Sie sind auf Seite 1von 14

International Journal of Machine Tools & Manufacture 46 (2006) 367–380

www.elsevier.com/locate/ijmactool

Predictive force model for ball-end milling and experimental


validation with a wavelike form machining test
M. Fontaine*, A. Devillez, A. Moufki, D. Dudzinski
Laboratoire de Physique et Mécanique des Matériaux, UMR CNRS 7554, ISGMP, Université de Metz, Ile du Saulcy, 57045 Metz, France

Received 20 December 2004; accepted 13 May 2005


Available online 1 August 2005

Abstract
This paper presents a predictive force model for ball-end milling based on thermomechanical modelling of oblique cutting. The tool
geometry is decomposed into a series of axial elementary cutting edges. At any active tooth element, the chip formation is obtained from an
oblique cutting process characterised by local undeformed chip section and local cutting angles. This method predicts accurately the cutting
force distribution on the helical ball-end mill flutes from the tool geometry, the pre-form surface, the tool path, the cutting conditions, the
material behaviour and the friction at the tool-chip interface. The model is applied for a complex surface which is a wavelike form used as a
validation machining test. The results are compared with experimental data obtained from ball-end milling tests performed on a 3-axis CNC
equipped with a Kistler dynamometer.
q 2005 Elsevier Ltd. All rights reserved.

Keywords: Cutting forces; Ball-end milling; Tool run-out; Thermomechanical modelling; Oblique cutting; Complex surface; Wavelike form; Validation test

1. Introduction used in the computation of the cutting forces acting on


the tool. Several published works proposed accurate
The prediction of cutting forces in machining is essential solutions to calculate forces distribution on rigid cutters
to enhance NC codes and then contributing to improve [1–4] or on flexible cutters [5–7]. Afterwards, these classical
reliability, accuracy and productivity in CNC machining. In models were extended to the case of curved or complex
fact, it may give information about cutter deflection, surfaces [8–16]. All these models are efficient to study tool
machine tool chatter, tool wear and breakage, and then deflection, chatter vibrations, surface errors, tool run-out
tool life and surface integrity can be optimised in selecting and cutting conditions optimisation. However, they are
appropriate cutting conditions. Modelling of cutting dependent on empirical laws giving the cutting force
operations is the focus of many research programs to follow coefficients necessary to perform the calculations. These
the development of High Speed Machining and to go cutting force coefficients are obtained by fastidious and
towards simulation of machining processes. The purpose of expensive experiments, orthogonal or oblique turning tests
this work is to develop a cutting force model for ball-end or directly milling tests, with varying cutting conditions.
milling process. This process is used extensively in the The purpose of this work is to present an original and
manufacturing of free form surfaces such as dies, moulds predictive model using the oblique cutting approach
and many aerospace components. proposed by Moufki, Dudzinski and Molinari [17–19]. In
The mechanistic approach is often used for the modelling this model, the cutting forces calculation is achieved by
of ball-end milling. This approach has demonstrated that introducing the thermomechanical behaviour of the work-
orthogonal or oblique cutting force data may be successfully piece material and the tool-chip interface friction charac-
teristics. In the same way as that in a mechanistic approach,
the ball-end milling operation is supposed carried out by a
* Corresponding author. series of infinitesimal cutting edge elements. Each elemen-
E-mail address: fontaine@lpmm.univ-metz.fr (M. Fontaine). tary straight line cutting element is machining the work-
0890-6955/$ - see front matter q 2005 Elsevier Ltd. All rights reserved. piece material in oblique conditions depending of
doi:10.1016/j.ijmachtools.2005.05.011 its position on the tool’s cutting edge. In other words,
368 M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380

the ball-end milling process is constructed as an aggregation surface which consists of a cylindrical surface with a radius
of oblique cuts. R0 and a hemisphere surface with the same radius R0. Each
The obtained model is applied to ball-end milling on cutting edge is considered as a helix with a constant lead, the
a 3-axis machine. The tool inclination in regard of helix angle on the cylindrical part of the tool is noted i0.
main machine axis is then not considered here but For a point with positive altitude z, the cutter radius in the
the proposed approach may be easily adapted for more (x, y) plane at this point is:
complex machining configuration. The case of sculptured pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
surface machining is treated by introducing a wavelike form If z% R0 : RðzÞ Z R20 K ðR0 K zÞ2
(1)
machining test which is very useful to validate a complete If zO R0 : RðzÞ Z R0
process modelling and to obtain information about the tool
response for variable cutting conditions (in terms of surface The local point P, on the cutting edge referenced j at
inclination and axial depth of cut). height z, is located by its angular position jj(z) mesured
Finally, to validate the modelling, the predicted forces positively from the y-axis and defined by:
are compared with measured forces during the chosen ball- 2p
end milling operation on a 42CrMo4 steel, for which jj ðzÞ Z q K Dj C ðj K 1Þ (2)
Nt
material characteristics such as strain hardening, strain rate
sensitivity and thermal softening are known. where Nt is the number of teeth (cutting edges) and q is the
rotation angle of the mill also measured positively from the
y-axis around the z-axis. Dj is the lag angle between the tool
tip (zZ0) and a point on the helical flute at height z. For a
2. Geometry of the ball-end mill
constant lead length, it is obtained from:
z z
2.1. Global geometry Dj Z tan i0 Z tan iðzÞ (3)
R0 RðzÞ
The global geometry of the ball-end mill is described in In this case, the local helix angle i(z) is defined by:
Fig. 1(a). The mill parameters are defined in a Local  
Coordinate System (x, y, z) (noted LCS) linked to the tool tip RðzÞ
iðzÞ Z tanK1 tan i0 (4)
or tool end E. The cutting edges lie on the tool envelope R0

(b) dFr
κ
P' E t er dFψ
z
P eψ
x n eκ
dFx x
P C
θ ∆ψ
dFy dFκ
y ψ y
dF z ψ

(a)

z
ω
z
Pn
i0 R0 αn
(c)
λS
C dz
n = -er
P' Ps t
z eψ P
V dw eκ λS
E R( z )
x dFψ
dFκ
dFr

