc + Pc (34)
FVc
ΔTC = (25)
ρwk Vt1 wCp FT = F t + P t (35)

After the tool-chip interface temperature is obtained, Eq. (1) is


√̅̅̅ 4. 3D turning force model
applied to determine the average flow stress in the chipKint = σ / 3,
wherein the effective strain of the chip can be calculated by the formula 4.1. Prediction of chip flow angle and equivalent cutting edge geometry
(26):
1 h During the process of oblique cutting, the three-dimensional plastic
εint = √̅̅̅ (26) flow of the chip formation is significantly more complicated than the
2 3 δt2
orthogonal machining process. In order to utilize the previous orthog­
onal cutting force model, there is a necessity to model the effect of the
3.3. Plowing force
nose radius during the cutting process. In previous studies, equivalent
straight edges were used to predict the influence of nose radius [38,39].
In the above chip forming force prediction model, it is assumed that

1309
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

On the basis of reasonable assumptions, Young et al. [38] constructed a calculated by applying the following formula:
relatively simple model, where the orthogonal machining condition(i = ⎧ ⎫
0α = 0) was applied to predict the flow direction of chips during bar ⎨ 1 + sini⋅tanCs ⋅tanα ⎬

Cs = cos − 1
(40)
turning. Subsequently, this model was extended to the circumstance [
⎩ (tanα⋅tanC + sini) + cos2 i⋅sec2 C
2 ]1/2 ⎭
s s
where the inclination angle and rake angle were not equal to zero [40],
and it was determined that the proposed method was suitable for obli­ ( )
ir θ = − sin− 1 [sinα⋅cosi⋅sin(θ + C’ s ) + sini⋅cos(θ + C’ s )] (41)
que cutting model. For the calculation of the chip flow angle, as shown
in Fig. 6, the undeformed chip contact area is split into three regions, The area of the region S can be determined by the relevant cutting
regions A and B respectively correspond to the nose radius portion, and geometry:
region S corresponds to the side edge straight portion.
The chip flow angle due to the impact of the tool nose radius is AS = f [d’ − r⋅(1 − sinC’ s ) − 1/4⋅f ⋅sin(2C’ s )] (42)
calculated based on geometric considerations and experimental obser­ d represents the projection of the depth of cutdin the plane, which

vations. It is assumed that the direction of the resultant force represents can be obtained from the equation
the direction of the chip flow, and the micro-component friction is in­
[ ]1/2
tegrated along the cutting edge to obtain the resultant friction. Fig. 6 d’ = d⋅cosCs ⋅seci⋅ (tanα⋅tanCs + sini)2 + cos2 i⋅sec2 Cs (43)
presents a schematic diagram of the nose radius smaller than the cutting
depth. At this time, only the tool side edge and the tool nose radius are The upper and lower limits of the integral are as follows.
involved in the cutting. The predicted chip flow direction can be f
θ1 = cos− 1 ( ) (44)
generally expressed as: 2r
(∑ ) ( )
NUMi
Ω = tan− 1 ∑ i = A, B, S (36) θ2 = π − tan− 1 r⋅sinCs’
(45)
DENi r⋅cosCs’ − f
Where
/ ∫ { ⎫ θ3 = π − Cs’ (46)
θ2 [ ]2 }


NUMA = − 1 2 cos(θ − ir ) r2 − f ⋅cosθ + (r2 − f 2 ⋅sin2 θ)1/2 dθ ⎪
⎬ The flow direction of chip can be predicted by using the following
equation.
θ1
/ ∫ θ2 { [ ]2 } ⎪

DENA = 1 2 sin(θ − ir ) r2 − f ⋅cosθ + (r2 − f 2 ⋅sin2 θ)1/2 dθ ⎪
⎭ π
θ1 ηc = − C ’ s − Ω (47)
2
(37)
Finally, the equivalent cutting edge of the known cutting geometry is
/ ∫ θ3 { [ ]2 } ⎫ determined,
NUMB = − 1 2 cos(θ − ir ) r2 − f ⋅cosCs’ − r sec2(θ + Cs’ ) dθ ⎪


