Chatter Stability of Machining Operations
Chatter Stability of Machining Operations
Gabor Stepan,
Professor of Applied Mechanics, Budapest University of Technology and Economics
[email protected]
Erhan Budak,
Professor, Faculty of Engineering and Natural Sciences, Sabanci University, Turkey
[email protected]
Tony Schmitz,
Professor, Fellow ASME
University of Tennessee, Knoxville
Joint Faculty, Oak Ridge National Laboratory
[email protected]
Abstract
This paper reviews the dynamics of machining and chatter stability research since the first stability
laws were introduced by Tlusty and Tobias in the 1950s. The paper aims to introduce the
fundamentals of dynamic machining and chatter stability, as well as the state of the art and research
challenges, to readers who are new to the area. First, the unified dynamic models of mode coupling
and regenerative chatter are introduced. The chatter stability laws in both the frequency and time
domains are presented. The dynamic models of intermittent cutting, such as milling, are presented
and their stability solutions are derived by considering the time-periodic behavior. The
complexities contributed by highly intermittent cutting, which leads to additional stability pockets,
and the contribution of the tool’s flank face to process damping are explained. The stability of
parallel machining operations is explained. The design of variable pitch and serrated cutting tools
to suppress chatter is presented. The paper concludes with current challenges in chatter stability of
1. Introduction
Chatter continues to be the main limitation in increasing material removal rates, productivity,
surface quality, and dimensional accuracy of machined parts. F. W. Taylor, who was considered
to be the initiator of manufacturing engineering, declared that chatter was the “most obscure and
delicate of all problems facing the machinist’’ in his 1906 ASME article [1]. The early
investigations on chatter mechanisms were conducted by Arnold [2], who classified machining
vibrations as forced and self-induced types. He conducted several turning experiments at varying
speeds and feeds and reported the influence of cutting conditions and tool wear on the shape of the
vibration waves and stability. Doi and Kato presented the effect of time lag in the thrust force
relative to the chip thickness variation as the source of chatter instability in turning [3]. The first
scientific stability laws were independently presented at almost at the same time by Tobias et al.
[4] and Tlusty et al. [5] in the 1950s. Tlusty presented an absolute stability law that predicted the
critical depth of cut proportional to the machine’s dynamic stiffness and inversely proportional to
the material’s cutting force coefficient in orthogonal cutting regardless of the spindle speed. Tobias
presented a similar stability law, but included the effect of spindle speed, i.e., the regenerative time
delay, which led to the stability pockets or “lobes” in orthogonal cutting. The dynamics of cutting
and chatter stability models were reviewed by Tlusty [6, 7] and Altintas and Weck [8]. General
literature reviews on the sources of nonlinearities in the dynamics of cutting [9] and chatter [10]
were also presented. Altintas et al. presented the frequency and time domain chatter stability
prediction laws [11].
The chatter theories have been applied to the analysis of machine tools to improve their dynamic
stiffness through design modifications or by adding passive and active dampers. Machine tools are
dynamically tested; the critical mode shapes which affect their chatter stability are identified and
modified to strengthen their stiffness by machine tool designers. Merritt converted the stability
equations of Tlusty and Tobias into a closed-loop system with unity and delayed feedback to
consider regeneration and solved the stability using the Nyquist criterion [12]. Tlusty showed that
the productivity gain is proportional to the dynamic stiffness improvements at the tool-workpiece
contact zone in the direction of chip generation [13]. The dynamics of various metal cutting and
grinding operations have been modeled by several researchers and their stability was solved by
applying the one-dimensional, frequency domain stability laws of Tlusty and Tobias. Tlusty
pioneered the development of high-speed milling machines by stating that the product of spindle
speed and the number of teeth on the cutter (i.e., the tooth passing frequency) should be matched
with the first bending mode of the spindle to operate the machine at the highest speed (first, or
rightmost) lobe which leads to highest material removal rate [7]. High speed-high power spindles
were developed to remove more than 95% of the material from aluminum slabs to produce
monolithic, lightweight parts for the aerospace industry [14, 15]. As the application domains have
grown, the limitation of Tobias’s and Tlusty’s one-dimensional stability theories have also been
investigated. It was observed that as the spindle speed is reduced relative to the natural frequency
of the machine, the stability of the process increases due to process damping. Tobias attributed this
process damping effect to the penetration of the cutting edge into the wavy cut surface [16] which
is still studied at the present. Process damping significantly depends on the ratio of the cutting
speed to the chatter frequency caused by the dominant mode. In the presence of low-frequency
Sridhar [19] showed that milling operations have coupled dynamics in two directions with periodic
coefficients; therefore, the stability of such multi-point machining applications cannot be solved
by the one-dimensional chatter theories of Tlusty and Tobias. Minis and Yanushevsky [20, 21]
applied Flouquet theory to solve the stability of two dimensional, periodic milling operations in
the frequency domain. Budak and Altintas determined the stability limit by considering the
harmonics of tooth passing frequencies in the eigenvalues [22, 23], as well as a direct solution of
stability analytically by considering the averages of directional factors [24] in the frequency
domain. The research group of Stepan presented a semi-discrete time-domain stability solution of
a single [25] and two-directional milling operations [26]. The research in chatter stability
continues due to its complexity and application to the design of machine tools and cutting tools,
as well as their chatter-free and productive use in machining operations.