Fig. 1. Geometry of the ball-end mill. (a) Global geometry, (b) local geometry, (c) elementary cutting edge.
M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380 369

For ball-end mills, a constant lead length is more 3. Geometry of the wavelike form machining test
common than a constant local helix angle. The constant
lead, which implies a variable helix angle along the flute, is 3.1. Description of workpiece surface
preferred by the cutter grinders to save material during re-
grinding operation. However, cutting mechanisms are more The initial pre-form surface, or uncut surface, is here a
uniform with constant helix cutters, which require varying
plane normal to the vertical machine Z-axis. The tool z-axis
lead [20]. These two cases are approximations because of
and the machine Z-axis are merged (3-axis milling
re-grinding and compensation operations used to optimise configuration). In this first approach, the mill is assumed
the tool design [21], then the real geometry is difficult to to be rigid (no deflection occurs). The reference point of the
obtain from tool manufacturers. A more accurate represen- Global Coordinate System (GCS) is taken at the intersection
tation of the helical cutting edge profile on the ball-end part of three solid planes at a corner of the workpiece, point O in
can be used for specific tools from measuring points along Fig. 2. The Local Coordinate System (LCS) associated to
the cutting edge and fitting this data by a polynomial the tool is fixed at the tool tip E, Figs. 1 and 2(a, b).
function [3,11,14,22]. Nevertheless, it is fastidious to re-
Two inclination angles could be add between LCS and
calculate the real cutting edge geometry for each new tool.
GCS, in particular, a tool-surface inclination could be
The presented approximation with a constant lead length imposed by the part design or imposed during machining to
gives a good correlation with the mills used in the limit cutting forces, avoid chatter or enhance surface finish
experimental validation. [4,10,23]. This inclination is typically constant on a 3-axis
milling machine but may vary with more degrees of freedom
2.2. Local geometry available. The influence of the inclination angles of the tool
or of the workpiece in the Global Coordinate System on the
The engaged cutting edges are decomposed into a series cutting forces are studied in parallel works. The purpose of
of linear elementary cutting edges corresponding to the this paper is only to test and validate a cutting force model
axial increment dz, Fig. 1(a). For a current point P at height for non-linear tool paths.
z, the local unit base vectors er, ek, ej are introduced, The local altitudes of pre-form and post-form surfaces
associated to the spherical coordinates (R0, k, j) of this can be calculated for a complex shape by an analytical
point, Fig. 1(b). They are in the radial direction and in relation (polynomial, canonical, NURBS) or by a discrete
transverse directions of increasing j and increasing k representation (Z-Map, Z-Displacement Vectors). It is
respectively. The angle k is function of height z: chosen here to define analytically the initial and final
  surfaces in order to simplify computing calculations.
K1 R0 K z The chosen machined surface is shown in Fig. 2(c), it is
If 0% z% R0 : k Z cos
R0 (5) comparable with those presented by several authors in
p recent works [14,16,22]. In the plane (X, Z), the final
If zO R0 : kZ
2 machined surface is composed of two circle arcs with a
The elementary cutting edge at point P is drawn in the same radius RS, one concave and the other convex. The
tangent plane Ps defined by the unit vectors ek and ej at P, difference of altitude DZ with the initial form varies
Fig. 1(c). The local cutting velocity vector is collinear to the between a maximum in the concave arc and a minimum
direction of ej and the local cutting edge inclination angle ls in the convex one, noted respectively DZmax and DZmin.
is the angle, measured in the tangent plane Ps, between the These extreme values correspond to the extreme axial depth
elementary cutting edge and the normal direction to the of cut that occurs along the given tool path. The same value
cutting velocity vector, it is given by: of DZ is chosen at the entry, at the middle and the exit of the
workpiece. It is noted DZ0 and verifies the following
sin ls tan ls relation:
tan i Z Z 5 ls Z tanK1 ðtan i sin kÞ (6)
cos ls sin k sin k
DZ0 Z ðDZmax C DZmin Þ=2 (9)
The local normal rake angle an, measured in the plane Pn
normal to the elementary cutting edge, is defined constant. RS can be calculated from:
The elementary width of cut dw is: ðL0 =4Þ2 C ðDZmax K DZ0 Þ2
RS Z (10)
dz 2ðDZmax K DZ0 Þ
dw Z (7)
sin k where L0 is the workpiece length on the X-axis.
The local cutting velocity V is simply calculated from:
V Z u RðzP Þ (8) 3.2. Description of tool path

with u the spindle frequency in rad/s and R the radius of the The tool path is here defined by the movement of a
mill at height zP, Eq. (1). fixed point of the mill: the ball part centre C. The contact
370 M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380

Fig. 2. Wavelike form machining test for 3-axis ball-end milling. (a) Machined workpiece, (b) geometrical representation, (c) pre-form and post-form surfaces
definition, (d) path of tool ball centre C.