/ ∫ θ3
θ2
{ } Cs∗ = C’ s + ηc (48)
[ ]2 ⎪

sin(θ − ir ) r2 − f ⋅cosCs’ − r sec2(θ + Cs’ ) dθ ⎭
DENB = 1 2
θ2
In accordance with the ISO standard of single point cutting tool, the
(38) maximum feed is limited to being less than 0.8 times of the nose radius,
suggesting that the straight line part of the end cutting edge is not
}
NUMS = cos(Cs’ + i)⋅AS involved in the cutting. In the actual range of cutting conditions, it is
(39) rare to observe that only the nose radius is used for machining. There­
DENS = sin(Cs’ + i)⋅AS
fore, the cutting geometry referred to above rules out the fact that cut­
Wheredindicates the cutting depth,frefers to the feed,ir denotes the ting occurs at the straight line of the cutting edge at the end and the nose
angle of the edge inclination that varies withθ, andC’s represents the radius is exclusively involved in the cutting depth direction.
projection of the side edge angleCs on the rake face plane, which can be

4.2. Coordinate transformation

In respect of orthogonal theory, the undeformed chip thicknesst1 is


equal to feedf, and the cutting widthwis equal to the cutting depthd.
With regard to oblique cutting, there is a need for the application of
orthogonal theory to be transformed into equivalent value. Firstly, the
equivalent inclination anglei∗ , the equivalent rake angleα∗ and the
equivalent chip flow directionη∗c are expressed as follows:

i∗ = sin-1 (cosηc ⋅sini − sinηc ⋅sinα⋅cosi) (49)


( )
secηc sini − sini∗
α∗ = sin− 1
(50)
tanηc cosi∗

η∗c = i∗ (51)

As shown in Fig.7, the feedf, the cutting depthdand the tool nose
radiusrare known, and the equivalent cutting edgeC∗S is obtained using
the predicted chip flow direction. Then, the equivalent cutting thick­
nesst1∗ and the cutting widthw∗ are obtained by applying the following
formula, respectively:
t1∗ = f cosCs∗ (52)

Fig. 6. Geometric relations of the chip flow model for nose radius tools.

1310
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

maximum effective strain rate along the shear band. Besides, the value
of shear angle can be obtained.
As shown in Fig.8, an obvious strain rate distribution can be observed
in the shear band region, and the maximum strain rate during the cutting
process is distributed in the radius of the tool edge. The shear angle
value can be measured with the assistance of computer aided design
software. Based on the obtained shear angle observation, a full quadratic
model describing the shear angle and cutting speed and feed is con­
structed. It is assumed that the shear angle equation is

ϕ = b0 + b1 X1 + b2 X2 + b3 (X1 )2 + b4 (X2 )2 + b5 X1 X2 (60)


X1 X2 represent cutting speedVand cutting thicknesst1 , respectively,
Fig. 7. Orientation for 3-D oblique cutting geometry [37].
and the full quadratic model is expressed as follows:
d ϕ = 19.8 + 0.02 × V + 37.5 × t1 − 3.2 × 10− 5 × V 2 − 56 × t12 − 0.0019 × V
w∗ = (53)
cosCs∗ × t1
When the Oxley model is used to predict cutting forcesFc andFt , the (61)
orthogonal cutting thickness and cutting width are replaced with­
A non-linear relationship between the shear angle and the cutting
t1∗ andw∗ . Meanwhile, the equivalent cutting width, cutting thickness and
parameters is determined, as shown in Fig. 9. The results demonstrate
cutting geometry are required to be factored into the previous plowing
thatϕis on the rise as cutting thickness or cutting speed increases, and
force formula for predicting the plowing forcesPc andPt .The cutting
the most influential one is cutting thickness, followed by cutting speed.
forcesFC ,FT andFR shown in Fig. 7 are obtained by the following formula.
The model is employed to input the shear angle of the modified Oxley
FC = F c + P c (54) model.

FT = F t + P t (55) 5.2. Solution analysis of Oxley model


( )
FC sini∗ − cosi∗ sinα∗ tanηc − FT cosα∗ tanη∗c As the shear angleϕis known, it is considered as necessary to itera­

