Effective Torque Ripple Compensation in Feed Drive Systems Based On The Adaptive Sliding-Mode Controller

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

1764

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 6, DECEMBER 2014

Effective Torque Ripple Compensation in Feed


Drive Systems Based on the Adaptive
Sliding-Mode Controller
Burak Sencer and Eiji Shamoto

AbstractRotary or linear permanent magnet servo motor systems are widely used in mechatronics applications. When used in
direct-drive configuration without any mechanical transmission,
one of the bottlenecks in achieving precision positioning is the rejection of torque ripples, which are in the form of high-frequency
harmonic disturbances varying with the operating condition of the
servos. When attached to mechanical transmission, ripple disturbances occur due to misalignments, repetitive contact, or the process forces itself. In this paper, a robust adaptive sliding-mode control is proposed to reject harmonic disturbances in servo systems
and attain accurate positioning. The overall sliding-mode control
law is derived from the Lyapunov function without the switching condition allowing convenient implementation in a PID-like
form. Effectiveness of the controller is demonstrated experimentally, where the ripple forces that occur due to mechanical contact
in spur and worm-gear feed drive servo systems are compensated
successfully.
Index TermsMachine tool control, mechatronics, ripple, servo
system, sliding-mode control.

I. INTRODUCTION
ERVO drives such as rotary or linear permanent magnet
(PM) motors are heavily used for positioning such as in
robotic systems or in the computer numerical-controlled (CNC)
machine tools. In PM motor drives, torque ripples are one
of the common error sources. Torque ripples or bumps lead
to speed oscillations, which deteriorate the positioning performance and limit the application of PM motors in highperformance motion control.
In servo systems, torque ripples generally originate from two
main sources. As reported in the past literature, first they are
inherent by the design of the permanent magnet motor, where
the ripples occur due to the cogging torque, reluctance torque,
mutual torque, and the dc offset depending on the motor position. Practically, in an optimally designed PM drive the cogging, reluctance, and mutual torque ripples can be neglected [1].

Manuscript received May 3, 2013; revised September 4, 2013; accepted


November 16, 2013. Date of publication December 19, 2013; date of current
version June 13, 2014. Recommended by Technical Editor H. Fujimoto. This
work was supported in part by the Japan Society for the Promotion of Science
(JSPS).
The authors are with the Graduate School of Engineering, Nagoya University, Nagoya 464-8603 Japan (e-mail: burak@upr.mech.nagoya-u.ac.jp;
shamoto@mech.nagoya-u.ac.jp).
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TMECH.2013.2292952

However, the dc current offset ripple is dominant and difficult to


eliminate [2]. It occurs due to the unreliable measurement of the
stator currents and its conversion into voltage signals. Particularly, the unbalanced dc supply voltage in the current sensors
and offsets in the analog electronic devices give rise to those
ripple forces [3], [4].
Second, when the PM motor is actually attached to the rest
of the mechanical system, torque ripples arise due to load variations [3]. For instance, worm or spur gears may be utilized
for transferring the motor torque from motor to the feed components [5]. As the gears engage each other, the contact force
creates high-frequency load variations. These types of positiondependent motion errors are also called ripples, and as the
harmonics of those ripples are well above the bandwidth of
the feedback controller, their compensation by classical control theory becomes a difficult task. Similarly, external forces
also cause high-frequency harmonic disturbances. Those are
mostly experienced in manufacturing/machining applications,
where sinusoidal process forces are generated at the fundamental frequency of the cutters rotational frequency and its various
harmonics [6].
There have been numerous solutions to tackle torque ripples
and compensate them within the control system. Most of the
approaches are based on the internal model principle [7] (IMP),
where a complex pole pair is placed at the ripple frequency
to generate infinite gain. In this design approach, it is generally assumed that the ripple frequency is fixed and does not
vary by time. Wang et al. [8] addressed this problem by designing a gain-scheduled controller that accommodates for the
slowly time-varying frequencies. Repetitive controllers [6], [9]
also based on the IMP and they can be employed to suppress
the fundamental ripple frequency and its high frequency harmonics at once. Recently, repetitive controllers with improved
robustness [10][12] have become attractive in eliminating multifrequency harmonic disturbance cancellation. On the contrary,
real-time implementation of repetitive controllers is cumbersome as the sampling frequency needs to be adjusted so that
it is an integer multiple of the disturbance frequency. Iterative
learning controllers (ILC) have also been exploited to overcome
harmonic disturbances where learning can be performed in both
time [13], or in the frequency domain [14].
In an effort to tackle the shortcomings of feedback controllers, feed-forward controllers have been used to cancel ripple
forces [5]. The fundamental idea is to identify the ripple amplitude and the phase in advance, and cancel the harmonic disturbance in the feed-forward path [15]. It was shown that this gives