point between the tool envelope and the specific post- If X1–2%XC%L0 then:
form surface has to be considered to respect geometrical qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
tolerances. The chosen reference tool path is shown in ZC ðXC Þ Z H0 K RS K DZmin C R2C2 K ð3L0 =4 K XC Þ2
Fig. 2(d). In the plane (X, Z), it is also composed of two (15)
circle arcs with adapted radius RC1 and RC2 centred
respectively in C1 and C2. This radius values are where XC and ZC are the coordinates of point C in the GCS,
deducted from RS by: H0 and L0 the initial dimensions of the workpiece
respectively on X and Z-axis, Rs the radius of circle arcs
RC1 Z RS K R0 (11) describing the post-form surface, RC1 and RC2 the radii of
the circle arcs centred at C1 and C2, respectively, and DZmax
and DZmin the extreme values of axial depth of cut.
RC2 Z RS C R0 (12) The feed vector per tool revolution noted f has a constant
modulus but varies continuously in direction in order to
where R0 is the radius of the mill ball part.
homogenise the cutting conditions along the tool path. The
The change of reference tool path between circle arc 1
feed per tooth vector ft is deduced from the feed vector, it has
and circle arc 2 is not located at L0/2, its position, noted X1–2
the same direction and its modulus is:
on Fig. 2(d), is calculated from:
  ft Z f =Nt (16)
L0 R0
X1–2 Z 2K (13) where f is the modulus of feed vector f and Nt the number of
4 RS
teeth of the mill.
The altitude of point C in GCS can be defined by: The feed per tooth vector ft is projected on the three axes of
If 0RXC%X1–2 then: the GCS and decomposed in three feed components noted ftx,
fty and ftz, Fig. 3. In our application, the resultant feed direction
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ZC ðXC Þ Z H0 C RS K DZmax K R2C1 K ðL0 =4 K XC Þ2 is only in the plane (X, Z), then the component fty is zero.
The feed per tooth vector makes a variable angle fz with
(14) the X-axis. This feed inclination is due to the tool trajectory
M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380 371

y
Previous tool Current point on
position the cutting edge

f tx P

Current position ∆p
of the mill at Z = ZP
Rp

Cp
Tool position on
previous path for XC = XP

Y
XP
Fig. 3. Axial depth of cut, decomposition of feed per tooth and undeformed
chip thickness t0 for a point P in (x, z) plane. X
Fig. 4. Relative position of current point P with previous tool positions in
the horizontal plane for (ZZZp).
and can be calculated from the inclination of the tangential
line to the tool path:
If 0%XC%X1–2 then:
   – the relative position of the workpiece uncut
L surfaces along the three axes and specifically on
4z Z sinK1 XC K 0 =RC1 (17)
4 Z-axis (upper initial or pre-form surface), Fig. 3,
– the previous tool path considered without any tool
If X1–2%XC%L0 then: deflection and with perfect surface finish, Fig. 4,
  
3L0 – and the path of the previous tooth which is defined
4z Z sin K1
K XC =RC2 (18) as a circle with the Martelloti’s equation
4
approximation [24], Fig. 4.
The angle fz can present negative values in case of
downward ramping. The position of point P is obtained by:
The feed components ftx and ftz can be expressed as: 0 1 0 1
Xp XC C x P
ftx Z ft cos 4z (19) B C B C
@ YP A Z @ YC C yP A (21)
ftz Z ft sin 4z (20) ZP GCS ZC K R0 C zP
with
0 1 0 1
xP RðzÞsinðjj ðzÞÞ
@ yP A B C
Z @ RðzÞcosðjj ðzÞÞ A (22)
4. Cutting force model
zP LCS z
4.1. Tool engagement determination where XC, YC and ZC are the coordinates of point C in GCS,
R(z) and jj(z) the radius and angular position at altitude z in
At any time or angular position of mill, each point LCS.
considered along each cutting edge is localised in the GCS Five conditions are necessary to check the engagement of
(X, Y, Z). To define if a current point P of a cutting edge is in the elementary cutting edge into the workpiece material:
working conditions, its relative position with the pre-form The length L0 and the width W0 of the initial workpiece
workpiece surface has to be determined. For each current on X and Y-axis respectively give simple conditions on local
point P, it is necessary to check: tool engagement. P is in cutting position (in GCS) if:
372 M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380

Conditions 1 and 2: where ftx and ftz are the projections of the instantaneous feed
per tooth ft respectively on the x and z-axis, Eqs. (19) and
0! XP ! L0 and 0! YP ! W0 (23)
(20).
The local altitude of uncut surface can be calculated for a A modified Martelloti’s equation is obtained; it takes into
complex shape by an analytical relation or by a discrete account of the position along the cutting edge and the two
representation (Z-Map [4], Z-Displacement Vectors [11], components of the feed. The sign of t0 has to be checked to
piece and tool meshing [14]). Here, to simplify calculation, determine the position of current point P beside the previous
the pre-form surface is a plane corresponding to a roughing tooth path:
or direct semi-finishing operation. Hence, the local altitude Condition 5:
of the uncut surface is directly obtained in the GCS from the
initial workpiece height H0: P is in cutting condition if t0 O 0 (32)
Condition 3:
If these five conditions are all verified, the elementary
P is in cutting position ðin GCSÞ if 0! ZP ! H0 (24) cutting edge defined at point P is considered engage in the
workpiece material.
Another method is to compare the position of point P in
the LCS with the axial depth of cut da, Fig. 3:
Condition 3 0 : 4.2. Tool run-out