FR = (56)
sini∗ sinα∗ tanη∗c + cosi∗ tively solve the strain rate coefficientCOxley of the first plastic deforma­
tion zone and the strain rate coefficientδof the second plastic
In order to represent the force in the direction measured by the
deformation zone. In this paper, the solution process is described by
dynamometer, there is a necessity to perform coordinate transformation
simplifying the parameter (w = 1, t1 = 0.15). Fig. 10 shows the process
of cutting forcesFC ,FT andFR . If the cutting, feed and axial forces are set to
in which the minimum stress difference varies with the coefficient of the
beP1 P2 andP3 , then the following formula is calculated as follows.
second deformation zone at different cutting speeds. The response value
P1 = FC (57) decreases at first and then increases, which indicates that the coefficient
corresponding to the minimum value in the middle is the coefficient of
P2 = FT cosCs∗ + FR sinCs∗ (58) the second deformation zone, and it will decline with the increase of
cutting speed, the minimum stress difference is calculated as follows.
P3 = FT sinCs∗ − FR cosCs∗ (59)
EτK = τint − Kchip (62)
5. Result and discussion As shown in Fig. 10, the minimum difference between the tool-chip
interface tangential stress and the strength of the chip root material is
5.1. Determination of shear angle by FEM changed primarily by the increase inδ, which is found according to the
modified Oxley model. Because the shear angleϕis constant and the­
In order to simplify the process of solving the shear angle, which COxley obtained by the solution remains unchanged. The chip forming
means the impact of the cutting depth on the shear angle is ignored, the forceFc in the final output cutting speed direction is kept constant as well.
three-dimensional cutting process is simplified to two-dimensional In addition, Adibi-Sedeh [10] found out that the cutting force is insen­
orthogonal cutting and the thermodynamic coupling finite element sitive toδafter the Johnson-Cook equation is applied. Therefore, δcan be
simulation is conducted. It is assumed that the workpiece is rectangular determined using the minimumEτK value that satisfies the stress equi­
and moves in the opposite direction to the cutting speed. Similar to the librium at the tool-chip interface. By contrast, in the conventional Oxley
previous three-dimensional simulation, the radius of the tool edge is model, δis determined in accordance with the principle of the minimum
30μm, and the material constitutive equation parameters, friction model cutting force. Therefore, a new Oxley model solving process is
and fracture criterion are identical to the previous three-dimensional established.
model. Fig. 11 (a) illustrates a process where the average temperature of
In order to determine the shear angle values over a large range of shear zone shows a tendency of iterative stabilization, before reaching
cutting parameters, a combination of 2 factors and 3 levels of cutting up to243∘ C. Fig. 11 (b) presents the temperature distribution of the shear
parameters was performed, as shown in Table 5. In the finite element zone, friction zone and plowing zone obtained by performing finite
simulation of different cutting thickness and cutting speed, the position element simulation. The temperature of the shear zone ranges
of shear surface AB can be determined by taking measurement of the between104∘ ∼ 358∘ , and the average value is calculated as approx­
imately231∘ . The average temperature of the shear zone is predicted to
be close to the finite element simulation results.
Table 5
Variable test parameters.
5.3. Experimental comparisons
parameters Low level Medium level High level

V(m/min) 50 150 300 5.3.1. Orthogonal cutting force model verification


t1(mm) 0.05 0.15 0.3
The 3D oblique cutting force model proposed in this paper is

1311
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

Fig. 8. Observation of shear angle from strain rate distribution.

developed on the basis of the effective orthogonal cutting force model,


as a result of which the correctness of the improved Oxely model needs
to be validated. For the above analysis, the force prediction was made
under different orthogonal cutting conditions. The cutting force data of
the orthogonal cutting Ti-6Al-4 V material [37] was applied to validate
the model. The cutting speed was set to 30 m/min, the rake angle to8∘ ,
the radius of the tool edge to13μm, the cutting width to 3.8 mm, and the
undeformed chip thickness to the range from0.025mmto 0.1524 mm.
The required Ti-6Al-4 V material properties and J-C equation parame­
ters are indicated in Tables 6 and 7.
Firstly, the shear angle value of the corresponding cutting parame­
ters is obtained by applying FEM. It can be seen from Fig. 12 (a) that the
Fig. 9. (a) Surface plot of shear angle vs cutting speed and feed. shear angle is on the rise as the cutting thickness increases, showing a
similar trend of change to the previous 304 stainless steel. The shear
angle is factored into the modified Oxely model to solve the chip forming
forcesFc andFt , as well as the plowing forcesPc andPt produced by the
cutting edge. The prediction results are indicated in Fig. 12 (a), which
shows that the error between the cutting force predicted by the
improved Oxely model and the observed cutting force is insignificant.
Meanwhile, in order to further validate the correctness of the cutting
force model in orthogonal cutting 304 stainless steel, the cutting force
data of the orthogonal cutting 304 stainless steel by Zhang [1] is refer­
enced. The rake angle of the tool is7∘ and the edge radius is40μm. The
cutting parameters are listed in Table 8, and the predicted results are

Table 6
Thermal and mechanical properties of Ti-6Al-4 V [13].
Density (kg/m3) 4420
Fig. 10. The minimum stress difference at different cutting speeds varies with
Thermal conductivity (w/m/K) 6.7
the coefficient of the second deformation zone (w = 1, t1 = 0.15).
Specific heat (J/kg/K) 560
6
Thermal expansion Coefficient (l/K) 9.2×10−
Young’modulus (GPa) 114
Poisson’s ratio 0.33

Fig. 11. (a)The iterative solution process of temperature in shear zone AB and (b) temperature field of FEM (w = 1, V = 150, t1 = 0.15).