The aim of this paper is not a general review of the chatter literature, but rather the presentation of
mathematical modeling of the dynamic cutting process and its stability solution in the frequency
and time domains; see Sections 2 and 3. The mechanism behind the appearance of stability islands
for highly intermittent cutting operations at high spindle speeds is explained in Sections 2.3 and
3.3. The design of special-purpose cutters to suppress chatter is presented with the aid of stability
models in Section 4, and the complexities of parallel machining dynamics are discussed in Section
5. The paper is intended to provide strong foundations in dynamic cutting and chatter stability
while explaining the current research challenges including improved accuracy (Section 6). The
paper is concluded in Section 7 by pointing out the future challenges in chatter research. The paper
is dedicated to the two pioneering scientists, the late Professors Tobias and Tlusty, who contributed
immensely to chatter stability literature that led to the present high-speed machine tool and high-
performance machining technologies.
h(t ) h0 hd (t ) h0 r (t ) r (t T ) , (1)
where the delay T is the time between the two passes, h0 is the commanded (static) chip thickness
which corresponds to the feed per revolution in orthogonal turning (Figure 1a), and ℎ𝑑 is the
dynamic (time-varying) chip thickness. The chip thickness (h(t )) produces variable tangential
( Ft ) and normal ( Fr ) cutting forces which form the resultant cutting force ( Fc ) as:
Ft (t ) Kt ah(t ) , Fr (t ) K r Ft (t ) K r Kt ah(t )
, (2)
Fc (t ) Ft 2 (t ) Fr 2 (t ) K c ah(t ) K c Kt 1 K r2
where Kt is the tangential cutting force coefficient, Kr Fr / Ft tan is the radial to tangential
force ratio, and a is the depth of cut. The edge components of the cutting forces, which depend
only on the depth of cut, are neglected since they don’t contribute to the regeneration of chip
thickness. The cutting force coefficients are either identified from the orthogonal to oblique cutting
transformation using Merchant’s thin shear plane model [27, 28] or mechanistically evaluated from
cutting tests by considering the chip thickness and thermal softening effects [29]. The resultant
cutting force excites each spring i attached to the tool by projection into the corresponding mode’s
direction:
Each mode with a transfer function of (ii ( s)) will cause vibration xi ( s) in the Laplace domain
( s ) as:
xi ( s) ii ( s) Fi ( s) . (4)
By projecting and superposing all the vibrations in the chip thickness (r) direction, vibration in
this direction is calculated according to:
l
r ( s ) o ( s ) Fc ( s ) o ( s ) sin i cos i i ( s) , (5)
i 1
where o ( s) is the oriented transfer function between the vibrations in the chip thickness and
resultant cutting force directions. By neglecting the static chip thickness ( h0 ), which does not
affect stability, the dynamic or regenerative chip thickness contributed by the vibrations is
evaluated as:
hd (t ) r (t ) r (t ) r (t T ) , (6)
where the time delay is T 60 / n for the spindle speed n [rev/min]. Substituting the dynamic
chip thickness into the resultant cutting force (Eq. (2)) yields the dynamic cutting force in the
Laplace domain as:
1 Kc a(1 e sT ) 0 ( s) 0 , (8)
which has an infinite number of roots due to the delay term e sT . By substituting 𝑠 = 𝑖𝜔𝑐 for the
critical stability condition, the characteristic equation of the dynamic cutting system in the
frequency domain becomes:
1 1
1 Λ 0 (ic ) 0 Λ aK c 1 eicT
0 (ic ) G0 (c ) iH 0 (c )
, (9)
where Go and H o are the real and imaginary parts of oriented frequency response function, or FRF
( 0 ). Expanding Eq. (9) gives:
1
Λ
G0 (c ) iH 0 (c )
aK c 1 e icT aK c 1 cos cT i sin cT
(10)
1 aK c
G0 1 cos cT H 0 sin cT iaK c G0 sin cT H 0 1 cos cT 0
and equating the imaginary part to zero leads to the critical spindle speed ( n ) :
By also equating the real part to zero and substituting H 0 / G0 sin cT / cos cT 1 , the critical
depth of cut alim is determined to be:
1 1
alim . (12)
H ( ) 2 KcG0 (c )
KcG0 (c ) 1 cos cT 0 c sin cT
G0 (c )
The solution leads to a positive depth of cut only when the real part of the oriented FRF (G0 (c ))
is negative. Tobias considered the entire negative region of the real part of the FRF which leads
to the critical depth of cut and corresponding spindle speeds for each integer number of vibration
waves ( kl ) generated within the time delay (i.e., spindle period) T that led to the stability lobes
[4]. Tlusty considered the worst case by taking the minimum of G0 (c ) which led to the minimum
depth of cut regardless of the spindle speed [5]:
1
amin . (13)
2Kc min G0 (c )
The eigenvalue (Eq.(9)) in mode coupling does not contain the delay term and the system must
have at least two flexibilities which affect the chip thickness through the excitation by the cutting
force. Consider a case with two flexibilities which are orthogonal to each other, but oriented from
the velocity and chip thickness directions, i.e. 2 1 / 2 . The characteristic equation (9)
becomes:
Since alim is a real number, the imaginary part of the oriented FRF must have a zero crossing, i.e.,
H 0 (mc ) 0 at the mode coupling chatter frequency mc . Mode coupling chatter is independent
of speed and has the limiting depth of cut:
which leads to about two times more stable depth than regenerative chatter. The orientation of the
cutting force and modes alter the stability as indicated by Eq. (14). If the imaginary part of the
oriented FRF does not cross zero, mode coupling would not occur. Ismail et al. showed that the
mode coupling stability can be improved by selecting the ratio of modes through the modification
of end mill stiffness [30].
