You are on page 1of 16

Mechanical Systems and Signal Processing 128 (2019) 266–281

Contents lists available at ScienceDirect

Mechanical Systems and Signal Processing


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

Model predictive control based active chatter control in milling


process
Denghui Li a,b, Hongrui Cao a,b,⇑, Xingwu Zhang a,b, Xuefeng Chen a,b, Ruqiang Yan a,b
a
School of Mechanical Engineering, Xi’an Jiaotong University, Xi’an 710049, PR China
b
State Key Laboratory for Manufacturing Systems Engineering, Xi’an Jiaotong University, Xi’an 710049, PR China

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

Article history: In high-speed milling process, chatter is often encountered due to the self-excitation mech-
Received 17 December 2018 anism, which not only limits the productivity, but also results in inferior work piece qual-
Received in revised form 28 March 2019 ity, shorter tool life, cutting noise and so on. In order to control chatter, an active chatter
Accepted 29 March 2019
control method based on model predictive control is proposed in this work. Firstly, an
Available online 6 April 2019
active control structure is designed by integrating the specific tool holder, piezoelectric
actuators and sensors into the spindle. Then, a time-varying model integrating the chatter
Keywords:
dynamics is developed comprehensively considering the designed structure and the
Chatter
Active control
milling process. The model is then transformed into a linear time invariant (LTI) approxi-
Model predictive control mate model through zero order approximation (ZOA) and Padé approximation. Next, on
Piezoelectric actuator the base of the model, a model predictive controller is designed to control only the chatter
dynamics caused by regenerative effect. Finally, both simulation analysis and milling
experiments are carried out to verify the effectiveness of the active chatter control method.
The results show that the chatter is effectively controlled in the milling process and the
milling stability is improved using the proposed active control method.
Ó 2019 Elsevier Ltd. All rights reserved.

1. Introduction

Chatter is a kind of self-excited vibration occurred in the milling process. It not only severely limits the improvement of
productivity but also affects the performance and the life of the machine components and even the whole system. With the
demand of efficient machining, chatter has become one of the main problems facing the machining industry. Various chatter
detection techniques [1–4] and chatter control methods have also been proposed to solve this problem.
The regenerative effect is reported as the main reason of chatter. So special tool geometries [5–8] and varying rotating
speed techniques [9] are proposed to control chatter by disrupting the regenerative effect. But compared with these passive
control methods, active control strategies that measure vibration response via sensors and execute control through actuators
are more flexible and effective, and thus have received extensive attention. In [10], Cao et al. summarized the state-of-the-art
of chatter monitoring and chatter control, and also proposed that active control is a promising method to improve the
machining stability.
In order to disrupt the regenerative effect, Denkena et al. [11] developed an adaptive spindle system with three piezo
actuators. In the machining process, the actuators generated superposed vibrations in the feed direction and disrupted
the regenerative effect. The cutting stability was improved but only within the speed range less than 5000 rpm. Sallese

⇑ Corresponding author at: Xi’an Jiaotong University, No. 28, Xianning West Road, Xi’an, Shaanxi 710049, PR China.
E-mail address: chr@mail.xjtu.edu.cn (H. Cao).

https://doi.org/10.1016/j.ymssp.2019.03.047
0888-3270/Ó 2019 Elsevier Ltd. All rights reserved.
D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281 267

et al. [12,13] designed an active fixture to suppress chatter vibrations by applying low-frequency excitations to the work
piece. The chatter vibration can be suppressed, but the excitation frequency depends on the tooth passing frequency which
varies with the change of cutting conditions.
Active methods which adjust system characteristics, such as damping and stiffness, through actuators are also often pro-
posed to control chatter. Using two pairs of piezoelectric actuators, Wang and Zhang et al. [14] proposed a milling dynamic
model of stiffness variation (SV) with two degrees of freedom and proved that the SV method was effective to suppress chat-
ter. Ries et al. [15] designed an active milling spindle. Two piezo-ceramic stack actuators were used to provide additional
damping to the spindle under the appropriate control strategy. Chen and Lu et al. [16,17] designed two types of controllers,
which were velocity feedback controller and loop shaping controller, to increase the damping and the stiffness of the boring
machine by applying a linear magnetic actuator to the boring bar. Kern et al. [18] designed a hybrid bearing spindle by instal-
ling an active magnetic bearing (AMB) between the tool and the front bearing of the spindle. The machining stability was
improved using robust feedback and adaptive feed forward control strategies. Dohner et al. [19] designed a controller based
on the linear quadratic Gaussian (LQG) method and increased the milling stability and the maximum metal removal rate
using two pairs of electrostrictive actuators. However, according to [20], this kind of controller lacks robustness and has dif-
ficulties when the machine’s dynamics change.
Parameter uncertainties in the models usually have a significant impact on the stability analysis and chatter control. Con-
sidering the parameter uncertainties, robust stability were analysed [21–24] and the results showed that the robust stability
diagrams were in better agreement with the experiment test. Thus, in order to suppress chatter effectively, robust tech-
niques are needed and also often used in active chatter control. Considering the uncertainties of the established model,
Van Dijk et al. [25–27] proposed a robust control method using l-synthesis. The chatter was effectively controlled in an a
priori specific milling condition with only the perturbation displacement as the feedback signal. On the basis of this study,
in order to reduce operations, such as chatter detection and perturbation signal separation, Wouw et al. [28] developed a
low-order Pyragas-type delayed feedback controller with the directly measured vibration displacement as input. Wu
et al. [29] also proposed a l-synthesis active chatter control approach to mitigate chatter in machining process with low
radial immersion. Building the perturbation model of the modal parameters and the milling parameters, Zhang et al. [30]
designed a robust controller and effectively suppressed chatter in the milling process. Based on H1 almost disturbance
decoupling problem, Shi et al. [31] presented a robust chatter control method in which the control effect could be adjusted
by changing the controller parameters. On the base of a spindle supported by AMBs, Huang et al. [32] developed an optimal
controller which was then enhanced by an adaptive algorithm to suppress chatter and expand the stability lobe diagram
boundary. By approximating the milling force with Fourier series and extending the system state with discrete strategy,
Zhang et al. [33] proposed a model predictive control method to improve the milling stability.
Although many active chatter control methods have been proposed, there are still some problems which limit their appli-
cation in industry. For example, there is a certain deviation between the established model and the actual system; the input
limit of the actuator has not been effectively dealt. In this paper, a model predictive control (MPC) based active chatter con-
trol method is proposed and some work is done on these problems. In detail, the frequency response functions (FRFs)
between different positions (tool tip, sensors and actuators) are firstly measured and the corresponding subsystems are then
identified. Based on the identified subsystem models, the accurate model of the designed active chatter control spindle sys-
tem is developed. In order to reduce the model uncertainty caused by chatter, the milling force is divided into two parts (sta-
tic milling force due to feed and dynamic milling force due to chatter) based on their generation mechanism and the dynamic
milling force, which is also the chatter dynamics, is integrated into the system model. Thus, the dynamic characteristics of
chatter can be considered in real time and the accuracy of the spindle-actuator-milling process system model is improved. Of
course, the strong robustness of the designed MPC controller can also help to deal with model uncertainty. According to the
characteristics of the piezoelectric actuator, its input constraint, which is also the output limit of the controller, can be deter-
mined and also set explicitly in the design process of the MPC controller. So the output command will be executed accurately
by the actuator in the real control process and the accuracy of the control can be guaranteed. What’s more, the static milling
force has no effect on chatter. So it is reasonably ignored in the design of the controller. In this way, only the chatter dynam-
ics is controlled, and the stable vibration caused by static milling force is hardly concerned. It is helpful for energy saving and
is also benefit to maximize system stability using the limited power of the actuator.
The remainder of this paper is organized as follows. Section 2 describes the structure of the designed active chatter con-
trol system and also develops a LTI model by zero order approximation and Padé approximation. Based on the established
model, a model predictive controller is designed in Section 3. In Section 4, a numerical simulation is carried out using the
experimentally identified system parameters. The milling experimental verification is presented in Section 5. Finally, conclu-
sions and discussions are drawn in Section 6.