P is in cutting position ðin LCSÞ if zP ! da (25) The previous equations and conditions describe a rigid
with and stable process where no deflection and no vibration
occur. These disruptive factors are easily avoidable by
da Z H0 K ZC C R0 (26) limiting and controlling the cutting conditions and by
The actual position of the tool in the plane (X, Y) containing choosing stiff tools and tool-holders. But there is another
the point P is also compared with the previous tool path, more problem which is difficult to control: the tool run-out. In
precisely, with the tool position on the previous tool path when milling, this geometrical default can be due to tool itself
its axis was at the coordinate XP, Fig. 4. The point of this axis at (wear, asymmetry, insert setting, dynamic imbalance and
altitude ZZZP is noted CP. A condition is obtained by thermal deformation) but it is mainly due to the offset
comparing the modulus of vector PCp (vector from P to Cp, between the position of the tool rotation axis and the spindle
Fig. 4) with the radius of the previous path mill RP: rotation axis. The consequence is a tool rotation around the
Condition 4: spindle axis with an eccentricity. This eccentricity modifies
the tool engagement and the local cutting conditions
P is in cutting position if kPCP kO RP (27) (cutting velocity and angles). Then, the run-out has a direct
with effect on the cutting forces level and variation. It depends on
the geometrical quality of the spindle and tool holder. Its
kPCP k Z jDp C yP j (28) effect is particularly significant when the undeformed chip
where Dp is the path interval (distance between two tool thickness t0 reaches small values. In this case, one or several
paths on Y-axis), and cutting edges may cut no material. The tool run-out has to be
considered first for the modelling of machining defaults in
RP Z R½zP C ZC ðXC Þ K ZC ðXP Þ (29) milling operations because of its large influence on cutting
From Eqs. (1), (14) and (15). forces and surface quality. An accurate modelling permits to
The current position of the point considered has to be take into consideration all existing machine configurations
compared with the path of the previous tooth to define if it is and propose fault diagnosis [26] and compensation methods
inside or outside the pre-cut region. The path of the previous as variation of feed rate [16,27].
cutting point is assumed to be a circle in the plane (x, y) at The tool run-out due to spindle/tool offset can be divided
the altitude zP as shown in Fig. 4. This Martelloti’s into an axis tilt and a centre offset. The axial run-out due to
approximation [24] gives a simplified relation for the axis tilt is usually taken into account for face-milling
calculation of the undeformed chip thickness t0 on the operations where the surface roughness is very sensitive to
cylindrical part of the mill (zPRR0): the insert cutters location and where the large values of tool
diameter increase its influence. In the case of ball-end
t0 Z ftx sin jj (30) milling operations, the limited tool diameter and the round
On the ball part of the mill (0%zP%R0), this value is form of the tool allows to neglect this axis tilt if the tool
obtained by a scalar product between the total feed vector ft length is limited.
and the vector er normal to the tool envelope at point P [25], The radial run-out due to centre offset is always modelled
Fig. 3: by a parallel axis eccentricity and a lag angle from the first
tooth position, noted respectively e and je on Fig. 5(a). For
t0 Z f t $er Z ftx sin k sinjj K ftz cos k (31) a ball-end mill, this locating angle je has to be considered
M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380 373

On the ball part of the mill (0%zP%R0), if the


(a) eccentricity e is smaller than the feed per tooth ft, the value
of undeformed chip thickness can be obtained by the
following scalar product:
t0 Z ðf t C e K ejK1 Þ$er Z f t $er C ðe K ejK1 Þ$er (35)
e ∆ψ
where ft is the feed per tooth vector with its components
CR e P defined by Eqs. (19) and (20), er the vector normal to the
tool envelope at point P, e the eccentricity vector for the
z x position considered and ejK1 the eccentricity vector when
Cm ψe the previous tooth was in the same angular position at the
previous position of the mill. These two vectors are defined
by:
e Z e sin q x C e cos q y;
θ (36)
ejC1 Z e sinðq K 2p=Nt Þ x C e cosðq K 2p=Nt Þ y
θ
The equivalent radius Re of the cutter at point P (radial
y distance of point P from the rotation axis) is calculated to
determine the value of cutting velocity V, Fig. 5(b). From
(b) the projection of point P on the direction of vector e, noted
CR P 0 , the following relation can be found:
x sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi
2p
Re(zP) Re ðzP ÞZ RðzP Þ2 Ce2 C2eRðzP Þcos je KDjCðjK1Þ
e Nt
R(zP)
(37)
Cm P
ψ e − ∆ψ The local cutting velocity is then calculated with:
θ VZu Re ðzP Þ (38)
y The cutting speed direction varies too and the local
angles an (normal rake angle) and ls (edge inclination
angle) slightly change. All these angle variations have no
P' significant effect on global forces and then are neglected in
this approach.
Fig. 5. Radial run-out parameters. (a) Radial run-out description,
(b) equivalent radius Re(zP) of the mill at point P.
4.3. Thermomechanical oblique cutting model

from the first cutting edge inclination at the tool tip. Cm(zP) For application to ball-end milling modelling, the main
is the centre of the mill at altitude zP and CR(zP) is the and useful equations of the thermomechanical oblique
projection of point Cm(zP) on z-axis which is the spindle cutting model [18,19] are reminded and summarised.
rotation axis. The vector formed between these two points For an elementary cutting edge, the chip element is
from CR(zP) to Cm(zP) is the eccentricity vector and is noted obtained under oblique cutting conditions, Fig. 6(a). The
e. The reference angular position of the spindle (qZ08) is primary shear zone is modelled as a shear band of constant
taken with the eccentricity vector e collinear to the y-axis. thickness h, as shown in Fig. 6(b). The strain due to the other
The angular position jj(zP) of point P on the cutting edge deformation zones is neglected.
j at height zP is now referenced by: The normal shear angle fn is assumed to be given by a
2p modified Merchant law:
jj ðzP Þ Z ðq C je Þ K Dj C ðj K 1Þ (33)
Nt fn Z A1 C A2 ðan K lf Þ (39)
The position of point P in LCS is now defined by: with
0 1 0 1
xP RðzÞsinðjj Þ C e sin q lf Z tanK1 ðmf Þ (40)
@ yP A B C
Z @ RðzÞcosðjj Þ C e cos q A (34)
where lf is the mean friction angle at the tool-chip interface
zP LCS z and mf the mean friction coefficient taken as an average
374 M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380

where t, g, g_ and T represent respectively the shear stress, the


shear strain, the shear strain rate and the absolute temperature.
The characteristics of material behaviour are the strain
hardening exponent n, the strain rate sensitivity coefficient
m, the thermal softening exponent n and constants A and B, g_ 0 ,
Tr (reference temperature) and Tm (melting temperature).
The strain rate can be extracted from the constitutive law
(43) and written as:
 pffiffiffi 
t 3 1
_ t0 Þ Z g_ 0 exp
g_ Z gðg; K (44)
mg1 ðgÞg2 ðTÞ m
  n    
g T K Tr n
g1 ðgÞ Z A C B pffiffiffi ; g2 ðTÞ Z 1 K
3 Tm K Tr
(45)
The shear strain rate g_ corresponds to the material
derivative of the shear strain g. The material flow through
the band is supposed stationary and depending only on one
dimension: the coordinate zs, Fig. 6(b). Then it is obtained
_
for the strain rate g:
dg _ t0 Þ
gðg;
Z (46)
dzs V cos ls sin fn
From considering that the deformation in the chip is
limited to the band, the shear strain at the entry of the shear
band (zsZ0) is equal to zero:
g0 Z 0 (47)