Case 2: Regenerative chatter stability with process damping
Consider that the system is flexible only in the chip direction ( r ) and the force is also in the same
direction for simplicity to explain the process damping mechanism ( / 2, / 2) . The
clearance face and edge radius of the tool have time-varying ploughing contact with the presently
cut wavy surface (Figure 1b) which generates a damping force, aC p r / vc , proportional to the
process damping constant of the material ( C p [ N / m] ) and the ratio of the vibration velocity (𝑟̇ =
𝑑𝑟/𝑑𝑡) to the cutting speed (vc ) as described by Das and Tobias [31]. Assuming a structure with
stiffness ( k ) , a viscous damping coefficient ( c ) , and a mass ( m ) with a corresponding natural
frequency n k / m and damping ratio c / (2 km ) is excited by the regenerative cutting
and process damping forces, Fr (t ) K r Kt ah(t ) aC p r / vc , the equation of motion for orthogonal
cutting is expressed as:
n2 n2 r
r 2n r n2 r Fr (t ) K r K t ah(t ) aC p (15)
k k vc
Note that the process damping term contains vibration velocity ( r ), which increases the viscous
damping of the system. Eq. (15) is expressed in the Laplace domain as:
2 2 1
s 2 2n s n 2 r ( s) n K r Kt a h0 1 e sT r ( s ) n
k k v C p sr ( s )
c
(16)
r (s) n 2 / k
( s) 2
Fr ( s) s 2n s n 2
where ( s) is the transfer function of the system’s structural dynamics in the chip thickness
direction. Eq. (16) was expressed as a closed-loop system by Merrit [12], but without the process
damping term which is now included in Figure 2. The closed-loop transfer function of the system
is expressed as:
Cp
s
V
e sT
Figure 2 Block diagram of dynamic orthogonal cutting with process damping and regenerative
chip feedback.
Cp
1 a ( s) s
h( s ) vc
. (17)
h0 ( s ) C
1 a (1 e sT ) K r K t p s (s)
vc
The characteristic equation of the system 1 a (1 e sT ) Kt (C p / vc )s (s) has an infinite
number of roots ( s ic ) due to the delay term e sT
. By considering the critical stability
limit, where the real part of the root is zero ( 0 ) and the structure’s FRF is represented by its
real and imaginary parts ( (ic ) G (c ) iH (c ) ), the characteristic equation becomes:
C p C
1 K r K t alim G 1 cos cT H sin cT c i K r K t alim G sin cT p H 1 cos cT 0 .
vc vc
(18)
Note that regardless of whether the system has a regenerative delay ( e jcT ) or not such as in mode
coupling, the process damping affects the system dynamics as shown in Eq. (18). Tlusty et al. [5]
and Tobias et al. [4] neglected the process damping term (i.e., C p 0 ) to find an analytical,
frequency domain stability solution in their initial publications as described in Case 2. By forcing
both the real and imaginary parts of Eq. (18) to zero, the critical depth of cut ( alim ) and spindle
speed ( n ) are found by Tobias and Fishwick [4] as :
1 c H (c )
alim [m] ; n[rev / min] 60 tan 1 , kl 0,1, 2,.. (19)
2 K r K t G (c ) (2kl 1) G (c )
which gives the absolute minimum depth of cut that is linearly proportional to the dynamic
stiffness of the machine ( 2k ) and inversely proportional to the product of the cutting force
coefficients for the selected work material ( Kr Kt ) . Tlusty’s simple formula (Eq.(20)) has been
widely used as a guide by machine tool designers and Tobias’s formula has been applied to select
the stable depth of cuts at high spindle speeds [7].
Sample stability lobes are shown for mode coupling, with and without the process damping effect,
Figure 3 Chatter stability diagrams and corresponding chatter frequency diagram for
mode coupling chatter, and regenerative chatter with and without process damping.
Simulation parameters: Cutting coefficients - Kt=1000 MPa, Kr=0.3; Modal parameters-
k1=15×106 N/m, k2=10×106N/m, n,1 150 Hz, n,2 250 Hz , 1 0.010, 2 0.012 . The
orientation of flexibilities- 1 70deg , 2 160deg ; process damping constant - Cp=2e6
N/m; workpiece diameter = 30 mm. Simulated cases: h0 0.1mm / rev, a 2.5mm. Cases-
A: n=700 rev/min (stable, process damping zone), B: n=5000 rev/min (stable, high speed
zone), C: n=8000 rev/min (unstable, high speed zone).
in
The spindle speed n is measured in [rev/min], so the angular velocity of the spindle is
f n / 60 1 / T in [Hz], and it is 2 / T 2 f in [rad/s], where the undamped natural
Kr Kt
r0 ah0 ah0
k
of the tool causes a surface location error. The resulting governing equation then assumes the form
2 / T / n f / f n ,
which is used to construct the stability charts in the parameter plane. The trial solution of the linear
delay-differential equation is the same as it is for the linear ordinary differential equations:
Due to delay term, there are infinite values for and that satisfy Eq.(24). For the stability of
stationary cutting, all the possible values must be negative for the exponentially decaying
vibrations. The values of and correspond to the characteristic exponents i in the
Laplace domain, which corresponds to the Laplace transformation applied to Eq. (16) in the
frequency domain with the relation s c . Accordingly, the stability boundaries can be found
~
in the (, a~) plane when Eq. (25) is substituted back into Eq. (24) with 0 . The separation of
~ ~
the coefficients of the cos(t ) and sin(t ) terms leads to two equations for the dimensionless
spindle speed () and the dimensionless depth of cut a ~ with the parameter . This can be
interpreted as a dimensionless chatter frequency since it relates to the chatter frequency c
appearing in Eq. (18) according to c / n fc / f n along the stability boundaries when process
damping is neglected. The corresponding closed-form expressions for the critically stable
dimensionless spindle speed () and depth of cut ( ( a ) ) are:
, kl 1, 2,...
1 1 2
kl arctan
2 (26)
( 1) 4
2 2 2 2
a , 1 .