2. System modeling

2.1. Design of active control structure

In order to control chatter, an active control structure is firstly designed. As shown in Fig. 1, a specific tool holder inte-
grated with a pair of bearings is customized, wherein the bearings ensure that the control forces of the actuators can be effec-
268 D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281

Piezoelectric Connecting Front Spindle Rear


actuator structure bearing housing Rotor bearing

Tool holder

Milling tool

Bearing

Ring

Displacement
sensor
Controller Amplifier

Fig. 1. Structure of the active control system.

tively applied to the rotating tool holder. In the milling process, the vibration displacements of the tool holder can be mea-
sured by a pair of displacement sensors in the orthogonal X and Y directions and then passed to the controller. Based on the
measured displacement response and the control objectives, the controller calculates appropriate voltages to drive the two
pairs of piezoelectric actuators, which are orthogonally distributed in X and Y directions, to exert active control forces to the
tool holder through the bearings.
In Fig. 1, three symbols are used to mark the key locations, where P t represents the tool tip position, P a represents actu-
ator excitation position, and Pd represents the measurement point of the displacement sensor.

2.2. Modeling of the spindle-actuator-milling process system

According to Fig. 1, the block diagram of the spindle-actuator-milling process system is shown in Fig. 2. In the block dia-
gram, Ft is the milling force; Fa is the control force; zt is the vibration displacement at point P t ; zd is the vibration displace-
ment at point Pd ; CðsÞ is the designed controller; Gij ðsÞ is the transfer function matrix from point Pj to point P i with i ¼ t; d and
j ¼ t; a. Then, the dynamic equation of the system can be described as Eq. (1).
    
zt ðsÞ Gtt ðsÞ Gta ðsÞ Ft ðsÞ
¼ ð1Þ
zd ðsÞ Gdt ðsÞ Gda ðsÞ Fa ðsÞ
 T  T
where zt ðsÞ ¼ ztx ðsÞ; zty ðsÞ ; zd ðsÞ ¼ zdx ðsÞ; zdy ðsÞ ; Ft ðsÞ ¼ ½F tx ðsÞ; F ty ðsÞT ; Fa ðsÞ ¼ ½F ax ðsÞ; F ay ðsÞT ; Gtt ðsÞ ¼ diagðGttx ðsÞ; Gtty ðsÞÞ;
Gta ðsÞ ¼ diagðGtax ðsÞ; Gtay ðsÞÞ; Gdt ðsÞ ¼ diagðGdtx ðsÞ; Gdty ðsÞÞ; Gda ðsÞ ¼ diagðGdax ðsÞ; Gday ðsÞÞ The subscripts ‘‘x” and ‘‘y” indicate
the X direction and the Y direction of the spindle system, respectively. Gijx and Gijy are the transfer functions from point
Pj to point P i in X direction and Y direction, respectively. In this work, the crossover effect between X and Y directions is
not considered. Compared with the model developed in [33], the actual application position of the control forces and the
transmission relationship from the position to the tool tip are considered in more detail in this work.
In order to facilitate subsequent calculations and analysis, the state space equation Eq. (2) of the spindle-actuator-milling
process system can be obtained based on Eq. (1).

Fig. 2. The block diagram of the spindle-actuator-milling process system.


D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281 269

_
 
xðtÞ ¼ A xðtÞ þ BFðtÞ

ð2Þ
zðtÞ ¼ C xðtÞ þ DFðtÞ

where zðtÞ ¼ ½zt ðtÞ; zd ðtÞT is the output; FðtÞ ¼ ½Ft ðtÞ; Fa ðtÞT is the input; x ðtÞ8q1 is the state vector and the value of q is equal
to the dimension of the subsystem identified by the transfer function Gijk with k ¼ x; y; A8q8q ; B8q4 ; C48q and D44 ¼ 0 are
the coefficient matrixes.
In order to separate the milling force Ft ðtÞ ¼ ½F tx ðtÞ; F ty ðtÞT and the active control force Fa ðtÞ ¼ ½F ax ðtÞ; F ay ðtÞT , Eq. (2) is
converted to Eq. (3) as below.
_
 
xðtÞ ¼ A xðtÞ þ ½Bt ; Ba ½Ft ðtÞ; Fa ðtÞT

ð3Þ
½zt ðtÞ; zd ðtÞT ¼ ½Ct ; Ca T xðtÞ

where ½Bt ; Ba  ¼ B and the dimensions of ½Bt 8q2 and ½Ba 8q2 are the same; ½Ct ; Ca T ¼ C and the dimensions of ½Ct 28q and
½Ca 28q are the same, too.