1312
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

Table 7 hard-to-machine materials in the nuclear and aerospace industries.


Johnson-Cook constants for Ti-6Al-4 V [41]. Fig. 13 (a) shows the lathe used in this experiment, model C6136HK
A B C n m ε̇0 Tm CNC lathe. The types of tool used for cutting 304 stainless steel and
Inconel 718 materials are CNMG120404 and CNMG120408, respec­
862Mpa 331Mpa 0.012 0.34 0.8 1 1950 K
tively, as shown in Fig. 13(b) and (c). The tool geometry is indicated in
Tables 3 and 9.
During the cutting process, as illustrated in Fig. 13 (d), the KISTLER
piezoelectric dynamometer 9257B is fitted on the machine tool holder to
measure the cutting force in three directions during the turning process
(Main cutting forceFx , Feed forceFy , Radial forceFz ). In order to reduce
the error to the minimum, each set of experiments was repeated three
times and then the average value was calculated. For validation of the
model and to analysis of the influence exerted by the cutting parameters
on the cutting force, a single-factor experiment is designed. The exper­
imental layout is indicated in Table 10. The work pieces are 304 stainless
steel bar and Inconel 718 bar with a diameter of 69 mm.
Figs. 14–16 present the predicted value and experimental value of
cutting force of 304 stainless steel. After calculation, the average errors
ofFx ,Fy andFz in the nine experiments are found to be 7%, 11 % and 11 %,
respectively. As revealed by the comparative relationship, the predicted
results of the model conform well to the experimental results.
The component forces Fx ,Fy andFz in the three directions vary with the
cutting parameters. That is to say, as the feed rises, Fx exhibits a trend of
linear increase, andFy Fz exhibit a trend of non-linear increase. This is
attributed to the increase of feed that exacerbates cutting deformation
and tool-chip friction [42], thus causing the cutting force to increase.
When the depth of cut is on the rise, Fx andFy exhibit a rapid increase in
nonlinearity, andFz exhibits an increase of non-linearity at a slow pace.
The contact area between the cutting tool and the chip expands as the
cutting depth is on the rise. Due to the increased contact area, the ma­
terial removal rate improves, and the pressure of the tool edge increases
[43], as a result of which the cutting force shows an upward trend.
Under the specified experimental conditions, the cutting speed increases
and the temperature near the tool-chip interface rises significantly. The
high temperature produces a thermal softening effect on the shear flow
stress, which could impact on the sliding friction coefficient, thus
resulting in a decline in the frictional stress of the secondary shear zone.
Fig. 12. Trends of cutting force and shear Angle with different cutting pa­ Therefore, the cutting speed has impact on the plastic flow stress during
rameters. (a) Ti-6Al-4 V, (b) 304 stainless steel. the cutting process, thus causing the predicted cutting force to decline as
the cutting speed is on the rise. As indicated by the experimentally ob­
tained force, when the cutting speed is on the increase, the two relatively
Table 8 large forcesFx andFy exhibit a rapid decline in nonlinearity, which con­
Cutting conditions for 304 stainless steel.
forms to the theoretical analysis. The observed value has no obvious
Case V(m/min) t1 (mm) w(mm) change, while the prediction shows a downward trend.
1 35 0.1 3 Within the range of selected parameters, when cutting 304, the
2 60 0.1 3 maximum influence on the main cutting force is exerted by feed, fol­
3 60 0.2 3 lowed by the depth of cut and the cutting speed. The most significant
4 85 0.1 3
5 85 0.2 2.8
influence on the feed force is the depth of cut, followed by the cutting
6 110 0.1 3 speed and feed rate. The feed rate has the most significant influence on
7 110 0.2 2.5 the radial force, followed by the depth of cut and the cutting speed.
The cutting force of the cutting Inconel 718 in Table 10 was pre­
dicted by taking the same cutting force prediction method as shown in
indicated in Fig. 12 (b). The prediction errors of cutting force in cutting
Fig. 1. The J-C model parameters of the model input are obtained from
direction and perpendicular direction are 6% and 7%, respectively.
the literature [44], as listed in Table 11. The tool geometry angle is
According to the results, the predicted cutting forces are broadly
indicated in Table 9, and the material properties are shown in Table 12.
consistent with the measured results. Therefore, the orthogonal cutting
The comparison made between the simulation results and the experi­
force prediction model can be taken as the basis for the prediction of 3-D
mental data is illustrated in Figs. 17–19. The average errors ofFx ,
cutting force.
Fy andFz in the seven groups were 6%, 8% and 4%, respectively.
The variation of force in three directions with cutting parameters is
5.3.2. 3-D turning force model verification
as follows. When turning Inconel 718, the feed rate and depth of cut
Based on the above analysis, a computer program was developed in
have the same effect on the cutting force, and the cutting force increases
MATLAB 2017a to construct a hybrid model, prior to the prediction of
with the feed rate and depth of cut. It can be seen from the experi­
the cutting force. A range of cutting experiments were conducted on
mentally obtained force that as the depth of cut increases,Fx andFy show a
various difficult-to-machined materials, and the applicability of the
trend of linear increase, whileFz declines at a slow pace. Besides, the
hybrid model proposed in this paper was verified. The two materials are
trend of change forFz is contrary to the predicted result. As the feed rate
304 stainless steel and Inconel 718, both of which are known as typical