2( 2 1)
These curves are also called stability lobes and they are presented in Figure 4 together with their
characteristic parameters. The major characteristic parameters of the chart can be identified from
Eq.(26). For example, the lower bound for all lobes is a straight-line at the absolute limit for the
which is the same absolute minimum as given in Eq. (20) via the frequency domain model for
small damping ratios. At these minimum parameter points of the lobes, the angular chatter
frequency c and the corresponding “worst” spindle speeds are:
1 2 4
c 1 2 n and n n , kl 1, 2,... (28)
1
kl arctan
1 4kl 1
1 2
Figure 4. Theoretical stability chart of turning processes in the plane of the dimensionless
spindle speed and the dimensionless depth of cut a . The dimensionless chatter
frequency fc / f n is greater than 1. The damping ratio is fixed at 0.05 , while
kl 1, 2,... refers to the sequence number of the lobes.
close to c n . Note that at the intersection of the lobes, quasi-periodic chatter may also occur
0
with two frequencies involved. The dimensionless stability model in the time domain provides the
critical speeds where the depths of cut are either maximum or minimum as a function of the natural
frequency of the structure.
2.3 Discrete time-domain analysis of highly interrupted cutting
As opposed to the continuous time domain analysis, the discrete time-domain analysis can also be
represented in case of the simplified model of highly interrupted orthogonal cutting shown in
Figure 5. Consider the mechanical model of the regenerative mechanism introduced in Figure 1,
where only one vibration mode is relevant at the angle 1 / 2 , so the tool position is given by
the single coordinate x : x1 r . The cylindrical workpiece has a rotation period T , which means
that its circumference is Tvc , where vc is the cutting speed. However, the cylindrical workpiece
has a large groove of width (1 )Tvc , and it is hypothetically assumed that this groove is almost
as large as the whole circumference of the workpiece. The tool is therefore in cut for only very
short time intervals T 0 , where the dimensionless parameter stands for the ratio of the time
spent in the cut to the time spent out of the cut. The depth of cut is still a , so during this very short
time of contact, the tool is subjected to a kind of impact load due to the cutting force applied during
the very short contact time T . Still, the regenerative mechanism appears again, since the impact
force will depend on the difference of the past and present positions of the tool during contact.
Note, that the same mechanical model can also be derived from the dynamics of milling presented
in Section 3 (see Figure 7). In that case, the rotating tool is considered to be rigid in the y direction
and flexible only in the x direction. Assume that the milling tool has only one cutting edge with
rotation period T , the cutting speed is vc again, and the circumference of the milling tool is Tvc .
position x j 1 x(t j 1 ) x(t j T ) and the initial velocity v j 1 x (t j 1 ) x (t j T ) . The tool has a
en t en t
x(t ) cos(dt ) x j 1 sin(d t ) v j 1
1 2 d
vi x((1 )T ) x(T ) , (30)
en t en t
x(t ) n sin(dt ) x j 1 cos(dt ) v j 1
1 2 1 2
where the phase angle is calculated from tan / 1 2 . The position and the velocity of the
tool can then be calculated when it enters the workpiece again by substituting t (1 )T T into
Eq. (30) :
where the negative sign in the superscript refers to the fact that these values occur at the start of
the short impact-like cutting. During this infinitesimally short cutting time of T , the variation of
the position of the tool is negligible, while the cutting force variation Fr K r Kt a x j 1 x j has
the linear impulse that causes the variation m v j v j of the tool’s linear momentum.
Accordingly, when the tool leaves the workpiece again at the time instant t j , its position and the
velocity can be calculated as:
kx kx
Substituting the values of x(0) and x(T ) from Eq. (30), a discrete time model is obtained that
describes the connection of the subsequent positions and velocities of the tool after each short
cutting period by means of the two-dimensional iteration:
x j A11 A12 x j 1
, j 1, 2,... (33)
v j A21 A22 v j 1
where
e nT e nT
A11 cos(dT ), A12 d sin(dT )
1 2
cos( T )
nT
A21 n e sin(dT ) T n2 a 1 e T
Kr Kt
1 2 kx 1 2 d
A22 e nT
1 2
cos( T ) T
d n a sin( T ) .
Kr Kt
kx d
The stability of the stationary highly interrupted cutting process is equivalent to the convergence
of the geometric vector series given in Eq.(33), which means that the quotients, 1,2 , that is, the
eigenvalues of the quotient matrix A in Eq.(33), must have moduli less than 1: 1, 2 1 .
In this discrete system (Eq.(33)), the counterpart of the classical chatter with angular frequency c
appears when the following characteristic equation has the complex conjugate roots
1,2 eic cos c i sin c on the unit circle of the complex plane which satisfy:
Using the same dimensionless quantities as defined in Eq.(23), the substitution of these roots into
Eq. (34) leads to the chatter boundaries a~c for the dimensionless axial depth of cut in the following
explicit form:
2
sinh
1 2 2 1 2
ac >ac,min , (35)
2
sin 1 2
where the lower estimate for the absolute stable region is not as sharp as it is in case of turning,
but the formula is quite good in the high spindle speed domain as shown in Figure 6. However,
there appears another kind of vibration that does not occur in turning. This is the result of a period-
doubling (or period-2) bifurcation, when the critical characteristic root is 1 1 , while 2 1 .