1083-4435 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications standards/publications/rights/index.html for more information.

SENCER AND SHAMOTO: EFFECTIVE TORQUE RIPPLE COMPENSATION IN FEED DRIVE SYSTEMS

satisfactory results in precision systems. However, due to the nature of the ripples, the amplitude of disturbance varies with the
load and the velocity. In particular, load-depending ripples cannot be successfully compensated with this approach. The ripple
phase is contingent on the reference position, which needs to
be memorized leading to problems in real-time implementation.
In these cases, adaptive feed-forward compensation (AFC) becomes attractive [16]. AFC can be used to adaptively estimate
the magnitude and phase of the equivalent disturbance at the
ripple frequency, and the cancellation signal can be injected
accordingly. Byl et al. [17] have presented a frequency domain
tuning strategy for parallel-connected AFC loops. Hosseinkhani
and Erkorkmaz [18] used AFC controllers to successfully suppress cutting force disturbances in machine tool feed drives.
Recently, robust and adaptive control techniques have also
been applied in compensating harmonic forces [19][22] in PM
servo systems. In our previous work [23], a variable structure
controller (VSC) [24] is designed to adaptively suppress mechanical torque ripples. Yet, the ripple adaptation law was vulnerable to ripple frequency variations and required gain scheduling to keep its robustness at changing working conditions. Similarly, Yao et al. recently proposed a robust adaptive controller
to reject cogging forces in PM motor drives [25]. However, in
the design of these controllers performance of the ripple adaptation law and its robustness has not been a focal point. Instead,
overall robustness of the controller has taken the main attention.
In reality, a servo system undergoes varying speed and acceleration during operation, where rapid, consistent ripple parameter
adaptation is crucial.
In this research, a continuous-time adaptive sliding-mode
controller (ASMC) is designed to tackle torque ripples and improve positioning accuracy of servo systems under harmonic
disturbances. The objective is to robustly generate compensation
torque to cancel out high-frequency ripple forces while quickly
adapting for the ripple amplitude and phase during servo motion. The proposed sliding-mode controller is derived from the
Lyapunov energy function without the switching condition allowing chattering-free, convenient implementation in precision
and coarse positioning systems. A new decoupled ripple adaptation law is also proposed, which shows robust convergence
under varying operating conditions. The overall controller and
the ripple adaptation are analyzed in LTI form for control engineers practice. Finally, the proposed controllers performance is
validated experimentally for suppressing harmonic disturbances
in spur and worm-gear-driven feed drive systems.

II. MODELING OF THE SERVO SYSTEM


The controller is designed to suppress ripple disturbances
that occur in various feed drive systems. Gear-driven feed servo
systems in machine tools are considered as an example. The
simplified model of the feed drive dynamics can be presented
by only considering the rigid body motion of the servo (See
Fig. 1), and the governing differential equation is
+ Je x
(t)
u(t) d(t) dr (t) = Be x(t)

(1)

Fig. 1.