2.3. Zero order approximation for the time-varying system

Regenerative chatter is caused by the regenerative effect. So, in the milling process with chatter, the milling force depends
not only on the feed rate but also on the vibration response of the tool in two adjacent cuttings. Thus, the milling force can be
described as below.
1 1
Ft ðtÞ ¼ beK t Hs ðtÞ þ bK t HðtÞðzt ðtÞ  zt ðt  sÞÞ ð4Þ
2 2
where b is the axial depth of cut; e is the feed per tooth; K t is the tangential milling force coefficient and can be obtained by
experiment. The time delay s can be calculated by s ¼ 60=ðN XÞ where N is the tooth number of the milling tool and X is the
rotating speed of the spindle in r/min.
The dynamic milling force coefficient matrix HðtÞ is shown as Eq. (5), and the static milling force coefficient matrix Hs ðtÞ is
equal to the first column of HðtÞ.
" PN1 PN1 #
j¼0  ½sinð2/j Þ þ K rc ð1  cosð2/j ÞÞ j¼0  ½1 þ cosð2/j Þ þ K rc sinð2/j Þ
HðtÞ ¼ PN1 PN1 gð/j Þ ð5Þ
j¼0 ½1  cosð2/j Þ  K rc sinð2/j Þ j¼0 ½sinð2/j Þ  K rc ð1 þ cosð2/j ÞÞ

where K rc ¼ K r =K t and K r is the radial milling force coefficient; /j is the instantaneous angular position of the tooth j. gð/j Þ is
a switch function and it can be described as
(
1; /en < /j < /ex
gð/j Þ ¼ ð6Þ
0; /j < /en or /j > /ex

where /en is the entry angle of the tool while /ex is the exit angle. In up-milling process, /en ¼ 0 and /ex ¼ arccosð1  2d=DÞ,
while in down-milling process, /en ¼ arccosð2d=D  1Þ and /ex ¼ p, in which D is the diameter of the tool and d is the radial
depth of cut.
The static and dynamic milling force coefficient matrixes Hs ðtÞ and HðtÞ are both periodic at tooth passing period s. There-
fore, they can be expressed as Fourier series [34]. In order to simplify the model while ensuring accuracy, zero order approx-
imation is applied and the periodic time-varying matrixes Hs ðtÞ and HðtÞ are approximated by constant matrixes Hs and H,
respectively. H is shown as Eq. (7).
 
hxx hxy
H¼ ð7Þ
hyx hyy
where hxx; hxy; hyx and hyy can be written as
/
hxx ¼ 4Np ½cosð2/Þ  2K rc / þ K rc sinð2/Þ/exen
/
hxy ¼ 4Np ½sinð2/Þ  2/ þ K rc cosð2/Þ/exen
/
ð8Þ
hyx ¼ 4Np ½sinð2/Þ þ 2/ þ K rc cosð2/Þ/exen
/
hyy ¼ 4Np ½cosð2/Þ  2K rc /  K rc sinð2/Þ/exen
T
Hs is equal to the first column of H and can be written as Hs ¼ ½hxx; hyx .
By substituting Eqs. (4) and (7) into Eq. (3), a time-invariant system is obtained as below.
270 D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281

_
  
xðtÞ ¼ ðA þ hÞ xðtÞ  h xðt  sÞ þ Bt Ftn ðtÞ þ Ba Fa ðtÞ

ð9Þ
½zt ðtÞ; zd ðtÞT ¼ ½Ct ; Ca T xðtÞ

where h ¼ bK t Bt HCt =2; Ftn ðtÞ ¼ beK t Hs =2 is the static milling force, which is also called nominal milling force.

2.4. Padé approximation for the time delay expression

 
Defining xT ðtÞ ¼ xðt  sÞ, then its Laplace transform is
 
xT ðsÞ ¼ diagðess ; ess ;    ; ess Þ xðsÞ ð10Þ
ss
The lag factor e is a transcendental function and it can be approximated by a polynomial as Eq. (11) through Padé
approximation.

a0  a1 s þ    þ ð1ÞL aL sL
ess  pd ðsÞ ¼ ð11Þ
a0 þ a1 s þ    þ aL sL
where s is the Laplace operator; a0 ; a1 ;    aL are the constants; L is the order of the Padé approximation.
For ease of calculation, the state space equation is used and Eq. (12) can be obtained based on Eqs. (10) and (11).

x_ d ðtÞ ¼ Ad xd ðtÞ þ Bd xj ðtÞ
  ð12Þ
xT;j ðtÞ ¼ Cd xd ðtÞ þ Dd xj ðtÞ
   
where xj ðtÞ and xT;j ðtÞ are the jth (j 2 N þ ) elements in xðtÞ and xT ðtÞ, respectively; xd ðtÞ is the state vector and its dimension is
equal to L.
  
By substituting Eqs. (11) and (12) into Eq. (10), xT ðtÞ which is also xðt  sÞ can be calculated by a LTI system with xðtÞ as
the input.

x_ de ðtÞ ¼ Ade xde ðtÞ þ Bde xðtÞ
  ð13Þ
xT ðtÞ ¼ Cde xde ðtÞ þ Dde xðtÞ
 
where xT ðtÞ 2 Rm ; xðtÞ 2 Rm ; xde ðtÞ 2 RmL ; Ade ¼ diagðAd ; Ad ;    ; Ad Þ; Bde ¼ diagðBd ; Bd ;    ; Bd Þ; Cde ¼ diagðCd ; Cd ;    ; Cd Þ; Dde ¼
diagðDd ; Dd ;    ; Dd Þ. The value of L is related to the milling parameters and the frequency range concerned. Compared with
extension of system state using discrete strategy [33], a system equation with smaller dimensions is easier to be obtained
using Padé approximation, which is conducive to reducing the computational cost in the control process and improving
the real-time performance of control.
So far, substitute Eq. (13) into Eq. (9) and a LTI state expanded system without time delay expression can be obtained.
_
XðtÞ ¼ Az XðtÞ þ Bz UðtÞ
ð14Þ
ZðtÞ ¼ Cz XðtÞ
 T
where XðtÞ ¼ ½xðtÞ; xde ðtÞ ; ZðtÞ ¼ ½zt ðtÞ; zd ðtÞT ; UðtÞ ¼ ½Ftn ðtÞ; Fa ðtÞT . It can be seen that the input UðtÞ of the system here is
different from that (½Ft ðtÞ; Fa ðtÞT ) in Eq. (3). The chatter dynamics caused by regenerative effect has been integrated into the
controlled system model. Az ; Bz and Cz can be written as
 
A þ h  hDde hCde
Az ¼
Bde Ade
   ; ð15Þ
Bt Ba Ct 0t2
Bz ¼ ; Cz ¼
0t1 0a1 Ca 0a2

where 0t1 ; 0a1 ; 0t2 and 0a2 are all zero matrixes but with different dimensions.

3. Model predictive controller

3.1. Model predictive control

Model predictive control is a kind of model-based control method with three control strategies: multi-step prediction,
receding horizon optimization and feedback compensation [35]. The mechanism of MPC is shown in Fig. 3. At each sampling
time, a finite time domain optimization problem is solved online according to the current measurement information, and the
first element of the control sequence is applied to the controlled object. At the next sampling moment, the above process is
repeated: refresh the optimization problem with a new feedback value and resolve it.
D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281 271

Fig. 3. Mechanism of model predictive control.

In the process of model predictive control, the predictive model is firstly developed based on the control object. The sys-
tem’s reference output Zref ðk þ jjkÞ; which means the output of the system at time k þ j predicted at time k, is also set. The
purpose of model predictive control is to make the output of the system change according to the set reference output.
At each time k, the output of the system in the next N p step is predicted. The control sequence in the next N c (N c  N p ) step
can be calculated based on the cost function Eq. (16) (where Q and R are output error weight matrix and control weight
matrix, respectively) as well as the constraints Eq. (17). It is noted that the control force does not change after the control
horizon ends, which can be described as Eq. (18).