It is called period-doubling because the corresponding time period of the critical chatter vibration
MANU-19-1385 Altintas p.15
is two times the time period of the spindle rotation. If this is substituted back into Eq. (34), the new
kind of instability lobes have the following closed-form expression:
2 2
cos 1 2 cosh
1 2
>a 1 2
apd . (36)
2 2 2
pd,min
sin 1
These stability boundaries are presented in Figure 6. Compared to the stability chart of turning in
Figure 4, the number of lobes is doubled, the classical chatter lobes become much thinner and,
although the period-doubling lobes show up between them, large stable pockets are still present at
~
1 / k l similar to the case of turning. These stable pockets are relevant at high spindle speeds.
The chatter and period-doubling vibrations at the stability limits have a rich frequency content due
to the parametric excitation in the time-delayed system. This means that the critical self-excited
vibrations are not harmonic anymore, although they are periodic, and they still have a relevant
fundamental harmonic vibration component. The corresponding dominant frequency is usually the
lowest frequency among the many frequencies presented above the stability chart in Figure 6, but
Figure 6 Theoretical stability chart of highly interrupted cutting in the plane of the
dimensionless spindle speed and the dimensionless (axial) depth of cut a. The
dimensionless chatter frequencies are fc / f n . The damping ratio is fixed at 0.05 ,
0.1 .
Multi-point machining operations, such as milling, are carried out with tools having multiple teeth
that periodically cut the material. The dynamics and stability of milling operations are summarized
in the following sections.
3.1 Stability of milling operations in the frequency domain
Milling cutters have multiple teeth that have intermittent engagements with the workpiece. A
diagram of milling with dynamic flexibilities in feed ( x ) and normal ( y ) directions is shown in
Figure 7. If the tooth j is at the angular immersion ( j ) which is measured clockwise from the
( y ) axis, the dynamic chip thickness in the radial direction is generated by the vibrations at the
present and previous tooth periods ( x(t ) x(t ) x(t T ) , y (t ) y (t ) y (t T ) ) as:
The dynamic chip creates tangential ( Ftj Kt ah( j ) ) and radial ( Ftj ( j ) K r Ftj ) cutting forces at
each engaged tooth, which are projected in the feed ( x ) and normal ( y ) directions, and they are
summed to find the total force components exciting the structure [24].
Fx (t )
1 axx axy x(t ) 1
aKt F(t ) aKt A(t )Δ(t ) (38)
Fy (t )
2 a yx a yy y (t ) 2
The directional factors, which exist when the cutter is engaged with the workpiece (i.e., when a
tooth is located between the start st and exit ex angles), are:
N 1 N 1
axx g j [sin 2 j K r (1 cos2 j )] ; axy g j [(1 cos 2 j ) K r sin2 j )]
j 0 j 0
N 1 N 1
a yx g j [(1 cos2 j ) K r sin 2 j ] ; a yy g j [sin 2 j K r (1 cos2 j )]
j 0 j 0
Since the cutter rotates at angular speed (rad / s ) , the immersion angle is time-varying
( j t ( j 1) p p 2 / N ) . Consequently, the directional factors ( A (t ) ) are also time-
varying and periodic at the tooth passing interval (T ) , or at the tooth passing frequency
(T 2 / T ) in the frequency domain, or at cutter pitch angle intervals ( p T 2 / N ) in the
angular domain. The periodic, dynamic cutting force (Eq.(38)) can be transformed into the
frequency domain as:
1
𝑭(𝜔) = 2 𝑎𝐾𝑡 [𝑨(𝜔)(1 − 𝑒 −𝑖𝜔𝑐𝑇 )𝒒(𝜔)] ← 𝒒(𝜔) = {𝑥(𝜔) 𝑦(𝜔)}𝑇 = 𝜱(𝑖𝜔)𝑭(𝜔), (39)
where Φ j is the FRF matrix of the structure in the ( xx, xy; yx, yy ) directions. The periodic
directional factors can be represented by their Fourier series components as [22]:
ex
N axx axy irN
A( ) Ae r
irT t
; Ar
2 a a yy
e d , (40)
r st yx
The stability condition provided in Eq. (41) is challenging to determine due to the presence of the
delay term eicT and the corresponding periodicity of the directional matrix A( ) at the unknown
tooth passing frequency T . Minis and Yanushevsky [21] applied the Floquet theory to identify
the critical stable depth of cut (alim ) and spindle speed ( n ) iteratively. Budak and Altintas
proposed two approaches: the zero-order solution [24], where only the average of directional
factors is considered; and the multi-frequency [22] solution, which includes the harmonics of the
directional factors for low immersion, highly intermittent milling operations.
Zero-Order Solution: In this case, only the average component of the directional matrix is
considered so that [24]:
ex
N axx axy
A0
2 a
yx a yy
d . (42)
st
The dynamic milling (Eq. (41)) process then becomes time-invariant and its critical stability can
be found from the characteristic equation using:
1
R i I K t a (1 e icT ) det I ] A 0 (i ) 0 a0 2 a1 1 0 . (43)
2
From the computed real ( R ) and imaginary ( I ) parts of the eigenvalues, the critical stable
depth of cut (alim ) and spindle speed ( n ) are evaluated directly as [24]:
2 R
alim 1 2 I
NKt R
(44)
2 tan 1 kl 2 60
T (sec) , n(rev / min) , kl 0,1, 2, ..
c NT
Figure 8 Stability lobes predicted with the zero-order solution analytically and time-domain
numerical simulations for a cutter with two circular inserts.