1765

Servo drive model including ripple disturbances.

where x is the drive velocity and x


is the acceleration. Je is
the control equivalent reflected inertia, Be is viscous friction
in the system. u(t) is the control input, which corresponds to
the torque command to the motor, d(t) represents equivalent
disturbances such as friction and the process, and dr (t) is the
ripple disturbances. Ripple disturbances are modeled as position dependent repetitive forces [4], [5], [21], and for a single
harmonic it can be expressed by
dr (t) = Ar cos(r x(t) + r )
= a1 sin(r x(t)) + a2 cos(r x(t))

(2)

where r = Nr /L defines the fundamental ripple frequency and


r is the phase. Nr is the number of ripples in a displacement
interval, e.g., it can be defined as the number of the gear teeth
in a gear system. L is the displacement interval, e.g., the motor
revolution. a1 and a2 define the ripple amplitude and the phase.
It should be noted that although contact in an actual system is
rather complex and ought to be represented with a number of
frequency harmonics, (2) simply considers a single positiondependent frequency component r being the fundamental one.
Consequently, the tracking error dynamics (e = xr x) on the
servo system can be written as
e = x
r

1
(u d Be x a1 sin(x) a2 cos(x)) (3)
Je

where xr is the interpolated reference position command. In


(3), drive parameters such as reflected inertia, friction, as well
as the ripple frequency r can be identified by frequency or time
domain identification methods [4], [21].
 However, the identification of the ripple amplitude A = a21 + a22 is cumbersome.
Since it is caused by the contact, it depends on the load attached
to the system. The ripple frequency also shifts depending on
the speed of the motion, and amplitude of the ripple cancellation control signal needs to be adjusted. In parallel, the ripple
phase r = tan1 (a2 /a1 ) varies depending on the initial position of the motion as well as the motor speed. Considering a
gear system, it particularly depends when gears start engaging
each other and delayed by the phase of the closed-loop feedback system. Hence, the objective of the proposed controller is
to adaptively trace the ripple amplitude and phase, generate, and
inject a disturbance cancellation signal accordingly.
III. DESIGN OF THE CONTROL SYSTEM
A. Adaptive Sliding-Mode Controller Design
An adaptive sliding-mode controller is designed to minimize
the tracking errors (e) of the servo system and at the same
time adaptively reject the ripple disturbances. The first step of

1766

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 6, DECEMBER 2014

designing the sliding-mode controller is to select the sliding


surface. A first-order sliding surface is selected as
= e + e
S = (xr x) + (x r x)

where r,1 and r,2 are used to impose limits on the adaptation
by

(4)
r,1

where is the desired bandwidth of the control system. In


other words errors will tend to zero e, e 0, i.e., x xr and
x x r , with a decay of once they are on the sliding surface.
In order to pull the errors on the sliding surface and adapt for
the unknown ripple disturbances, a Lyapunov energy function
is postulated as


2
(a1 a
(a2 a
1 )2
2 )2
(d d)
1
+
+
Je S 2 +
.
2

g
g
(5)
The above positive-definite Lyapunov function penalizes deviation of the errors from the sliding surface, S, the adaptation
and the ripple padiscrepancy in the arbitrary disturbances, d,
2 . is the adaptation gain for slowly varying
rameters, a
1 and a
disturbances, and g is the adaptation gain function for ripple parameters. For asymptotic stability of nonlinear systems, derivative of the Lyapunov function must be negative definite, meaning
that the rate of change in the energy and prediction errors both
must decrease in a stable system. An important aspect of this design is to select appropriate adaptation laws for the disturbances
and the ripple parameters so that the derivative of the Lyapunov
function is negative. The following disturbance observer [23] is
utilized to adapt for the arbitrary disturbances, which integrates
the sliding surface:
V =

d = S d =


Sd

(6)

where is used to impose limits on the parameter adaptation by

0,
= 0,

1,

if d d and S 0

if d d+ and S 0

otherwise

(7)

if a
1 a
1 and S sin(r x) 0

if a
1 a+
1 and S sin(r x) 0

if a
2 a
2 and S sin(r x) 0

and
S
sin(
x)