X
Np
T
X
N c 1
T
J¼ ðZðk þ jjkÞ  Zref ðk þ jjkÞÞ Q ðZðk þ jjkÞ  Zref ðk þ jjkÞÞ þ Fa ðk þ jjkÞ RFa ðk þ jjkÞ ð16Þ
j¼1 j¼0

Famin  Fa ðk þ jjkÞ  Famax


ð17Þ
DFamin  DFa ðk þ jjkÞ  DFamax

Fa ðk þ jjkÞ ¼ Fa ðk þ Nc  1jkÞ; j > Nc ð18Þ


At time k, the control sequence ½Fa ðkjkÞ;    ; Fa ðk þ N c jkÞ; Fa ðk þ N c þ 1jkÞ;    ; Fa ðk þ N p  1jkÞ is obtained. However, only
the first element Fa ðkjkÞ is applied to the controlled object. At time k þ 1, new measured output is fed back to the controller
and the above operation is repeated.

3.2. Design of model predictive controller

3.2.1. The development of the predictive model


On the basis of the LTI system developed in Section 2, a discrete model of the spindle-actuator-milling system is estab-
lished as below.
Xðk þ 1Þ ¼ wXðkÞ þ c1 Ftn ðkÞ þ c2 Fa ðkÞ
ð19Þ
ZðkÞ ¼ Cz XðkÞ
where w ¼ wðTÞ, c1 ¼ c1 ðTÞ and c2 ¼ c2 ðTÞ where T is the discrete period and is a constant; for conciseness, the variable ðkTÞ
is abbreviated as ðkÞ in the text, for example XðkTÞ ¼ XðkÞ; Ftn ðkÞ has no effect on the occurrence of chatter and is therefore
regarded as external interference. w and c ¼ ½c1 ; c2  can be calculated by Eq. (20).

w ¼ eAz T
R ð20Þ
c ¼ 0T eAz t Bz dt
Eq. (19) reflects the change of the vibration displacement ZðkÞ with the active control force Fa ðkÞ under the interference
Ftn ðkÞ. However, in the process of model predictive control, it is more common for DFa ðkÞ ¼ Fa ðkÞ  Fa ðk  1Þ, the change of
the active force, to be punished in the cost function. It is also more convenient and effective to control DFa ðkÞ in a discrete
system, because the integral nature from DFa ðkÞ to Fa ðkÞ enables the dynamic system to achieve error-free control. Hence, an
T
augmented state space model is developed as below by defining state vector as fðkÞ ¼ ½XðkÞ; Fa ðk  1Þ .
        
Xðk þ 1Þ w c2 XðkÞ c2 c
¼ þ DFa ðkÞ þ 1 Ftn ðkÞ
Fa ðkÞ 0 I Fa ðk  1Þ I I
  ð21Þ
XðkÞ
ZðkÞ ¼ ½ Cz 0
Fa ðk  1Þ
272 D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281

where I 2 R22 is an unit vector.


In the milling process, the static milling force Ftn cannot be measured directly, which means that it is unknown in the
process of controller design. In the work, the goal is to control only chatter dynamics and to maximize the stability of the
milling process with limited control forces. As mentioned above, Ftn has no effect on chatter and the vibration of the system
under Ftn is stable. So, in order to avoid the increase of the control cost due to Ftn , the stable vibration caused by Ftn is not
considered in the control process. Thus, the interference Ftn can be set to a zero vector in the process of controller design.
Therefore, it can be ignored in the prediction process and the model can be simplified to Eq. (22).
      
Xðk þ 1Þ w c2 XðkÞ c2
¼ þ DFa ðkÞ
Fa ðkÞ 0 I Fa ðk  1Þ I
  ð22Þ
XðkÞ
ZðkÞ ¼ ½ Cz 0
Fa ðk  1Þ
According to Eq. (22), the predicted state Xðk þ jjkÞ; j  N c can be described as Eq. (23).
2 3
DFa ðkjkÞ
Pj1 6 .. 7 Xj1
Xðk þ jjkÞ ¼ W j XðkÞ þ ½ i¼0 W
i
c2    c2 6
4 .

5 Wi c2 Fa ðk  1Þ ð23Þ
i¼0
DFa ðk þ j  1jkÞ
When N c < j  N p , Eq. (18) is applied and the predicted state Xðk þ jjkÞ can be obtained as Eq. (24).
2 3
DFa ðkjkÞ
Pj1 PjNc i 6 .. 7 Xj1
Xðk þ jjkÞ ¼ W j XðkÞ þ ½ i¼0 W
i
c2  6
i¼0 W c2 4 .

5 Wi c2 Fa ðk  1Þ ð24Þ
i¼0
DFa ðk þ Nc  1jkÞ
Summing up the above, the predictive model can be written as below.
2 3 2 3 2 3 2 3
Xðk þ 1jkÞ W c2 c2   0
6 7 6 7 6 7
6 .. 7 6 .. 7 6 .. 7 6 .. .. .. 72
6 . 7 6 6 . 7 7 6 . 7 6 . . . 7 DFa ðkjkÞ
3
6 7 6 PNc 1 i 7 6 PNc 1 i 7
6 Xðk þ Nc jkÞ 7 6 Nc 7 6 7 6 7
6 7 6 W 7 6 i¼0 W c2 7 6 i¼0 W c2   c2 76 .. 7
6 7 ¼ 6 Nc þ1 7 7 XðkÞ þ 6 PNc i 7 Fa ðk  1Þ þ 6 PNc i 76 .
7
6 Xðk þ Nc þ 1jkÞ 7 6 6 c 7 6 c   c þ c 7 4 5
6 7 6 6
W 7
7 6 i¼0 W 2 7 6 i¼0 W 2 W 2 2 7
6 .. 7 6 .. 7 6 .. 7 6 .. .. .. 7 DFa ðk þ Nc  1jkÞ
6 . 7 4 . 5 6 7 6 7 |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
4 5 4 . 5 4 . . . 5
  Np PNp 1 i PNp 1 i PNp Nc i DFaz ðkÞ
X k þ Np jk W
|fflfflfflfflfflffl{zfflfflfflfflfflffl} W c W c   W c
|fflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflffl}
i¼0 2
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
i¼0 2 i¼0 2
Ap Bp1 Bp2

ð25Þ

2 3 2 32 3
Zðk þ 1jkÞ Cz Xðk þ 1jkÞ
6 Zðk þ 2jkÞ 7 6 7 6 7
6 7 6 Cz 7 6 Xðk þ 2jkÞ 7
6 .. 7 ¼ 6 .. 7 6 .. 7
6 7 6 76 7 ð26Þ
4 . 5 4 . 54 . 5
   
Z k þ N p jk Cz X k þ Np jk
|fflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Zz ðkÞ Cp

According to Eqs. (25) and (26), Eq. (27) can be obtained.