Multi-Frequency Solution: When the harmonics of tooth passing frequency are included in the
directional matrix A( ) , the dynamic milling equation (Eq. (39)) is expanded using Floquet
theory to obtain [22]:
F( ) Pl c lT
l
(45)
1
F( ) A r l Φ c lT F( ) aKt 1 eicT
r l 2
This system has an infinite dimension, but it is truncated to few harmonics in practice. If the
intermittency is severe, such as the case for small radial depths of cut, more harmonics of the tooth
passing frequency (T ) need to be considered to obtain accurate results, which increases the size
of the eigenvalue problem. For example, if only one harmonic is used ( (r , l ) (0, 1) ), the
eigenvalue problem becomes:
P0 A
0 A1 A1 c P0
P1 A1 A0 A2 c T
P1
(46)
P
1 (61) A1
A2 A0 (66) c T P
1 ( 61)
(62)
which leads to six eigenvalues and the eigenvalue that gives the minimum depth of cut is selected
as a solution. Since the spindle speed is needed to find the tooth passing frequency (T ) , the
solution is obtained at each given spindle speed. Typically, even with the most severe intermittent
conditions, including two to three harmonics is sufficient for stability convergence. The multi-
frequency method is able to predict the added stability pockets (i.e., period-doubling lobes) at very
high spindle speeds which are beyond the natural frequency of the structure; see Figure 9 [42]. An
efficient elimination of false eigenvalues, which improves the computational speed, was developed
by Merdol [33, 47].
where the time delay is the tooth passing period T 2 /( N) , and the directional factor
A(t ) A(t T ) is time-periodic with the same time period T as the delay (for details, see Section
3.1).
In this case, full discretization may be applied, where the time-periodic coefficients and the time
derivatives are all discretized with discrete time intervals. This brute force method is successful,
but, as explained in Section 2.2, the construction of a stability chart takes several hours on a
standard personal computer. However, the combination of the continuous and discrete-time
domain methods explained in Sections 2.2 and 2.3, respectively, results in a numerical method
where the time derivatives are not discretized, but only the time periodicity and the time delay are.
The basic idea of the discretization of the delay is not trivial. However, a discrete-time mapping
can be constructed similar to the quotient matrix A of highly interrupted cutting, but with a much
larger size. The elements of the matrix can be calculated for each time step in closed form;
analytical solutions of linear non-homogeneous ordinary differential equations can be used; and
the computation time of the stability analysis can be reduced radically. Then, the stability
boundaries can be reconstructed in the same way as in the case of Eq.(33). Although the size of
the quotient matrix A is large, there are several advanced and efficient numerical methods and
related routines to check whether all eigenvalues of a large matrix have moduli less than 1.
The basic semi-discretization is now introduced. While it is straightforward to approximate the
time-periodic function A(t ) in Eq.(47) by the piecewise constant function:
where the discrete time step is t T / n ; see Figure 10a. It is more completed with respect to
the time delay to obtain an approximate system in an analytically manageable system of linear
is defined that refers back to the same time instant in the past for each time step:
x(t T ) x(t (t )) x(t (t (n int(t / t ))t )) x( j t T ) x(( j n)t )
(50)
t [ j t , ( j 1)it ), j 0,1, 2,...
The two kinds of discretization can be carried out with the same approximation number n (number
of time steps) since the delay and the time period are the same. While the time-periodicity of the
time delay seems to be a further complication, it makes the approximate system simpler since it
refers back to the same past value within one time step. This way, the time-periodic delay-
differential equation provided in Eq. (47) can be approximated by the linear non-homogeneous
ordinary differential equation:
aK t aK
x(t ) 2n x(t ) n2 1 A( j t ) x(t ) n2 t A( j t ) x(( j n) t ) (51)
kx kx
for each time interval t [ jt , ( j 1)it ), j 0,1,2,... . Although the average time delay in the
approximate system is somewhat larger than the exact delay T , the approximation is convergent
by decreasing the size of the time step t , that is, by increasing the approximation number n (see
[48] for details). Since the right hand side is piece-wise constant, the closed form solution of this
approximate system can be obtained similarly to Eq.(30) for the case of highly interrupted cutting.
This time, the discrete map expressed in Eq.(33) will have a much larger size. While the piece-
wise constant approximation of the periodic function A(t ) does not increase the size of this matrix,
the intermittent delayed state values do, since not just x (t ) and x(t T ) , but all the discrete states
x(( j i )t ) , ( i 0,1,..., n ) and their time derivatives have to be used as state variables at the j th
time instant t j jt , j 0,1,2,... . This way, the size of the quotient matrix will be at least
2(n 1) 2(n 1) , and its eigenvalues must be checked with appropriate numerical method to have
absolute values less than 1.
A typical stability chart is presented in Figure 11, where the effect of the time periodicity is not as
strong as in the case of highly interrupted cutting because the milling process with higher number
of teeth (N = 4), while the radial immersion is still far from full immersion. In this case, the
presence of the period-doubling instabilities is not so characteristic and is less relevant. Moreover,
as it was proven by Szalai and Stepan [49], these are not lobes anymore, but rather unstable islands
(or unstable pockets/lenses), which are mostly surrounded by the classical chatter instability lobes.
These unstable islands can still be separated and shifted to the stable domains as fully isolated
islands as shown experimentally by Zatarain et al.[50], which discusses also further complications
when modeling helical cutting edge tools.
Figure 12 Top) Time domain results for a period-2 bifurcation, i.e. chatter with added lobes.
(Bottom) Poincaré map for period-2 bifurcation. The sampled points align at a two fixed
locations for the period-2 bifurcation.
.
MANU-19-1385 Altintas p.25
of the forcing frequency). In this case, the once-per-tooth sampled points alternate between two
solutions. For period-n bifurcations, the sampled points appear at n distinct locations in the
Poincaré map.
An example period-2 bifurcation, the chatter with added lobes, is depicted in Figure 12. In the top
panel, the time domain displacement and velocity are shown individually. The inset provides a
magnified view; it is observed that the sampled points repeat every other tooth period, rather than
every period. The bottom panel shows the Poincaré map, where the sampled points appear at two
distinct locations.