Fig. 6. Elementary cutting edge and chip formation. (a) Direction of chip
At the outflow (zsZh), there is an additional boundary
flow (zfl) on the rake face and elementary cutting forces, (b) shear band condition:
parameters in plane (Pn).  
1 cos an
gh Z (48)
value in this study. A1 and A2 are constants depending on the cos hs sin fn cosðfn K an Þ
workpiece material. Equations of motion are reduced to a single relation:
The chip flow angle hc on the tool rake face, Fig. 6(a), is
given by the oblique cutting model with the assumption that dt
Z rV cos ls sin fn g_ (49)
the frictional force and the chip flow directions are collinear. dzs
It is calculated from the following implicit equation: with relation (46) and after integration, that leads to:
2
cosðfn Kan Þsin fn sin hc Ktan ls cos ðfn Kan Þcos hc t Z tðg; t0 Þ Z rðV cos ls sin fn Þ2 g C t0 (50)
Cðcos an Ksinðfn Kan Þsin fn Þtan lf sin hc cos hc where t represents the stress distribution in the primary
Ctan lf tan ls sinðfn Kan Þcosðfn Kan Þcos2 hc Z0 ð41Þ shear band, r the material density, V the cutting velocity, ls
the edge inclination angle, fn the normal shear angle, g
The angle hs characteristic of the shear direction in the the shear strain in the band and t0 the stress at the entry
primary shear plane is determined by: of the band.
  The integrated form of Eq. (46) and the boundary
K1 tan hc sin fn Ktan ls cosðfn Kan Þ
hs Z tan (42) conditions (47) and (48) are used to determine the stress t0
cos an
at the entry of the band (for zsZ0):
Theworkpiecematerialissupposedtobeisotropicandrigid ð gh
V cos ls sin fn
(elastic deformations neglected). The thermomechanical dg K h Z 0 (51)
0 _ t0 Þ
gðg;
response is described by the following Johnson-Cook law:
  n      A non-linear equation is obtained, with Eqs. (44), (45)
1 g g_ T KTr n
tZ pffiffiffi ACB pffiffiffi 1Cmln 1K and (50), the only unknown parameter is the shear stress t0.
3 3 g_ 0 Tm KTr The last important relation for the material flow through
(43) the band is given by the conservation of energy written with
M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380 375

the assumption of adiabatic shearing: represented on Figs. 1 and 6(a), are obtained by:
0 1 0 1
dFx ðq; z; jÞ sin jj sin k sin jj cos k cos jj
dT B C B C
rcV cos ls sin fn Z btg_ (52) B dFy ðq; z; jÞ C ZB cos jj sin k cos jj cos k Ksin jj C
@ A @ A
dzs
dFz ðq; z; jÞ Kcos k sin k 0
0 1
dFr ðq; z; jÞ
The first term corresponds to the material derivative of
B C
absolute temperature T in the band, r and c are respectively !B C
@ dFk ðq; z; jÞ A ð56Þ
the material density and its heat capacity. The second term
assumes that the fraction b (Taylor–Quinney coefficient) of dFj ðq; z; jÞ
plastic work is converted into heat. Using the previous They are summed for each edge j of the elementary disk
relations (46) and (50), it is obtained: considered:
0 1
  XNt

b g2 B dFx ðq; z; jÞ C
T Z Tðg; t0 Þ Z Tw C rðV cos ls sin fn Þ2 C t0 g 0 1 B B jZ1
C
C
rc 2 dFx ðq; zÞ BX Nt C
(53) B C B B
C
dFy ðq; z; jÞ C
@ dFy ðq; zÞ A Z B C (57)
B jZ1 C
dFz ðq; zÞ B N C
BX t C
where Tw is the absolute temperature in the workpiece @ dF ðq; z; jÞ A
z
before cutting. jZ1
The elementary cutting forces dFr, dFk and dFj
and summed for each elementary disk of the engaged length
represented on Fig. 6(a) correspond respectively to the
of the mill (axial depth of cut da) in order to obtain the
thrust force, the lateral force and the cutting force applied to
instantaneous forces on the whole cutter function of mill
the elementary cutting edge at point P in the (er, ek, ej)
angular position q:
frame. Their values on corresponding axes are evaluated
0 da 1
from the equilibrium of the forces applied to the elementary ð
chip after the outflow of the primary shear zone: B dFx ðq; zÞ C
B C
B C
0 1 B z¼0 C
Fx ðqÞ B dða C
dFr Z KdFs cos hs sin fn K Ns cos fn ; B C B
B C
C
@ Fy ðqÞ A Z B dFy ðq; zÞ C (58)
dFk Z dFs cos hs ½tan hs cos ls C cos fn sin ls  B C
Fz ðqÞ B C
B z¼0 C
B dða C
K dNs sin fn sin ls ; (54) B C
@ dF ðq; zÞ A
z
dFj Z KdFs cos hs ½tan hs sin ls K cos fn cos ls  z¼0

K dNs sin fn cos ls


5. Experimental validation

where dFs is the elementary shear force value on xs-axis 5.1. Experimental setup
(Fig. 6(c)), and dNs the elementary normal force value on
zs-axis (Fig. 6(a)), at the exit of the primary shear zone: A large number of cutting tests were performed without
lubricant on a vertical 3-axis milling machine to validate the
proposed modeling of ball-end milling. The cutting tests
t0 dw
dFs Z K t ; were conducted on a 42CrMo4 steel (equivalent to AISI
cos ls sin fn h 4142). It has the following chemical composition, in
(55)
tanðfn K an Þ C tan lf cos hc addition of iron (%mass):
dNs Z cos hs dFs C(0.425), Cr (1.180), Mn (0.749), Si (0.266), Ni (0.232),
tan lf cos hc tanðfn K an Þ K 1
Cu (0.227), Mo (0.154), S (0.029), Al (0.027), As (0.022), P
(0.016), N (0.008), Ti (0.005).
where th is the shear stress at the exit of the band given by A Kistler six-components dynamometer (model 9265B)
Eq. (50) for gZgh. has been used to measure cutting forces. The output signals
Finally, the elementary oblique force components dFr, were recorded and stocked on a PC Dewetron console
dFk and dFj are projected on the LCS axes corresponding through a height-channel Kistler charge amplifier.
here to the machine or table axes. The elementary values of A two fluted uncoated ball-end mill (referenced 51221)
cutting forces dFx, dFy and dFz (in LCS or CGS) as from Diager-Industrie company with a nominal diameter of
376 M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380