Zz ðkÞ ¼ UXðkÞ þ CFa ðk  1Þ þ KDFaz ðkÞ ð27Þ

where U ¼ Cp Ap ; C ¼ Cp Bp1 ; K ¼ Cp Bp2 .

3.2.2. The solving of the controller


In normal milling process, the vibration displacement of the tool stably fluctuates within low amplitude. When chatter
occurs, the milling force sharply increases and the vibration displacement at tool tip also increases significantly. As a result,
the milling process becomes unstable. Hence, the reduction of the vibration displacement caused by chatter is the target of
chatter control and is also the inevitable result of chatter control. In order to achieve better chatter control effect, the refer-
ence output of the predictive model needs to be as small as possible. For quantitative analysis, the reference output is set to a
zero vector and it can be described as Zref ðk þ jjkÞ 0. As stated above, the predictive model does not consider static milling
forces, so the control objective of this study is to reduce the system vibration caused by chatter to zero. It means that under
control, chatter is suppressed and the system stably vibrates under static milling forces.
D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281 273

Cost function is an important basis for calculating active control forces. Substitute Zref ðk þ jjkÞ 0 into Eq. (16), the cost
function can be obtained as below.

X
Np
T
X
N c 1
T
J¼ Zðk þ jjkÞ QZðk þ jjkÞ þ DFa ðk þ jjkÞ RDFa ðk þ jjkÞ ð28Þ
j¼1 j¼0

where Q 2 R44 and R 2 R22 are constant matrixes, and N p and N c are constants. All of their values have great influence on
the control effect and they can be set based on experience and the characteristics of the system model.
According to Eqs. (25)–(27), Eq. (28) can be also described as
T T
J ¼ Zz ðkÞ Q z Zz ðkÞ þ DFaz ðkÞ Rz DFaz ðkÞ ð29Þ

where Q z ¼ diagð Q    Q Þ; Rz ¼ diagð R    R Þ.


Define EðkÞ ¼ UXðkÞ  CFa ðk  1Þ, then Zz ðkÞ ¼ KDFaz ðkÞ  EðkÞ and Eq. (29) can be written as
1 z T
J¼ DF ðkÞ bDFaz ðkÞ þ dT DFaz ðkÞ þ const ð30Þ
2 a
where b ¼ 2ðKT Q z K þ Rz Þ; d ¼ 2KT Q z EðkÞ.
The power of the piezoelectric actuator is limited. So the constraints of the output force of the actuator driven by the con-
troller must be considered. It means that Eq. (17) needs to be satisfied in the control process.
It is clear that Faz ðkÞ and DFaz ðkÞ satisfy the relationship shown in Eq. (31). Then the constraints Eq. (17) can be described as
Eq. (32).
2 3 2 32 3 2 3
Fa ðkjkÞ I 0  0 DFa ðkjkÞ I
6 F ðk þ 1jkÞ 7 6 I I    0 7 6 DF ðk þ 1jkÞ 7 6 I 7
6 a 7 6 76 a 7 6 7
6 7 ¼ 6. . . 76 7 þ 6 . 7 Fa ðk  1Þ
6
4
.. 7 6. .
5 4. . . . ... 7 54
6 .. 7 6.7
5 4.5 ð31Þ
. .
Fa ðk þ Nc  1jkÞ I I  I D Fa ðk þ Nc  1jkÞ I
|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |ffl{zffl}
g k

2 3 2 3
I ^amax
DF
6 I 7 6 7
^amin
6 7 z 6 DF 7
6 7DFa ðkÞ  6
6 F
^
7
7 ð32Þ
4 g 5 4 amax  kFa ðk  1Þ 5
g ^
Famin þ kFa ðk  1Þ

where F^amin ¼ ½ Famin    Famin T ; F


^amax ¼ ½ Famax    Famax T ; Famin and Famax are the minimum and maximum output forces
which are proportional to the lower and upper limit input voltages of the piezoelectric actuators.
So far, the optimal or sub-optimal control sequence DFaz ðkÞ at time k can be calculated by solving the quadratic program-
ming (QP) standard problem as below.
1 z T
min DF ðkÞ bDFaz ðkÞ þ dT DFaz ðkÞ s:t: Eq: ð32Þ ð33Þ
DFaz ðkÞ 2 a

Among the control sequence, only the first element DFa ðkÞ ¼ ½ I 0    0 DFaz ðkÞ is used and only
Fa ðkÞ ¼ DFa ðkÞ þ Fa ðk  1Þ is output and executed by the actuators based on receding horizon optimization.
At time k þ 1, the optimization problem is refreshed by the real-time feedback of the vibration displacements measured
in the milling process and the above process is repeated. Through these operations (receding horizon optimization and feed-
back compensation), the uncertainty of the developed model can be considered and a strong robustness of the controller is
guaranteed.
The proposed MPC method of active chatter control is shown in Fig. 4. The real-time control process can be described as
below.

(1) According to the characteristics of the spindle-actuator-milling system, the predictive model Eq. (27) is developed.
(2) At time k, the vibration displacements are measured and fed back to the controller. The reference output
Zref ðk þ jÞ 0; j ¼ 1; 2;    ; N p is also set.
(3) The cost function J is obtained and the constraints for the controller output are also set.
(4) Calculate the optimal or sub-optimal control increment sequence DFaz ðkÞ by solving the QP standard problem.
(5) Based on receding horizon optimization, only the first element DFa ðkÞ is preserved and the control force
Fa ðkÞ ¼ DFa ðkÞ þ Fa ðk  1Þ is applied to the spindle-actuator-milling system.
(6) At time k þ 1, repeat the above control process with a refreshed optimization problem.
274 D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281

Fig. 4. The proposed MPC method of active chatter control.

4. Control simulation

4.1. Milling force coefficient identification

According to the modeling method proposed in Section 2, the radial and tangential milling force coefficients K r and K t
need to be identified to establish the system model.
As seen in Fig. 5, milling experiments are carried out and the milling forces are measured by the Kistler9443B dynamome-
ter. The parameters of the milling tool and the milling process are listed in Tables 1 and 2, respectively. With sampling fre-
quency 10,000 Hz, the milling forces under different feeds per tooth are measured and recorded. Then the coefficients K r and
K t are calculated based on the method mentioned in [34]. The results are shown in Table 3.

4.2. Modal parameter identification

As shown in Fig. 6, the impact experiment and swept frequency experiment are carried out to identify the model param-
eters of the spindle-actuator system. The milling tool is the same as that in Table 1, but the overhang is 75 mm. The test

Milling tool

Workpiece

Kistler
dynamometer

Fig. 5. The setup for milling force measurement.


D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281 275

Table 1
Parameters of the milling tool.