1313
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

Fig. 13. Cutting test platform: (a) Turning test; (b) A turning tool of model CNMG120404; (c) A turning tool of model CNMG120408; (d) KISTLER piezoelectric
dynamometer.

Table 9
Tool geometry and working angles.
tool types Main cutting edge angle Kr(◦ ) Rake angleα(◦ ) Clearance angle γ(◦ ) Inclination angle λs(◦ ) edge radius re(μm) nose radius r(mm)

CNMG120408 93 8 5 − 6 20 1.30

Table 10
Machining parameters for 3D oblique turning.
Materials Test Number ap (mm) V(m/min) f(mm/rev)

1 1 90 0.05
2 1 90 0.1
3 1 90 0.15
4 1 150 0.1
304 stainless steel 5 1.375 150 0.1
6 1.75 150 0.1
7 1.75 90 0.15
8 1.75 150 0.15
9 1.75 210 0.15
10 1 40 0.1
11 1.5 40 0.1 Fig. 14. Main cutting force predictions for Case 1-9.
12 2 40 0.1
13 1.5 40 0.1
Inconel 718 14 1.5 40 0.15
15 1.5 40 0.2
16 1.5 40 0.1
17 1.5 70 0.1
18 1.5 85 0.1

is on the rise,Fx appears to increase linearly, whileFy andFy exhibit a slow


increase in nonlinearity. As the cutting speed increases,Fx Fy andFz show a
rapid decline in nonlinearity. As revealed by test16-test18 shown in
Fig. 17, though the predicted cutting force in the direction of cutting
speed is consistent with the experimental numerical trend, the reduction
in cutting force obtained from the experiment is less than obvious. As
Fig. 15. Feed force predictions for Case 1-9.
shown in Fig. 15, this can also be observed in the trend of change in the
feed force of 304 stainless steel. Meanwhile, the predicted result exhibits
a slow-paced decline in nonlinearity, which is due to that, in actual thermal softening of the material during the cutting process is found to
processing, the cutting speed increases, the temperature rises, and the be more severe than in the standard compression test, and the J-C ma­
material is subject to impact by high temperature and high strain. The terial model parameters are measured in standard compression tests,

1314
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

Fig. 19. Radial force predictions for Case 10-18.


Fig. 16. Radial force predictions for Case 1-9.

some potential defects under high strain [14].


In the range of selected parameters, when cutting Inconel 718, the
Table 11
Johnson-Cook constants for Inconel 718 [44]. maximum impact on the main cutting force is caused by the depth of cut,
followed by the feed rate and cutting speed. In addition, the cutting
A B C n m ε̇0 Tm
depth is the most important influencing factor in the feed resistance,
1200Mpa 1284Mpa 0.006 0.54 1.2 0.001 1800 K followed by the cutting speed and feed rate. Feed rate exerts the most
significant influence on cutting depth resistance, followed by cutting
speed and cutting depth.
Table 12
Thermal and mechanical properties of Inconel 718. 5.3.3. The effect of cutting parameters(f, ap , V)
The impact of such cutting parameters feed, cutting depth and cut­
Density (kg/m3 ) 8200
ting speed on cutting force is predicted using hybrid model. When the
Thermal conductivity (w/m/K) 11.4 cutting material is chosen as 304 stainless steel and the type of tool is set
Specific heat (J/kg/K) 4350 to CNMG120404, the three-dimensional curvature is obtained for the
Thermal expansion Coefficient (l/K) 11.8×10− 6
effect of cutting parameters on the Fx, Fy and Fz shown in Fig. 20. It can
Young’ modulus (GPa) 185 be seen from this figure that the component of cutting force shows a
Poisson’s ratio 0.33 strong clearly mapping relationship with the change in cutting param­
eters. In general, the component of cutting force shows an increasing
trend with the increase of feed and cutting depth, but a downward trend
with the rise of cutting speed.