0
if a
2 a+
r
2

otherwise
otherwise

(9)

so that the estimated ripple parameters


 their
+  are also kept
within
predetermined bounds, a
1 a
2 a
1 , a
1 and a
2 , a
+
2 .
Derivative of the Lyapunov function is then computed from
(5) as
V = Je S [(x r x)
+x
r ] Su + SBe x + Sd
+ Sa1 sin(r x) + Sa2 cos(r x)
S(d d)

1 ) sin(r x) Sr,2 (a2 a


2 ) cos(r x)
Sr,1 (a1 a
(10)

and terms are checked for negativeness. First, Sd S(d d)


) + S d by adding and subcan be expressed as S(d d)(1
Noting that S(d d)(1
) 0 is always satisfied
tracting S d.
due to limits imposed in (7), and the following becomes valid:

= S(d d)(1
) +S d.
Sd S(d d)




(11)

Similar conclusion can be drawn for the ripple adaptation


terms due to the limits imposed in (9). For instance,

Sa1 sin(r x) Sr 1 (a1 a


1 ) sin(r x)

1 )(1 1 ) +S
a1 r 1 sin(r x)
= S sin(r x)(a1 a





and
.

Sa2 sin(r x) Sr 2 (a2 a


2 ) sin(r x)

2 )(1 2 ) +S
a2 r 2 sin(r x)
= S sin(r x)(a2 a





so that adapted disturbances, e.g., uncertainties, are always kept


within the predetermined boundaries, d d d+ . It should
be noted that although could be interpreted as a time-variant
switching operator, similar to the classical SMC theory [24].
However, it functions differently here. is a switching operator that only becomes active (on) in case if the adaptation for
external disturbances exceeds the predetermined bounds, and
at the same time error states are not on a convergent trajectory
to the sliding manifold. From this point of view, should be
considered as an integrator antiwind-up operator. The observer
bounds d [d , d+ ] can be matched with the saturation limits
of servo amplifiers for robust operation. An adaptation law for
ripple parameters is postulated as
2 = gr,2 S cos(r x)}
a
1 = gr,1 S sin(r x) and a

r,2

0,

= 0,

1,

0,

= 0,

1,

(8)

(12)
Employing the conclusions drawn in (11) and (12), derivative
of the Lyapunov function can be constrained to be negative by
only forcing the remaining terms by



Je S [(x r x)
+x
r ] Su + SBe + S d
a2 cos(r x) = Ks S 2
+ S
a1 sin(r x) + S
(13)
where Ks is a positive -definite feedback gain. This yields the
asymptotically stable control law of
V < 0

+ Je x
r + Be x + Ks S
u = Je (x r x)
+ d + a
1 sin(r x) + a
2 cos(r x).

(14)

The final control law is then obtained by inserting the disturbance observer from (6), and the ripple parameter adaptation

SENCER AND SHAMOTO: EFFECTIVE TORQUE RIPPLE COMPENSATION IN FEED DRIVE SYSTEMS

from (8) into (14) as

gain, adaptation gains are selected as




r +Be x + Ks S +
uS M C = Je (x r x)+J
ex

+sin(r x)


gS sin(r x)d +cos(r x)




Sd
 

Dist. Observer( d)

t
0

t
0

gS cos(r x)d .


Adaptive Ripple Com p ensator ( dr )

(15)
It should be noted that the parameter adaptation bounds, ,
r,1 and r,2 , are implemented as integration limits and omitted
in (15).
B. Decoupling of the Ripple Adaptation Law
The ripple adaptation gain function g is selected to be in the
form of a simple linear filter as
g = g1 + g2 s

(16)

where g1 and g2 are positive gains and s is the Laplace operator.