Material Tooth number Diameter (mm) Overhang (mm)


High speed steel 3 10 60

Table 2
Parameters of the milling process.

Material of the work piece Radial depth of cut (mm) Axial depth of cut (mm) Spindle speed (r/min)
Aluminum alloy 7075 5 0.5 8000

Table 3
Identified milling force coefficients.

Milling force coefficient Radial (MPa) Tangential (MPa)


Value 466 1432

(a) (b)
Displacement
sensor

Piezoelectric
actuators

Accelerometer Hammer Accelerometer

Fig. 6. (a) Impact experiment (b) Swept frequency experiment.

equipment and their technical parameters are listed in Tables 4 and 5. The excitation signal (force for impact experiment or
voltage signal for swept frequency experiment) and the response signal (acceleration at point P t and displacement at point
P d ) are measured and recorded at the same time by the ECON data acquisition system with sampling frequency 10240 Hz.
The impact experiment is firstly implemented by hammering the tool tip. By impacting the tool tip in the X direction and
measuring the responses at point P t and P d in this direction at the same time, the FRFs from P t to Pt and Pd are obtained in x-
axis. The FRFs from P t to Pt and P d in y-axis are also measured by a same operation in the Y direction. Then the swept fre-
quency experiments are carried out. A sine sweep voltage signal with range from 20 Hz to 3000 Hz is applied to the piezo-
electric actuator in the X direction and the responses at point P t and Pd are both measured in this direction. Thus, the FRFs

Table 4
Test equipment and their technical parameters.

Test equipment Brand Parameter values


Impact hammer PCB 086C01 Sensitivity 2.25 mv/N
Accelerometer DYTRAN 3032A Sensitivity 10.08 mv/g
Displacement sensor Micro-Epsilon CS02 Sensitivity 50000 mv/mm
276 D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281

Table 5
Parameters of the piezoelectric actuator.

Model Nominal Stroke (lm) Stiffness (N/lm) Nominal thrust (N) Resonant frequency (kHz)
Pst80VS15 76 25 2300 12

from P a to Pt and P d in x-axis are obtained. The FRFs from Pa to P t and P d in y-axis are also measured by doing a same oper-
ation in the Y direction. Since the output force of the actuator is proportional to the input voltage, the actuator and the spin-
dle are considered as a whole, and the input of the FRFs is the voltage in the swept frequency experiment. In this work, the
cross FRFs between X-direction and Y-direction are not considered. Take X-direction for example, the measured FRFs are
shown in Fig. 7. The first two natural frequencies are around 1094 Hz and 1516 Hz. The identified model parameters are
shown in Table 6. The model outputs of the corresponding systems identified are also presented in Fig. 7. It can be seen that
the characteristics of the identified system are in good agreement with that of the measured system. Thus, the values of
A; Bt ; Ba ; Ct ; Ca and h horizon ends, which can be described in the system equation developed in Section 2 can be obtained
from the identified results.

(a) (b)

(c) (d)

Fig. 7. Comparison of FRFs between experiment data and identified model in X-direction (a) from P t to P t (b) from P t to P d (c) from P a to P t (d) from P a to P d .

Table 6
Identified model parameters.

FRFs Order Mass (kg) Stiffness (N/m) Damping ratio


From P t to P t 1st 0.0835 3.9465e6 0.0220
2nd 0.0289 2.6196e6 0.0227
From P t to P d 1st 0.9654 4.6169e7 0.0187
2nd 0.5000 4.5159e7 0.0236
From P a to P t 1st 0.1228 5.8539e6 0.0205
2nd 0.1116 1.0066e7 0.0236
From P a to P d 1st 1.0664 5.0711e7 0.0205
2nd 2.0111 1.8283e8 0.0218
D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281 277

4.3. Selection of the milling parameters

According to the milling force coefficients and model parameters identified in Sections 4.1 and 4.2, the stable lobe
diagram (SLD) of the milling process can be plotted. In up-milling process, when the radial depth of cut is 2 mm, the SLD
calculated by semi discretization method, full discretization method and ZOA method are all described in Fig. 8. It can be
seen that the SLD obtained by ZOA method agrees well with those of the other two methods, which means that ZOA used
in Section 2 is reasonable and effective. It is also clear that the chatter-free axial depth of cut at the speed of 8000 r/min is
near the lowest value of the whole speed range. So the simulation analysis and the control experiment are both carried out
on this machining condition.

4.4. The order of Padé approximation

With the increase of the order of Padé approximation, the developed LTI system can approximate the original time delay
system more accurately. However, the model predictive controller will be more complex when the approximate order is
higher. So the choice of the order of Padé approximation is important. Since the amplitude of Padé approximation is always
one, it is enough to only analyze the phase-frequency characteristics of the approximate system with different orders. In
Fig. 9, the phase angle curves of different approximate orders are described with s ¼ 0:0025 s which is the tooth passing
period of a milling tool with three teeth at the speed of 8000 r/min. As we can see from Section 4.2, the two main modal
frequencies of the system are around 1094 Hz and 1516 Hz. It is noted that the chatter frequency in the milling process
locates near the natural frequency of the spindle-milling system [26]. So only the fitting accuracy of the low frequency band
less than 1600 Hz of the approximate system is concerned. Hence, the order of Padé approximation is set to 14 based on
Fig. 9. As a result, Ade ; Bde and Cde in the system equation in Section 2 can be obtained.

Fig. 8. SLD of the up-milling process.

Fig. 9. Phase angle curves of the approximate system with different orders.
278 D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281

4.5. Simulation analysis

Based on the results of above Sections, the augmented state space model in Section 3 can be developed and the simulation
analysis can be implemented.
In addition to the control model, the parameters such as N p ; N c ; Q ; R and so on also need to be set firstly to develop the
controller. These parameters can be determined by prior knowledge as well as the response characteristics of the system in
the milling process. For example, the controller shows better robustness but poor real-time performance with larger N p ,
while better rapidity but worse stability when N p is smaller. In this work, N p ¼ 15 and N c ¼ 10 are set. As for Q and R, they
are adjusted based on the amplitude of the displacement response. Considering the limits of the piezoelectric actuator and
the amplification factor of the power amplifier, the output of the controller is limited between 10 V to 10 V and the change
of output is in the range 20 V to 20 V. When the output is less than 0 V, it means that a positive voltage equal to the abso-
lute value of the output is input to the opposite actuator.
Based on the SLD developed above, an up-milling process is analyzed. The milling parameters are listed in Table 7. The
milling tool and the work piece material are the same as those in Tables 1 and 2, but the overhang of the tool is 75 mm. Take
X-direction for example, the results without control and under control are shown in Fig. 10. It can be seen that without con-
trol, the displacement at tool tip and the displacement measured by the displacement sensor are both significantly larger
than those under control. Except for the tooth-passing frequency and its harmonics, there are two main chatter frequencies,
which are f c1 ¼ 1151 Hz and f c2 ¼ 1552 Hz, in the frequency spectrums of the vibration displacements without control. With
control on, the chatter frequencies disappear, and only the tooth-passing frequency and its harmonics are left, which means
the system stably vibrates under static milling forces. Therefore, it can be concluded that chatter under the machining con-
dition listed in Table 7 is effectively controlled by the controller designed in this work. The control voltage calculated by the
controller in the control process is shown in Fig. 11. It is clear that the value of the control voltage is less than 10 V, which
indicates that the control process is feasible.