A second analysis tool is the bifurcation diagram. Here, the independent variable, such as axial
depth of cut, is plotted on the horizontal axis against the once-per-tooth sampled displacement on
the vertical axis. The transition in stability behavior from stable (at low axial depths) to period-n
or secondary Hopf instability (at higher axial depths) is then directly observed. This diagram
represents the information from multiple Poincaré maps over a range of, for example, axial depths,
all at a single spindle speed. A stable cut appears as a single point (i.e., the sampled points repeat
when only forced vibration is present). A period-2 bifurcation, on the other hand, appears as a pair
of points offset from each other in the vertical direction. This represents the two collections of
once-per-tooth sampled points from the Poincaré map. A secondary Hopf bifurcation is seen as
vertical distribution of points; this represents the range of once-per-tooth sampled displacements
from the elliptical distribution of points in the Poincaré map.
Figure 13 Example bifurcation diagram. Stable behavior is observed up to an axial depth (b)
of approximately 2.5 mm. It is indicated by the single value of the sampled feed direction (x)
tool displacements. Between 2.5 mm and 4 mm, secondary Hopf chatter with added lobes
occurs (i.e., a distribution of sampled points). Between 4 mm and 5.5 mm, a period-3
bifurcation is seen (i.e., three distinct sampled points).
3.4 Chatter stability of multi-point tools with lateral, torsional, and axial flexibilities
Any flexible direction that affects the regenerative chip thickness must be considered in the
equation of motion with time delay. For example, twist drills have lateral ( x, y ) and torsional ( )
flexibilities that create the axial ( z ) vibrations which change the regenerative chip thickness as
shown in Figure 14. The dynamics of a twist drill with such flexibilities may be expressed as:
Mq(t ) Cq(t ) Kq(t ) F(q(t ), q(t T )) , (52)
where the vibration vector q(t ) x(t ) y (t ) z (t ) (t ) has lateral (𝑥, 𝑦), axial (𝑧) , and
torsional ( ) vibrations. The force vector is F(t ) Fx (t ) Fy (t ) Fz (t ) Tc (t ) where (Tc ) is
the cutting torque. The FRF in Eq. (44) becomes [62]:
xx xy 0 0
yy 0 0
Φ( ) yx , (53)
0 0 zz z
0 0 z
where is the torsional FRF excited by the drilling torque and z is the FRF contributed by the
torsional-axial coupling of vibrations. The eigenvalue problem becomes four-dimensional in Eq.
(44), but the frequency and semi-discrete time-domain solution methods remain the same. Boring
heads and plunge mills have the same dynamics as twist drills [63]. The same argument is valid
for workpieces where there may be cross-coupling of the structure in different directions. Milling
Figure 15 Rectilinear representation of variable pitch milling tool proposed by Slavicek [71].
Fj Kt a r re
i j
for j 1, N , (54)
where Fj is the dynamic cutting force on tooth j, r is the vibration amplitude, and φ𝑗 is the phase
shift or delay between the waves, i.e., the inner and outer modulations. The phase shift is different
for each cutting tooth due to the unequal distance between each successive tooth:
where c is the chatter frequency and vc is the cutting speed. This simple relation shows that the
tooth spacing and the cutting speed affect the delay, hence the tooth spacing l j can be selected to
minimize the delay φ𝑗 to increase the stability.
Opitz et al. [72] considered milling tool rotation using average directional factors in analyzing
stability with irregular tooth pitch. They also considered alternating pitch with only two different
pitch angles. Average directional factors which relate the dynamic chip thickness to the vibrations
do not accurately represent the tool rotation which causes time-varying milling dynamics. A
comprehensive study on the stability of milling tools with non-constant pitch was carried out by
Vanherck [73] using computer simulations. His study demonstrated that a certain pitch variation
pattern was effective, i.e., increased the stability limits significantly for a given milling system
over a certain cutting speed range. This fundamental idea was later used in several works for
optimal design of variable pitch tools.
Tlusty et al. [74] presented the stability of milling cutters with special geometries such as irregular
pitch or serrated edges using numerical simulations. Their numerical simulations showed the
effectiveness of irregular milling tool geometries and cutting edges in increasing stability against
chatter. Later, Altintas et al. [75] analyzed the stability of variable pitch cutters accurately using
their analytical milling stability model [24]. The analytical model considers tool rotation, time-
varying dynamics, and multiple vibration modes, hence their experimentally verified predictions
were more accurate. They considered both linear and alternating pitch variations and showed that
each variation pattern had an effective zone where chatter stability limits were increased
substantially compared to standard milling tools. Olgac et al. used the same dynamic model of
variable pitch milling operations, but proposed a parametric stability analysis [76]. These studies
mainly concentrated on the effect of pitch variation on the stability limit; they did not directly
address the cutting tool design to determine the optimal pitch variation, although they can be used
to see the effectiveness of various tool designs. Budak [77] proposed an optimization methodology
for the design of variable pitch tools considering the chatter frequency and spindle speed. The
spacing variation amount is related to the chatter waves left on the surface in order to establish the
linear pitch angle variation as p , p , p 2 ,.. where p is the base pitch angle and is
the pitch angle increment between the successive teeth. Budak [76] showed that the eigenvalue
solution in the analytical chatter stability model takes the following form for variable pitch tools:
N
a
Kt 1 e c j ,
i T
(56)
4 j 1
where T j is the tooth period which is variable due to non-constant pitch angles and a is the depth
of cut. Budak obtained the following simple equation for the critically stable depth of cut in the
case of variable pitch milling tools:
represents the summation of the phase delays caused by each tooth interval. This equation implied
that S must be minimized to maximize the chatter stability limit using variable pitch end mills. The
effect of the phase variation due to the variable pitch tool on the stability gain, which is defined as
the stability limit for the variable pitch tool over the one with the standard tool, was investigated
and it was shown that for particular values of the delay , the stability limit was maximized as
illustrated in Figure 16.