5.2. Model parameters

The shear angle was obtained by using coefficients which


are, for this material, closely the same than those proposed
by Merchant: A1Z408; A2Z0.5.
The main friction coefficient was here an average value
determined from those identified in works of Moufki et al.
[28]. It depends mainly on the cutting and feed speeds. For
example, for fZ0.1 mm/tooth:

mfZ1.11 (lfZ488) for UZ2653 rev/min


mfZ1 (lfZ458) for UZ3980 rev/min
mfZ0.90 (lfZ428) for UZ5000 rev/min

The friction coefficient value can be superior to 1


because it reflects coupled phenomena appearing at the tool-
chip interface (seizure and sliding with plastic shearing).
These values are approximations but the model could be
used in an enhanced form with the introduction of a
temperature dependant friction law [19] to enhance the
model accuracy and study the temperature distribution
along the cutting edge. Here, an average value gives good
Fig. 7. Wavelike tests on a specific workpiece device. (a) Slotting: DpZ
results for cutting forces prediction.
12 mm, (b) roughing: DpZ3 mm. In a previous work [29], quasi-static and high strain rate
compression tests were conducted on a 42CrMo4 steel over
the strain rate range 10K3 to 104 sK1 and at temperatures
12 mm, a nominal helix angle of 178 and a normal rake from room temperature to 700 8C. The flow stress data of
angle of 08 on ball part was used in the experiments. these tests were used to determine the parameters of the
R0Z6 mm; i0Z178; anZ08. Johnson–Cook constitutive law (43):
The workpieces were parallelepipeds with a height H0 of
20 mm, a length L0 and a width W0 of 50 mm, Fig. 7. The AZ612 MPa; BZ436 MPa; g_ 0 Z0.001 sK1;
wavelike form machining validation test was defined by the nZ0.15; mZ0.008; nZ1.46;
surface topography presented in Fig. 2 with DZmaxZ3 mm TrZTwZ293 K; TmZ1793 K.
and DZminZ0.2 mm corresponding to typical depths of cut in
roughing and finishing operations respectively. In this view, The other useful material parameters are:
the results presented below give some information about
rZ7800 kg/m3; cZ500 J/(kg K), bZ0.9.
cutting force levels that can occur in such typical operations.
From these values and by using Eqs. (6)–(9), it is obtained the The primary shear zone thickness is taken as hZ0.025 mm
additional values DZ0Z1.6 mm, RSZ56.5 mm, RC1Z which is a typical value found in works of Shaw [30].
50.5 mm and RC2Z62.5 mm, characterising the machined
surface. The experiments were performed with three values
of spindle frequency: UZ2653, 3980 and 5000 rpm, and four 6. Results
different values of feed rate: ftZ0.05, 0.1, 0.15, 0.2 mm/tooth
and per revolution. Slot tests were first conducted, Fig. 7(a), The results presented here correspond to a slotting test in
then successive tool paths through the workpiece were up-milling, Figs. 8 and 9, and a roughing test conducted in
carried out with several values of the path interval DpZ6 and up-milling too, Fig. 10, but all the stable signals were
3 mm (roughing), Fig. 7(b), 1.5 mm (direct semi-finishing), checked.
and 0.6 mm (direct finishing). In each case, up-milling and
down-milling were tested. 6.1. Experimental cutting forces
The tool run-out was measured with a dial indicator fixed
on the machine table and touching the cylindrical part of the The signals were recorded for all the length of the
mill. The eccentricity value e is obtained by dividing workpiece [11] and the voltage/time signals were trans-
the measured lag value on a tool revolution by two, and the formed into force/rotation angle ones [22]. The signals on X-
locating angle is observed from the mill angular position axis are lightly low-pass filtered at 360 Hz (slotting test) and
corresponding to the maximum measured value of 540 Hz (roughing test) to suppress some accelerometer noise
eccentricity. due to vibration. This axis is more sensible to vibration
M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380 377

Fig. 8. Measured and predicted cutting force Fx for slotting and tool run-out influence. UZ2653 rev/min, ftZ0.1 mm/tooth, DpZ12 mm, eZ0.01 mm, jeZ08.
(a) Global force envelope, (b) zoom for daZ3 mm.

because of cutting configuration and because it is the main simulation for the slot cutting test at cutting speed of
axis of the piezoelectric sensors which tends to present a 2653 rpm and with a feed rate of 0.1 mm per tooth. Fig. 8(a)
typical accelerometer response when frequency grows up. corresponds to the complete wavelike form machining test
(a full simple path), Fig. 8(b) gives the measured and
6.2. Simulated cutting forces predicted force Fx only for five revolutions of the mill
during the test and corresponding to an axial depth of cut
The program was all executed with Fortran 8.5 on a PC with daZ3 mm. More precisely, the first Fig. 8(b) gives the
a 2.6 GHz single processor. The computation time is less than cutting force on each tooth and during these five revolutions.
3 min (slotting) and 2 min (roughing) for calculation It can be observed that the difference of cutting force level
increments on q and z as dqZ108 and dzZ0.1 mm. between the two teeth of the mill is regular and important. In
fact, at each mill rotation, one tooth is more engaged in the
workpiece than the other and that leads to higher local
7. Discussion cutting forces. It can be determined a ratio of about 30% on
force levels between the two teeth. The two other Fig. 8(b)
7.1. Influence of tool run-out show that it is essential to take into account of tool run-out
in the ball-end milling model. With run-out simulation, the
The Fig. 8 shows measured and predicted cutting force force evolution predicted by the model is very closed to the
Fx (main feed direction) without and with run-out measured one. In the same way, for the full slotting test,
378 M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380

Fig. 9. Measured and predicted cutting forces Fy and Fz for slotting. UZ2653 rev/min, ftZ0.1 mm/tooth, DpZ12 mm, eZ0.01 mm, jeZ08.