5. Experiment verification

In order to verify the method proposed in this work, two milling operations with and without control are carried out on
the three-axis milling machine (VMC-V5) based on above identified results and simulation analysis.

Table 7
Parameters of the milling process in simulation.

Spindle speed (r/min) Radial depth of cut (mm) Axial depth of cut (mm) Feed per tooth (mm)
8000 2 1.2 0.03

(a)

(b)

Fig. 10. (a) Displacement at point pt and its frequency spectrum (b) Displacement at point pd and its frequency spectrum. ‘‘fc1” and ‘‘fc2” represent chatter
frequencies.
D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281 279

(a)

(b)

Fig. 11. (a) Control voltage for the actuator in +X direction (b) Control voltage for the actuator in X direction.

Vibration
displacement

PC dSPACE

Amplifier
PC
ECON
Control
voltage
Milling process Control unit Monitoring and recording unit
Fig. 12. Control experiment set up.

(a)

(b)

Fig. 13. (a) Vibration displacement at P d in X-direction and its frequency spectrum (b) Vibration displacement at P d in Y-direction and its frequency
spectrum. ‘‘fc1” and ‘‘fc2” represent chatter frequencies; ‘‘fn” represents the frequency of spindle.
280 D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281

Fig. 14. The finished surface quality (a) without control, chatter (b) under control, stable.

The setup of the control experiment is shown in Fig. 12. It consists of three parts which are milling process (above
designed spindle), control unit (dSPACE, PC and an amplifier for the piezoelectric actuator) and monitoring and recording
unit (ECON and PC). In the milling process, the vibration displacement is firstly measured by the displacement sensor and
then passed to the dSPACE with frequency 10,000 Hz. Next, the control voltage is calculated by the designed MPC controller
based on the real-time measured displacement signal and then applied to the piezoelectric actuators through the amplifier.
Thus, the milling process can be controlled by the proposed MPC method. During the control process, the vibration displace-
ments and the control voltages are both monitored and recorded by the ECON data acquisition system.
The milling tool and the material of the work piece are listed in Table 1 (the overhang of the tool is 75 mm) and Table 2,
while the machining parameters are the same as those in Table 7. The chatter control effect can be analyzed from three
aspects.
First of all, in the two milling processes (without control and under control), the vibration displacements measured by the
displacement sensors at point P d in the X and Y directions and their frequency spectrums are shown in Fig. 13. It can be seen
that under control, the vibration displacements in the X and Y directions are both smaller than those without control.
Secondly, in the milling process without control, there are two chatter frequencies (f c1 ¼ 1159 Hz and f c2 ¼ 1562 Hz).
Similar to the simulation results, when the controller is on, the chatter frequencies disappear and only spindle revolution
frequency f n , the tooth-passing frequency and theirs harmonics are left in the frequency spectrums.
Thirdly, the finished surfaces of the work piece in two milling processes are shown in Fig. 14. Obviously, there are obvious
chatter marks on the finished surface without control, while the surface under control has much better machining quality.
All the results of these three aspects indicate that chatter is effectively controlled in the milling process with the MPC
controller designed in this work. Under control, the vibration caused by chatter is suppressed and only the steady vibration
due to static milling forces is left. The quality of the finished surfaces of the work piece is also significantly improved.

6. Conclusions and discussions

In this work, a model predictive control based method is developed to control chatter in the milling process. Firstly, we
design an active chatter control structure by integrating a specific tool holder, two pairs of piezoelectric actuators and two
displacement sensors to the spindle. Then, based on the designed structure and the milling process, a linear time-varying
model of the spindle-actuator-milling system is proposed. In order to eliminate the periodic variation of the system and
the existence of time-delay expression, zero order approximation and Padé approximation are applied and a LTI approximate
model integrating the chatter dynamics is finally developed, which is benefit to reduce the uncertainty caused by chatter. On
the base of the model, a model predictive controller is designed in the next. In the designing process of the controller, the
input constraint of the piezoelectric actuator is considered explicitly. Thus, the output of the controller can be the optimal
or suboptimal result within the limit of the actuator and the control command can be executed accurately and efficiently in
the real control process. Since the static milling force has no effect on chatter, it is reasonably ignored in the predictive
model. It means that only chatter dynamic is controlled in the milling process and the spindle system still vibrates steadily
under the static milling force. In this way, the required control voltage is smaller than that controlling the whole vibration of
the spindle system in the same milling process. So it is beneficial to saving energy and maximizing the stability of the spindle
system with limited power. In addition, the effect of uncertainty of the established model can be considered by receding
horizon optimization and feedback compensation, which are the two control strategies of the controller. So, strong robust-
ness of the controller can be achieved. Finally, in order to verify the effect of the active control method, simulation analysis
and milling experiments are carried out after two identification experiments. The vibration displacements are smaller and
the chatter frequencies are eliminated in the milling process under control compared with those without control. What’s
more, the finished surface of the work piece under control is also much better than that without control. All of these verify
that the MPC method developed in this work is effective to control chatter in the milling process.
In the work, the chatter dynamic caused by regenerative effect are integrated into the system model as well as the pre-
dictive model. It means that the controller designed based on the proposed method is related to the milling parameters.
When the milling parameters change, the controller will automatically change accordingly. As for the constant matrixes
and constants in the controller, such as the output error weight matrix Q , the control weight matrix R, the prediction horizon
N p and the control horizon N c , they can be fine-tuned based on the milling conditions. So the controller designed based on
the proposed MPC method remains active under different milling conditions. In the text, only one milling condition is ana-
lyzed. In the future research, the optimal settings of constant matrices and constants in the controller will be studied and
more milling conditions among the whole rotating range will be analyzed. The maximized SLD under control will also be
constructed and compared with that without control.
D. Li et al. / Mechanical Systems and Signal Processing 128 (2019) 266–281 281

Acknowledgments

This work is supported by the National Natural Science Foundation of China (Nos. 51575423, 11772244).