10
6
r
4
1/4 1 13/2
2
0
0 0.25 0.5 0.75 1
(rad) x2
Budak [77] showed that this condition could be achieved if the variation amount was selected as
follows:
for even N
c
(58)
( N 1)
for odd N
c N
where is the spindle speed in (rad/sec), is the pitch variation (rad), c is the chatter
frequency (rad/sec), and N is the number of teeth. Note that the additional phase delay introduced
by the optimal pitch variation is set as integer divisions of the original delay of the milling system
in a standard tool. The pitch variation should cancel the “remainder wave” for maximized stability.
Figure 18 Chatter stability of regular and serrated end mills, after Merdol et al. [81].
a1 z1 t z 2 t
F1 t
2 ,
Kr (59)
2
F t
a1 z2 t z1 t a2 a1 z2 t z 2 t
2
where Kr is the cutting force coefficient in the feed direction, z1 and z2 are the displacements of the
first and second tool in the feed direction, respectively; and τ is the workpiece rotation period. The
depth of cuts for the first and second tools are represented by a1 and a2, (a2 > a1), respectively
(Figure 20a). In this model, two regions on the shared surface can exist since the tools’ depths of
cut can be different. In the first region, both tools remove the same depth of cut a1. Therefore, the
displacement of each tool at any instant is influenced by the other tool’s displacement at the half
rotation period of the workpiece before. However, in the second region, the depth (a2-a1) is
removed by the second tool solely. Thus, the displacement of the second tool at time t is affected
by its displacement at the previous rotation period of the workpiece. The eigenvalue problem for
the marginally-stable case is formulated in the frequency domain and solved numerically. The
solution is provided for the first tool’s depth of cut over a range of spindle speed values while the
Figure 20 (a) parallel turning configuration, (b) stability lobe diagram for a1 versus spindle
speed (n) when a2= 1.5 mm and (c) time domain simulation for different points f (unstable), e
(stable) and d (unstable).(From [85])
Fx1 Fx1
F F
y1 ict 1 y1
e CPM DM OTF eict , (60)
Fx 2 4 Fx 2
Fy 2 Fy 2
where [CPM] is the cutting parameters matrix, [DM] is the delay matrix which includes the phase
delays for the milling tools, and [OTF] is the oriented transfer function matrix. Equation (60)
results in an eigenvalue problem that provides the experimentally proven stable depths for different
spindle speeds; see Figure 21.
7. Conclusion
This paper presents the fundamentals of chatter stability laws developed in the frequency and
discrete-time domains. It is shown that the dynamics of the machining process need to be modeled
by considering the interaction between the machining process and the structural dynamics of both
the machine and the machined part. The dynamics can be stationary with time-invariant
coefficients, such as in turning, or time-periodic with time-varying delays as in the case of turn-
milling operations. Although the understanding of stability for dynamic machining has increased
significantly during last six decades, the accuracy of stability predictions still suffers due to
measurement uncertainties, nonlinearities in the machine structure and process, time-varying
dynamics of machine tools and parts, and ploughing-based material removal by a worn tool or
chamfered cutting edge. The integration of physics-based off-line chatter prediction models needs
to be combined with on-line, adaptive learning and tuning techniques for robust chatter detection
and avoidance. Complex processes such as gear shaping with form tools, the stability of machining
composite metals, and additively manufactured parts with non-uniform material properties remain
to be studied.
Acknowledgments: The authors acknowledge the Natural Sciences and Engineering Council of
Canada Grants (IRCPJ 260683-18 and CANRIMT NETGP 479639-15); USA National Science
Foundation Grant (CMMI-1561221); TUBITAK-Turkey Grant Numbers: 105M032, 108M340,
110M522, 217M210; Hungarian National Research, Development and Innovation Fund
(TUDFO/51757/2019-ITM Thematic Excellence Program; and the Research Excellence Program
under Grant no. KKP133846.
References
[1] Taylor, F. W., 1906, "On the art of cutting metals," Transactions, The American Society of
Mechanical Engineers, 28, pp. 70-350.
[2] Arnold, R. N., 1946, "Cutting tools research: report of subcommittee on carbide tools: the
mechanism of tool vibration in the cutting of steel," P I Mech Eng B-J Eng, 154/1, pp. 261-284.
[3] Doi, S., and Kato, S., 1956, "Chatter vibration of lathe tools," Trans. ASME 78, p. 1127.
[4] Tobias, S., and Fishwick, W., 1958, "Theory of regenerative machine tool chatter," The
engineer, 205/7, pp. 199-203.
[5] Tlusty, J., and Polacek, M., 1963, "The stability of the machine tool against self excited
vibration in machining.," Proceedings of the ASME International Research in Production
Engineering.Pittsburgh, pp. 465-474.
[6] Tlusty, J., 1978, "Analysis of the state of research in cutting dynamics," Annals of the CIRP,
27/2, pp. 583-589.
[7] Tlusty, J., 1986, "Dynamics of High-Speed Milling," ASME Journal of Engineering for
Industry, 108(2), pp. 59-67.
[8] Altintas, Y., and Weck, M., 2004, "Chatter stability of metal cutting and grinding," Cirp
Annals-Manufacturing Technology, 53(2), pp. 619-642.
[9] Wiercigroch, M., and Budak, E., 2001, "Sources of nonlinearities, chatter generation and
suppression in metal cutting," Philos T R Soc A, 359(1781), pp. 663-693.