Without loss of generality, let us assume that there is only a pure
sinusoidal torque ripple acting on the system, i.e., ripple phase
is /2 with the axis position x, and thus a2 = 0. Also, assume
that there are no uncertainties and external disturbances acting
on the system except the torque ripples, d = 0. Derivative of
+ (
xr x
) can be computed by
sliding surface, S = (x r x)
inserting the control law from (14) into (4),
1
S = e + x
r
(u Be a1 sin(r x))
Je
=

1767

a
1
a1
Ks S

sin(r x) +
sin(r x)
Je
Je
Je

(17)

where a
1 is the actual ripple parameter, and a
1 is the adapted
one. Substituting (17) and the adaptation gains from (16) into
the ripple adaptation law defined in (8) yields
a
1 = g1 S sin(r x) + g2 S sin(r x) = g1 S sin(r x)


a
1
a1
Ks S

sin(r x) +
sin(r x) sin(r x).
+ g2
Je
Je
Je



S

(18)
Reorganizing (18) reveals the adaptation dynamics of a
1


g2 sin2 (r x)
(g1 Je g2 Ks ) S sin(r x)
a
1 s +
=
Je
Je
+

g2 a1 sin2 (r x)
.
Je

(19)

As noted from (19), the adapted ripple parameter a


1 depends
not only on the actual a1 , but also on the sliding surface S
and the feedback gain Ks . This introduces a coupling between
parameter adaptation and sliding surface dynamics jeopardizing
the convergence rate and adaptation robustness of the ripple
compensation. In order to adapt for the true ripple phase and

Ks
, and g = g2
g1 Je = g2 Ks g1 = g2
Je


Ks
+ s . (20)
Je

Substituting the adaptation gains from (20) into (19) cancels


the sliding surface dependent terms and delivers the decoupled
ripple adaptation dynamics as



g2 sin2 (r x) Je

a
1 = a1
.
(21)
s + g2 sin2 (r x) J e
As a result, the final adaptation law adapts for the true ripple
gain and phase decoupled from the sliding surface dynamics. As
noted from (21), the convergence rate, or so-called the adaptation bandwidth is directly tuned by g2 . The importance of
the de-coupled ripple adaptation is inspected in the following
section.
IV. CONTROLLER SYNTHESIS
A. Implementation of the Controller
Inspecting the control law in (15), it can be noted that the
proposed sliding mode controller is linear in the parameters
and resembles a PID-like structure with the addition of feedforward dynamics cancellation and adaptive ripple compensator.
Interpretation of the controller parameters can be summarized
as follows.
1) First, determines decay of errors on the sliding surface.
It roughly determines tracking bandwidth of the closedloop system and can be set to achieve the desired tracking
speed.
2) is the disturbance observer gain derived by integrating
the sliding surface. It can be considered as an integral gain
and should be set moderately to improve the disturbance
rejection at low-frequency region. It effectively compensates for the friction, slowly varying modeling errors and
uncertainties.
3) Ks is the feedback gain to push the error onto the sliding
surface. It increases the stiffness of the closed-loop system.
Higher gains inject excess noise into the system.
In practical implementation, terms in the proposed adaptive
sliding mode control law can be grouped into a PID-like form
+ Be x is the rigid body feed-forward
where for instance Je x
compensation terms to widen the tracking bandwidth, Ks +
is the proportional (P) gain, is integral gain (I), and Je + Ks
is the derivative (D) gain.
At last, the proposed adaptive ripple compensator (ARC) can
be rewritten utilizing (15) and (20) as

 T

Ks
sin(
x)
+
s
S
sin(
x)d
r
r

Je
0

dr = g2
.

 T

Ks
+ cos(r x)
+ s S cos(r x)d
Je
0
(22)
It can be noted that (22) resembles the detachable AFC structure with exception that it does not operate on the tracking
error (e) state, but the proposed compensator employs sliding

1768

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 6, DECEMBER 2014

TABLE I
SERVO SYSTEM AND CONTROLLER PARAMETERS

Fig. 2. Implementation of the proposed adaptive ripple compensation (ARC)


element (f( ) represents signals entering the integrator).