Fig. 8(a), modelling radial tool run-out reduces significantly It is mainly due to the fact that the material flow which
the discrepancy between predicted and measured forces, and occurs around the cutting edge is neglected in the modelling
finally, the predicted global force envelope is very similar to approach, in particular, the cutting edge is supposed
the measured one. A good correlation appears between perfectly sharp. The material flow and associated shearing
experimental and simulated cutting forces when radial run- occurring at the clearance face lead to a ploughing force.
out is considered. The same observation can be made for the The ploughing force level becomes very significant around
other force components, Fig. 9, and tests with other cutting the tool end when cutting velocity and undeformed chip
conditions and successive tool paths with a smaller interval thickness tend to zero. These limit cutting conditions appear
Dp, Fig. 10. The shape of the envelope of measured and at the tool tip and the resultant ploughing force value is high
predicted cutting forces is similar and the amplitude slightly in this region. The direction of this ploughing force is
differs. mainly normal to the tool envelope (er-direction) and at the
tool tip, this direction is closed to be the z-direction. That is
7.2. Shape of force envelope why the force Fz is more affected by this phenomenon and
the predicted Fz values are lower than the measured ones.
The fast increase followed by a maximum peak at the For the same reason, the asymmetry between downward
beginning of the cutting test represents clearly the entry of ramping sequence and upward ramping sequence is more
the mill in the workpiece and it seems to be well evident on Fz measured curves. The explanation is that in
predicted by the model, Figs. 8–10. The fast decrease downward ramping configuration, the tool tip is widely used
followed by instabilities at the end of the machining test and the resultant ploughing force increases. Hence, the
due to the exit of the mill from the workpiece can be difference between measured and predicted forces is
observed too in both measured and calculated cutting proportional to the existing ploughing force.
forces. The asymmetry between downward ramping This ploughing force can be added in the model by
sequence and upward ramping sequence, for same values introducing edge coefficients as in mechanistic approaches
of depth of cut, is retrieved too. However, it is more [2], but in this case, the level of prediction decreases. The
pronounced on the measured signals. ploughing effect can be limited by using a controlled tool-
workpiece inclination in 5-axis machining, for example.
7.3. Amplitude Finally, according to the fact that its influence occurs mainly
on Fz force component, which is less important for tool
The global force amplitude is well predicted for Fx and deflection, tool vibration calculation and then for surface
Fy cutting force components, but a more important finish prediction, it was not taken into account of this effect
amplitude offset appears on Fz (up to 30%), Figs. 8–10. in our approach.
M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380 379

Fig. 10. Measured and predicted cutting forces for roughing. UZ3980 rev/min, ftZ0.1 mm/tooth, DpZ3 mm, eZ0.01 mm, jeZ208.

7.4. Influence of cutting conditions down milling gives better surface finish, particularly for
small values of tool path interval Dp.
The global forces level decreases with increase of cutting
velocity V tested values. In this context, higher values of V
tend to stabilise the process and signals. The difference 8. Conclusions
between measured and calculated forces tends to decrease
when V increases. The thermomechanical approach Ball-end milling is here modelled by using an analytical
adopted here gives better results at higher cutting speed approach of oblique cutting, applied for each active cutting
[19]. The feed rate variation affects proportionally the edge element. The tool engagement in the workpiece
cutting forces due to the variation of the undeformed chip material is defined by the relative position of this cutting
thickness t0. The tests corresponding to fZ0.05 mm/tooth, edge element with the initial workpiece dimensions, the
not presented here, give unstable signals. The decrease of local altitude of uncut surface and the previous tool path.
the tool path interval Dp naturally reduces the cutting forces This approach can be adapted for more complex surfaces
values, Fig. 10, but the recorded signals for small values of and tool paths by modifying the description of the machined
Dp (finishing operations) are often unstable and then surface and the boundary conditions. A wavelike form
difficult to analyse. These signals could be compared with machining test was introduced and used to validate the
an enhanced version of the model taken into account the tool proposed model and to gather very complete information
deflection and chatter vibrations calculated from cutting about the process, the tool and material behaviour.
forces for rigid case. Finally, the up cutting configuration The obtained results give very good approximation for
tends to stabilise the cutting process and the signals but the the cutting forces and the main experimental tendencies are
380 M. Fontaine et al. / International Journal of Machine Tools & Manufacture 46 (2006) 367–380