5.3.4. The effect of tool geometry (re , r)


A hybrid model is applied to predict the impact of tool geometry
(edge radius and nose radius) on cutting force. When the cutting ma­
terial is chosen as 304 stainless steel and the tool model is selected as
CNMG120404, edge radius and nose radius range from 0.01 mm to 0.06
mm and from 0.4 mm to 1.2 mm, respectively. Then, the three-
dimensional surface diagram is obtained for the impact on the compo­
nents of cutting force, including Fx, Fy and Fz as shown in Fig. 21. It can
be seen from the figure that the components of cutting force increase
linearly as the radius of the edge, which is because the larger the radius
Fig. 17. Main cutting force predictions for Case 10-18.
of the edge, the higher the proportion of the edge involved in the
operation, which makes cutting deformation and friction more signifi­
cant, As a result, cutting force is increased. As the radius of the nose
increase, only the main cutting force and depth resistance show an up­
ward trend, while the feed resistance exhibits a downward trend, despite
to an insignificant extent, which is consistent with the trend indicated by
Fu et al. [23].
In summary, by validating and analyzing of the cutting force of
304stainless steel and Inconel 718, the results are obtained to demon­
strate that the predicted value is excellently consistent with the exper­
imental value, which evidences the effectiveness of the cutting force
modeling method proposed in this paper. Through the above analysis, it
can be concluded that the cutting force can be reduced by the combi­
nation of cutting parameters with low feed rate, low cutting depth and
Fig. 18. Feed force predictions for Case 10-18.
high cutting speed, or the structure of small edge radius and nose radius.

resulting in a predicted cutting force that is higher than the actual cut­ 6. Conclusions
ting force. As indicated by Sun [45], the flow stress obtained under high
strain at cutting is significantly lower than the flow stress measured in In this paper, a hybrid FEM-Oxley model is constructed by combining
the compression test. Therefore, the J-C material model measured ac­ finite element simulation and Oxley model. In consideration of the in­
cording to the Hopkinson pressure bar experiment continues to show fluence exerted by the edge radius and the nose radius of cutting tool, a

1315
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

Fig. 20. The impact of cutting parameters on the components of cutting force: (a), (b) and (c) changes in feed and cutting depth (= 90 m/min); (d), (e) and (f)
changes in cutting speed and depth (= 0.1 mm/rev); the cutting speed and feed variation of (g), (h) and (i) (= 1 mm).

Fig. 21. The impact of tool edge radius and nose radius on the components of cutting force (= 90 m/min, = 0.15 mm/rev, = 1.75 mm): (a) Fx, (b) Fy and (c) Fz.

prediction is made of the cutting force applied during the actual turning iteration. The strain rate coefficient in the second plastic deformation
process. Based on the results of this work, the following conclusions can zone is determined by the minimum difference between the
be drawn: tangential stress at the tool-chip interface and the material strength
at the root of the chip in each iteration.
1 2-D finite element simulations are performed of simplified orthog­ 3 In combination with 304 stainless steel and Inconel 718 turning
onal cutting, the shear angle is estimated by observing the maximum experiments, the proposed model is validated. It is concluded that
strain rate in the shear zone, and a shear angle regression model is the average error between the predicted values of cutting forces and
established. The shear angle is on the rise with the increase in cutting the experimental values ranges between 4 % and 11 %, and the
thickness or cutting speed. The most significant one is cutting calculated results conform well to the measured results.
thickness, followed by cutting speed. 4 In the range of selected process parameters, the cutting force declines
2 The shear angle is known as the Oxley model input, and the method with the rise in cutting speed, and increases with the rise in feed and
for determining the strain rate coefficient of the first deformation cutting depth. Among them, the impact of cutting depth and feed on
zone and the second deformation zone is modified. The Strain rate cutting force is found to be significant. The combination of low depth
coefficient in the first deformation zone ceases to change during the