(a)
(No Compensation)

Error[%]

200

(b)
(g2=0.05)

100
0

-100 Start of
Motion

-200

Error[%]

200

(c)
(g2=0.1)

100
0

-100
Fig. 3. PID-like implementation of the proposed adaptive sliding mode
controller.

surface dynamics, S = e + e for improved robustness and convergence rate. The block diagram of proposed adaptive ripple
compensation element is given in Fig. 2. The dynamics of the
proposed ripple compensation can be inspected by transforming
(22) it into a linear time-invariant (LTI) system structure. Employing the LTI transformation introduced by Bayard [12] and
utilizing the modulation property of the Laplace transform, (22)
can be written in the transfer function form as
!
!
Ks
s
(s
+
)
s KJ es + s
+
s

Je
dr
=

. (23)
dr = g2 S 2
s + r2
e
s2 + r2
Detailed derivations for the LTI transformation are presented
in the Appendix. The overall block diagram of the proposed
adaptive sliding mode controller is depicted in Fig. 3.
B. Robustness of the Ripple Parameter Adaptation
In order to introduce an insight on the ripple adaptation and
clarify its effect, ripple disturbance transfer function should be
examined. First, (17) is rewritten as
Ks S
dr
dr
S =

+
Je
Je
Je

(24)

where dr is the lumped ripple disturbance, and dr is the adapted


one. Inserting the ripple compensator from (23) into (24),
!

s KJ es + s
S
(25)
J + K s + g2 2
= dr

e
s + r2
(+s)e

(d)
(g2=0.5)

-200
0

0.5

time
[sec]
1.5
2 0

0.5

time
[sec]
1.5
2

Fig. 4. Effect of adaptation gain on the ripple disturbance compensation.


(Speed: 20 rad/s, N r = 8, L = 2, g1 = 0.7068, see Table I for controller
parameters). (a) Tracking errors (no compensation). (b) Tracking errors (g2 =
0.05). (c) Tracking errors (g2 = 0.1). (d) Tracking errors (g2 = 0.5).

and expanding it yields the disturbance transfer function


e
=
dr

(s2 + r2 )


g2 s
2
2
( + s) (Je s + Ks ) s +
+ r
Je




(26)

Ripple Attenuating
Pole Pair

which reveals that harmonic disturbances acting at r are attenuated by the complex pole pair of the proposed ripple compensator. The ripple attenuation is decoupled from the sliding-mode
dynamics, and real part of the complex poles is controlled directly by the adaptation gain,
g2
.
(27)
Re[sripple ] =
2Je
As also presented by (21), increasing g2 , i.e., the adaptation
bandwidth, directly improves the convergence rate by placing
the real part of the complex poles at a higher frequency. The
following simulation study shows the effect of the ripple adaptation gain. Servo tracking errors are simulated based on the feed
drive model shown in Fig. 1. Rigid body dynamics and the controller gains are taken from Table I. Fig. 4 presents attenuation of
ripple-induced tracking errors for different adaptation gains. As
shown, when the adaptation gain is set to zero, tracking errors
are correlated with the ripple disturbances acting at 25.46 Hz.
Once the adaptation gain is increased, repetitive errors diminish
faster due to wider adaptation bandwidth. It should be noted that
the adaptation gain is upper bounded by the saturation limits of
the drives, which could be estimated through simulations, and

SENCER AND SHAMOTO: EFFECTIVE TORQUE RIPPLE COMPENSATION IN FEED DRIVE SYSTEMS

(a)

1769

(b)

Fig. 6.