References

[1] H. Cao, Y. Lei, Z. He, Chatter identification in end milling process using wavelet packets and Hilbert-Huang transform, Int. J. Mach. Tools Manuf. 69
(2013) 11–19.
[2] H. Cao, Y. Yue, X. Chen, X. Zhang, Chatter detection in milling process based on synchrosqueezing transform of sound signals, Int. J. Adv. Manuf.
Technol. (2016) 1–9.
[3] H. Cao, K. Zhou, X. Chen, Chatter identification in end milling process based on EEMD and nonlinear dimensionless indicators, Int. J. Mach. Tools Manuf
92 (2015) 52–59.
[4] S. Xi, H. Cao, X. Zhang, X. Chen, Zoom synchrosqueezing transform-based chatter identification in the milling process, Int. J. Adv. Manuf. Technol.
(2018).
[5] N.D. Sims, B. Mann, S. Huyanan, Analytical prediction of chatter stability for variable pitch and variable helix milling tools, J. Sound Vib. 317 (2008)
664–686.
[6] Z. Dombovari, Y. Altintas, G. Stepan, The effect of serration on mechanics and stability of milling cutters, Int. J. Mach. Tools Manuf. 50 (2010) 511–520.
[7] Ahmad R. Yusoff, Neil D. Sims, Optimisation of variable helix tool geometry for regenerative chatter mitigation, Int. J. Mach. Tools Manuf. 51 (2) (2011)
133–141.
[8] V. Sellmeier, B. Denkena, Stable islands in the stability chart of milling processes due to unequal tooth pitch, Int. J. Mach. Tools Manuf. 51 (2011) 152–
164.
[9] M. Zatarain, I. Bediaga, J. Muñoa, R. Lizarralde, Stability of milling processes with continuous spindle speed variation: analysis in the frequency and
time domains, and experimental correlation, CIRP Ann. – Manuf. Technol. 57 (2008) 379–384.
[10] H. Cao, X. Zhang, X. Chen, The concept and progress of intelligent spindles: a review, Int. J. Mach. Tools Manuf. 112 (2017) 21–52.
[11] B. Denkena, O. Gümmer, Process stabilization with an adaptronic spindle system, Prod. Eng. Res. Devel. 6 (2012) 485–492.
[12] L. Sallese, A. Scippa, N. Grossi, et al, Investigating Actuation Strategies in Active Fixtures for Chatter Suppression, Procedia CIRP (2016) 311–314.
[13] L. Sallese, G. Innocenti, N. Grossi, A. Scippa, R. Flores, M. Basso, G. Campatelli, Mitigation of chatter instabilities in milling using an active fixture with a
novel control strategy, Int. J. Adv. Manuf. Technol. 89 (2017) 2771–2787.
[14] C. Wang, X. Zhang, Y. Liu, H. Cao, X. Chen, Stiffness variation method for milling chatter suppression via piezoelectric stack actuators, Int. J. Mach. Tools
Manuf 124 (2017).
[15] M. Ries, S. Pankoke, K. Gebert, Increase of material removal rate with an active HSC milling spindle, Conference proceedings of the adaptronic congress,
2006.
[16] F. Chen, X. Lu, Y. Altintas, A novel magnetic actuator design for active damping of machining tools, Int. J. Mach. Tools Manuf. 85 (2014) 58–69.
[17] X. Lu, F. Chen, Y. Altintas, Magnetic actuator for active damping of boring bars, CIRP Ann. – Manuf. Technol. 63 (2014) 369–372.
[18] S. Kern, A. Schiffler, R. Nordmann, E. Abele, Modelling and active damping of a motor spindle with speed-dependent dynamics, 2008.
[19] J.L. Dohner, J.P. Lauffer, T.D. Hinnerichs, N. Shankar, M. Regelbrugge, C.-M. Kwan, R. Xu, B. Winterbauer, K. Bridger, Mitigation of chatter instabilities in
milling by active structural control, J. Sound Vib. 269 (2004) 197–211.
[20] R. Kleinwort, M. Schweizer, M.F. Zaeh, Comparison of different control strategies for active damping of heavy duty milling operations, Procedia Cirp 46
(2016) 396–399.
[21] E. Graham, M. Mehrpouya, R. Nagamune, S.S. Park, Robust prediction of chatter stability in micro milling comparing edge theorem and LMI, CIRP J.
Manuf. Sci. Technol. 7 (2014) 29–39.
[22] D. Hajdu, T. Insperger, D. Bachrathy, G. Stepan, Prediction of robust stability boundaries for milling operations with extended multi-frequency solution
and structured singular values, J. Manuf. Processes 30 (2017) 281–289.
[23] M. Löser, A. Otto, S. Ihlenfeldt, G. Radons, Chatter prediction for uncertain parameters, Adv. Manuf. (2018).
[24] D. Hajdu, T. Insperger, G. Stepan, Robust stability analysis of machining operations, Int. J. Adv. Manuf. Technol. (2016) 1–10.
[25] N.V. Dijk, N.V.D. Wouw, E. Doppenberg, O. Han, Chatter control in the high-speed milling process using l-synthesis, Am. Control Conf. (2010) 6121–
6126.
[26] N.J.M.V. Dijk, N.V.D. Wouw, E.J.J. Doppenberg, A.J.O. Han, H. Nijmeijer, Robust active chatter control in the high-speed milling process, IEEE Trans.
Control Syst. Technol. 20 (2012) 901–917.
[27] N.V.D. Wouw, N.J.M.V. Dijk, A. Schiffler, H. Nijmeijer, E. Abele, Experimental Validation of Robust Chatter Control for High-Speed Milling Processes,
Springer International Publishing, 2017.
[28] N.V.D. Wouw, N.J.M.V. Dijk, H. Nijmeijer, Pyragas-type feedback control for chatter mitigation in high-speed milling, Ifac Papersonline 48 (2015) 334–
339.
[29] Y. Wu, H.T. Zhang, T. Huang, G. Ren, H. Ding, Robust chatter mitigation control for low radial immersion machining processes, IEEE Trans. Autom. Sci.
Eng. 1 (2018).
[30] X. Zhang, C. Wang, J. Liu, R. Yan, H. Cao, X. Chen, Robust active control based milling chatter suppression with perturbation model via piezoelectric
stack actuators, Mech. Syst. Sig. Process. 120 (2019) 808–835.
[31] F. Shi, H. Cao, X. Zhang, X. Chen, A chatter mitigation technique in milling based on H 1 -ADDPMS and piezoelectric stack actuators, Int. J. Adv. Manuf.
Technol. (2018).
[32] T. Huang, Z. Chen, H.-T. Zhang, H. Ding, Active control of an active magnetic bearings supported spindle for chatter suppression in milling process, J.
Dyn. Syst. Meas. Contr. 137 (2015) 111003.
[33] H.T. Zhang, Y. Wu, D. He, H. Zhao, Model predictive control to mitigate chatters in milling processes with input constraints, Int. J. Mach. Tools Manuf 91
(2015) 54–61.
[34] Y. Altintas, Manufacturing Automation: Metal Cutting Mechanics, Machine Tool Vibrations, and CNC Design, Cambridge University Press, 2012.
[35] L. Wang, Model Predictive Control System Design and Implementation Using MATLABÒ, 39 (2009) 25–26.

You might also like