1316
T. Zhou et al. Journal of Manufacturing Processes 58 (2020) 1304–1317

of cut, low feed rate and high cutting speed could facilitate the cut­ [17] Zou Z, He L, Jiang H, Zhan G, Wu J. Development and analysis of a low-wear
micro-groove tool for turning Inconel 718. Wear 2019;420:163–75.
ting force to be minimized.
[18] Jiang H, He L, Yang X, Zou Z, Zhan G. Prediction and experimental research on
5 With the increase of edge radius, each component of the cutting force cutting energy of a new cemented carbide coating micro groove turning tool. Int J
exhibits a linear increase. As the radius of nose increases, the main Adv Manuf Technol 2017;89:2335–43.
cutting force and cutting depth resistance show an increasing trend, [19] Sheikh-Ahmad J, He Y, Qin L. Cutting force prediction in milling CFRPs with
complex cutter geometries. J Manuf Processes 2019;45:720–31.
while the feed resistance drops slightly. In general, the combination [20] Deng B, Zhou L, Peng F, Yan R, Yang M, Liu M. Analytical model of cutting force in
of small edge radius and nose radius is conducive to reducing the micromilling of particle-reinforced metal matrix composites considering interface
cutting force. failure. J Manuf Sci Eng 2018;140:081009.
[21] Weng J, Zhuang K, Zhu D, Guo S, Ding H. An analytical model for the prediction of
force distribution of round insert considering edge effect and size effect. Int J Mech
Funding Sci 2018;138:86–98.
[22] Molinari A, Moufki A. A new thermomechanical model of cutting applied to
turning operations. Part I. Theory. Int J Mach Tools Manuf 2005;45:166–80.
This work was supported by National Natural Science Foundation of [23] Fu Z, Chen X, Mao J, Xiong T. An analytical force mode applied to three-
China [Grant No. 51765009 and No. 51665007]. dimensional turning based on a predictive machining theory. Int J Mech Sci 2018;
136:94–105.
[24] Laakso SVA, Agmell M, Stahl J-E. The mystery of missing feed force - the effect of
Declaration of Competing Interest friction models, flank wear and ploughing on feed force in metal cutting
simulations. J Manuf Processes 2018;33:268–77.
The authors declare that they have no known competing financial [25] Johnson GR, Cook WH. Fracture characteristics of three metals subjected to various
strains, strain rates, temperatures and pressures. Eng Fract Mech 1985;21:31–48.
interests or personal relationships that could have appeared to influence
[26] Fangjuan Z. Research on machined surface characteristics of 304 stainless steel.
the work reported in this paper. Huazhong University of Science and Technology; 2014.
[27] Cockcroft MG, Latham DJ. A simple criterion of fracture for ductile metals. 1966.
References [28] Jing X, Li H, Wang J, Tian Y. Modelling the cutting forces in micro-end-milling
using a hybrid approach. Int J Adv Manuf Technol 2014;73:1647–56.
[29] Bosetti P, Maximiliano Giorgio Bort C, Bruschi S. Identification of Johnson–Cook
[1] Zhang W, Wang X, Hu Y, Wang S. Predictive modelling of microstructure changes, and tresca’s parameters for numerical modeling of AISI-304 machining processes.
micro-hardness and residual stress in machining of 304 austenitic stainless steel. J Manuf Sci Eng 2013;135.
Int J Mach Tools Manuf 2018;130:36–48. [30] Lotfi M, Amini S, Sajjady SA. Development of a friction model based on oblique
[2] Zhu Z, To S, Zhu W-L, Huang P, Zhou X. Cutting forces in fast-/slow tool servo cutting theory. Int J Mech Sci 2019;160:241–54.
diamond turning of micro-structured surfaces. Int J Mach Tools Manuf 2019;136: [31] Waldorf DJ, DeVor RE, Kapoor SG. A slip-line field for ploughing during
62–75. orthogonal cutting. J Manuf Sci Eng 1998;120:693–9.
[3] Sahoo P, Pratap T, Patra K. A hybrid modelling approach towards prediction of [32] Pu C, Zhu G, Yang S, Yue E, Subramanian S. Effect of dynamic recrystallization at
cutting forces in micro end milling of Ti-6Al-4V titanium alloy. Int J Mech Sci tool-chip interface on accelerating tool wear during high-speed cutting of AISI1045
2019;150:495–509. steel. Int J Mach Tools Manuf 2016;100:72–80.