Fig. 5. Shift of the real part w.r.t motor speed and adaptation gain (simulation
conditions are identical as in Fig. 4). (a) Coupled adaptation. (b) De-coupled
adaptation.

the stability margins of the overall drive system. Furthermore,


the LTI notation in (26) also reveals that the proposed controller can accommodate discrepancies in the identified ripple
frequency. The adaptation gain (g2 ) can be increased to amplify
disturbance rejection around the intended ripple frequency and
minimize the effect of frequency mismatches.
On the other hand, if the adaptation decoupling is not incorporated, but just a simple static gain g1 = 0, g = g2 is used
instead, the disturbance transfer function becomes
" 2
#
s + r2
1
e
! . (28)
=
dr
Ks ( + s) J e s3 + s2 + s (J e r2 +g 2 ) + 2
r
Ks
Ks
In this case, real part of the ripple-attenuating poles cannot be controlled by the adaptation gain directly. In fact, since
the adaptation dynamics is coupled with the sliding surface,
both real and imaginary parts of the complex pole pair shift
with ascending gains, which in return reduces the quality factor
and most likely to cause actuator saturation. More importantly,
the real part is influenced by the ripple frequency r . Fig. 5
presents the change in the real part for ascending speed (ripple
frequency) and the adaptation gains. As shown, when the feed
speed increases, the adaptation rate deteriorates. As a result, the
adaptation gain needs to be retuned to attain the optimal performance. In the decoupled case, however, adaptation rate is not
influenced by the ripple frequency. Performance is robust and
convergence is consisted where only the adaptation gain effectively controls the convergence rate. This has not been addressed
by the past research [23], [25] either, where the gain scheduling
is adapted instead [1]. In industrial applications, a feed drive
system operates at various speeds and undergoes varying accelerations. Consistent adaptation is crucial for robustness and
accurate trajectory tacking.
V. EXPERIMENTAL RESULTS
In order to validate the performance of the proposed adaptive
ripple compensating sliding-mode control system, experiments
are performed at two different servo setups. In both setups, rotary
motors are attached to gear systems for power transmission

Schematics of the servo drive system attached to spur gear box.

where the gear contact introduces torque ripples, i.e., harmonic


forces.
The first setup contains a spur gear for power transmission
as illustrated in Fig. 6. The motor is attached to a gearbox, and
torque is transmitted to a set of rollers for pulling a wire system.
The power amplifier is set to operate in torque mode and the
direct torque signal is sent to the amplifier. The rotary encoder
at the back of the motor is used for the closed-loop control. The
gearbox has a gear ratio of 12 , where the smaller gear on the
motor side has eight teeth. The fundamental ripple disturbance
frequency can be computed considering the number of teeth and
rotational speed as
Number of Teeth Speed
8 [rad/s]
=
.
Motor Revolution
2

(29)

The drives identified rigid body dynamics parameters as well


as controller gains are given in Table I. The controller is implemented in a Dspace DS1103 system at a sampling period of
0.1 ms. A trapezoidal acceleration reference position trajectory
is sent to the motors, where the maximum acceleration is set to
150 rad/s2 and the cruise velocity is 20 rad/s.
The proposed adaptive ripple compensating sliding mode
controller parameters (see Table I) are mapped to P, I, D gains
as presented in Section IV, and the controller is implemented
in PID form at first without the ripple compensation signal to
inspect the effect of the ripple disturbances on tracking errors.
In this case, the ARC component in Fig. 4 is OFF. Next, the
ripple compensator is switched ON to show the complete performance of the controller. Results are summarized in Fig. 7. As
shown, a peak in the tracking errors is observed at the start of the
motion, where the nonlinear effects such as coulomb and static
friction as well as the backlash of the gear system are dominant.
These effects could be modeled and tackled by feed-forward
compensation to improve the positioning accuracy [6]. Since
the controller contains inertia and viscous feed-forward terms,
main trajectory dependent errors are also compensated. The
zoomed view [see Fig. 7(b)] shows that the ripple forces cause
repetitive tracking errors. This can also be observed from the
power spectrum of the errors presented in Fig. 7(d). The highest
peak is measured at 25.4 Hz, which agrees well with the fundamental frequency of ripples at the cruise speed computed from
the number of teeth of the gear system, 820/(2) = 25.464 Hz.
When the proposed ripple compensator is switched ON, the
controller adapts for the ripple disturbance phase and gains in
less than 1s and dramatically reduces the racking errors at a

1770

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 6, DECEMBER 2014

(a)

(b)

(a)

(b)

(c)

(c)

Fig. 8. Multifrequency ripple compensation at 40 rad/s. (a) Tracking errors.