retrieved. The modelling accuracy allows to analyse the [11] R. Zhu, S.G. Kapoor, R.E. DeVor, Mechanistic modeling of the ball-
effect of different defaults as tool run-out or ploughing on end milling process for multi-axis machining of free form surfaces,
J. Manufact. Sci. Eng. 123 (2001) 369–379.
cutting force level. This model is useful to gain an [12] B.W. Ikua, H. Tanaka, F. Obata, S. Sakamoto, Prediction of
understanding of the cutting phenomena and to simulate cutting forces and machining error in ball end milling of curved
the machining process in order to enhance surface integrity, surfaces—I. Theoretical analysis, J. Int. Soc. Prec. Eng.
tool life, stability and productivity by optimising cutting Nanotech. 25 (2001) 266–273.
conditions, tool path, tool-workpiece inclination and even [13] B.W. Ikua, H. Tanaka, F. Obata, S. Sakamoto, T. Kishi, T. Ishii,
Prediction of cutting forces and machining error in ball end milling of
tool geometry. curved surfaces-II Experimental verification, J. Int. Soc. Prec. Eng.
Nanotechnol. 26 (1) (2002) 69–82.
[14] I. Lazoglu, Sculpture surface machining: a generalized model of ball-
Acknowledgements end milling force system, Int. J. Mach. Tools Manufact. 43 (5) (2003)
453–462.
[15] A. Lamikiz, L.N. Lopez de Lacalle, J.A. Sanchez, M.A. Salgado,
The authors would like to thank sincerely their colleagues Cutting force estimation in sculptured surface milling, Int. J. Mach.
from the ENSAM of Metz (Ecole Nationale Supérieure Tools Manufact. 44 (2004) 1511–1526.
d’Arts et Métiers) and particularly Olivier Bomont for his [16] J.H. Ko, D-W. Cho, Feed rate scheduling model considering
help in preparing and programming the experimental tests. transverse rupture strength of a tool for 3D ball-end milling, Int.
J. Mach. Tools Manufact. 44 (2004) 1047–1059.
They also thank Lionel Schivre from their laboratory for his
[17] D. Dudzinski, A. Molinari, A modelling of cutting for viscoplastic
response time and accuracy in manufacturing the workpiece materials, Int. J. Mech. Sci. 39 (4) (1997) 369–389.
device and Florent Gaillard from Diager-Industrie for his [18] A. Moufki, A. Molinari, D. Dudzinski, M. Rausch, Thermovisco-
help in choosing and obtaining tools. plastic modelling of oblique cutting: forces and chip flow, Int.
J. Mech. Sci. 42 (6) (2000) 1205–1232.
[19] A. Moufki, A. Devillez, D. Dudzinski, A. Molinari, Thermomech-
anical modelling of oblique cutting and experimental validation, Int.
References J. Mach. Tools Manufact. 44 (2004) 971–989.
[20] S. Engin, Y. Altintas, Mechanics and dynamics of general milling
[1] M. Yang, H. Park, The prediction of cutting force in ball end milling, cutters, Part I: helical end mills, Int. J. Mach. Tools Manufact. 41
Int. J. Mach. Tools Manufact. 31 (1) (1991) 45–54. (2001) 2195–2212.
[2] P. Lee, Y. Altintas, Prediction of ball-end milling forces from [21] Y-C. Tsai, J-M. Hsieh, A study of a design and NC
orthogonal cutting data, Int. J. Mach. Tools Manufact. 36 (9) (1996) manufacturing model of ball-end cutters, J. Mater. Proc. Technol.
1059–1072. 117 (2001) 183–192.
[3] A Azeem, H-Y Feng, L Wang, Simplified and efficient calibration of a [22] B.U. Guzel, I. Lazoglu, Increasing productivity in sculpture surface
mechanistic cutting force model for ball-end milling, Int. J. Mach. machining via off-line piecewise variable feedrate scheduling based
Tools Manufact. 44 (2–3) (2003) 291–298. on the force system model, Int. J. Mach. Tools Manufact. 44 (1)
[4] G.M. Kim, C.N. Chu, Mean cutting force prediction in ball-end (2003) 21–28.
milling using force map method, J. Mater. Proc. Technol. 146 (3) [23] I. Lazoglu, S.Y. Liang, Modeling of ball-end milling forces
(2003) 303–310. with cutter axis inclination, J. Manufact. Sci. Eng, Trans.ASME 122
[5] C. Sim, M. Yang, The prediction of the cutting force in ball-end (2000) 3–11.
milling with a flexible cutter, Int. J. Mach. Tools Manufact. 33 (1993) [24] M.E. Martelloti, An analysis of the milling process, Trans. ASME 67
267–284. (1941) 233–251.
[6] H.Y. Feng, C.H. Menq, A flexible ball-end milling system model for [25] T-I. Seo, Intégration des effets de déformation d’outil en génération de
cutting force and machining error prediction, J. Manufact. Sci. Eng. trajectoire d’usinage, Ecole Centrale de Nantes-Université de Nantes,
118 (1996) 461–469. PhD Thesis, 1998.
[7] G.M. Kim, B.H. Kim, C.N. Chu, Estimation of cutter deflection and [26] R. Zhu, R.E. DeVor, S.G. Kapoor, A model-based monitoring and
form error in ball-end milling processes, Int. J. Mach. Tools Manufact. fault diagnosis methodology for free-form surface machining process,
43 (2003) 917–924. J. Manufact. Sci. Eng. 125 (2003) 397–404.
[8] E.M. Lim, H-Y. Feng, C-H. Menq, Z-H. Lin, The prediction of [27] C.C. Tai, K.H. Fuh, A predictive force model in ball-end milling
dimensional error for sculptured surface productions using the ball- including eccentricity effects, Int. J. Mach. Tools Manufact. 34 (7)
end milling process. Part 1: Chip geometry analysis and (1994) 959–979.
cutting force prediction, Int. J. Mach. Tools Manufact. 35 (8) [28] A. Moufki, A. Devillez, D. Dudzinski, A. Molinari, Thermomech-
(1995) 1149–1169. anical modelling of cutting and experimental validation Metal Cutting
[9] E.M. Lim, C-H. Menq, The prediction of dimensional error for and High Speed Machining, Kluwer Academic, Darmstadt, 2002, pp.
sculptured surface productions using the ball-end milling process. Part 51–67.
2: Surface generation model and experimental verification, Int. [29] A. Molinari, A. Moufki, D. Dudzinski, Study on behaviour of
J. Mach. Tools Manufact. 35 (8) (1995) 1171–1185. 42CrMo4 Steel CREAS Ascometal, 1997.
[10] B.M. Imani, M.A. Elbestawi, Geometric simulation of ball-end [30] M.C. Shaw, Metal Cutting Principles, Oxford Science Publications,
milling operations, J. Manufact. Sci. Eng. 123 (2001) 177–184. Oxford, 1984.

Das könnte Ihnen auch gefallen