[4] Maurel-Pantel A, Fontaine M, Thibaud S, Gelin J-C. 3D FEM simulations of [33] Shaw MC, Cookson J. Metal cutting principles. New York: Oxford University Press;
shoulder milling operations on a 304L stainless steel. Simul Modell Pract Theory 2005.
2012;22:13–27. [34] Grzesik W. Advanced machining processes of metallic materials: theory, modelling
[5] Wan M, Ye X-Y, Yang Y, Zhang W-H. Theoretical prediction of machining-induced and applications. Elsevier; 2008.
residual stresses in three-dimensional oblique milling processes. Int J Mech Sci [35] Jomaa W, Songmene V, Bocher P. Predictive analytical modeling of cutting forces
2017;133:426–37. generated by high-speed machining of ductile and hard metals. Mach Sci Technol
[6] Bai W, Sun R, Leopold J, Silberschmidt VV. Microstructural evolution of Ti6Al4V in 2017;21:335–61.
ultrasonically assisted cutting: numerical modelling and experimental analysis. [36] Administration I, Group EPGAM, Boothroyd G. Temperatures in orthogonal metal
Ultrasonics 2017;78:70–82. cutting. Proc Inst Mech Eng 1963;177:789–810.
[7] Wang Z. High-speed milling of titanium alloys: modeling and optimization. [37] Su J-C. Residual stress modeling in machining processes. Georgia Institute of
National University of Singapore; 2005. Technology; 2006.
[8] Merchant ME. Mechanics of the metal cutting process. I. Orthogonal cutting and a [38] Young H, Mathew P, Oxley P. Allowing for nose radius effects in predicting the
type 2 chip. J Appl Phys 1945;16:267–75. chip flow direction and cutting forces in bar turning. Proc Inst Mech Eng Part C-J
[9] Oxley P. B. Mechanics of machining, an analytical approach to assessing Eng Mech Eng Sci 1987;201:213–26.
machinability. New York, USA. 1989. [39] Arsecularatne J, Mathew P, Oxley P. Prediction of chip flow direction and cutting
[10] Adibi-Sedeh AH, Madhavan V, Bahr B. Extension of Oxley’s analysis of machining forces in oblique machining with nose radius tools. Proc Inst Mech Eng Part B-J Eng
to use different material models. J Manuf Sci Eng 2003;125:656–66. Manuf 1995;209:305–15.
[11] Lalwani D, Mehta N, Jain P. Extension of Oxley’s predictive machining theory for [40] Young HT, Liao YS, Yang YC, Mathew P. Predicting the chip flow for nose radius
Johnson and Cook flow stress model. J Mater Process Technol 2009;209:5305–12. tools under oblique machining conditions. J Chin Inst Eng 1993;16:825–34.
[12] Huang Y, Liang S. Cutting forces modeling considering the effect of tool thermal [41] Moufki A, Dudzinski D, Le Coz G. Prediction of cutting forces from an analytical
property—application to CBN hard turning. Int J Mach Tools Manuf 2003;43: model of oblique cutting, application to peripheral milling of Ti-6Al-4V alloy. Int J
307–15. Adv Manuf Technol 2015;81:615–26.
[13] Pan Z, Shih DS, Tabei A, Garmestani H, Liang SY. Modeling of Ti-6Al-4V machining [42] Orra K, Choudhury SK. Mechanistic modelling for predicting cutting forces in
force considering material microstructure evolution. Int J Adv Manuf Technol machining considering effect of tool nose radius on chip formation and tool wear
2017;91:2673–80. land. Int J Mech Sci 2018;142:255–68.
[14] Davis B, Dabrow D, Ifju P, Xiao G, Liang SY, Huang Y. Study of the shear strain and [43] Suresh R, Basavarajappa S, Gaitonde V, Samuel G. Machinability investigations on
shear strain rate progression during titanium machining. J Manuf Sci Eng 2018; hardened AISI 4340 steel using coated carbide insert. Int J Refract Met Hard Mater
140:051007. 2012;33:75–86.
[15] Pang L, Hosseini A, Hussein H, Deiab I, Kishawy H. Application of a new thick zone [44] Echávarri BE. Flow and fracture behaviour of high performance alloys. Universidad
model to the cutting mechanics during end-milling. Int J Mech Sci 2015;96: Politécnica de Madrid; 2012.
91–100. [45] Sun J, Guo Y. Material flow stress and failure in multiscale machining titanium
[16] Li B, Wang X, Hu Y, Li C. Analytical prediction of cutting forces in orthogonal alloy Ti-6Al-4V. Int J Adv Manuf Technol 2009;41:651–9.
cutting using unequal division shear-zone model. Int J Adv Manuf Technol 2011;
54:431–43.

1317

You might also like