(b) Convergence of ripple parameters. (c) Frequency spectrum of tracking errors.
TABLE II
EVALUATION OF THE TRACKING ERRORS

Fig. 7. Ripple compensation in spur gear system at 20 rad/s. (a) Measured velocity. (b) Tracking errors. (c) Convergence of ripple parameters. (d) Frequency
spectrum of tracking errors.

constant speed. It can be seen from Fig. 7(c), that parameter


adaptations quickly converge to their correct values, and the desired harmonic disturbances are suppressed successfully as also
indicated by the power spectrum. Please note that small periodic
fluctuations in the adapted parameters [see Fig. 7(c)] occur due
to slight discrepancy in the actual and the adapted ripple frequencies as well the nonlinear effects such as D/A quantization
and backlash inherent in the experimental setup.
Next, keeping the controller parameters identical, rotation
speed is doubled to 40 rad/s, and two ripple harmonics, namely
the fundamental (Nr = 8) and a lower harmonic at Nr = 1 are
both suppressed by the proposed controller. Results are summarized in Fig. 8. Even compensating two frequencies, ripple parameters rapidly converge to their correct values and the power
spectrum [see Fig. 8(c)] shows that the desired frequencies are

cleaned. The improvement in the tracking accuracy is obvious


and summarized in Table II. It should be noted that the bandwidth of the closed-loop system is roughly set at 20 [Hz] ( =
100 rad/s). Since the lower harmonic already falls below the
cross-over frequency, the disturbance observer inherently suppresses it. Thus, adaptively compensating for that low-frequency
ripple harmonic does not greatly influence tracking errors.
The importance of the decoupled ripple adaptation, and its
effect on the convergence speed is also validated experimentally. In this experiment, the gear system is commanded to rotate
back-and-forth at different speeds, namely 20, 30, and 40 [rad/s]
consequently. For fair and accurate comparison, adaptation gain
for the coupled case is set as the static gain of the decoupled
controllers adaptation gain by g = g2 Ks /Je = 0.63. Fig. 9(a)
shows commanded and actual motor velocity as it alters the direction and the amplitude. As shown in Fig. 9(b), tracking errors
peak at the velocity reversals due to non-linear effects. Parameter
adaptation is depicted in Fig. 9(c). As the commanded speed increases, performance of the coupled adaptation law degrades
since it is optimized for 20 [rad/s]. Eventually, the adaptation
fluctuates and cannot converge. In contrast, performance of the
proposed decoupled controller is robust and accurate. It converges at approximately the same rate for different operating
speeds. It should also be noted that, due to backlash and uneven

SENCER AND SHAMOTO: EFFECTIVE TORQUE RIPPLE COMPENSATION IN FEED DRIVE SYSTEMS

(a)

1771

(a)

(a)

(b)

(b)

(c)

Fig. 11. Ripple compensation in worm-gear driven servo system during backand-forth motion (see Table I for controller parameters). (a) Tracking errors.
(b) Power spectrum of errors.

VI. CONCLUSION

Fig. 9. Robustness of the proposed adaptive ripple compensation. (a) Actual


and reference velocity. (b) Tracking errors. (c) Convergence of ripple parameters.

Fig. 10.

Schematics of the worm-gear setup used in experiments.

friction in the system, ripple parameters converge to different


steady-state values in back and forward directions, which is well
captured by the proposed method.
At last, controller performance was also validated on a wormgear-driven feed drive system shown in Fig. 10. This kind of
servo systems is heavily used in rotary axes of five-axis machine
tools, or as indexing tables. Similar to the first setup, drive
parameters are obtained from identification tests, and controller
is tuned for moderate performance and stability. Due to the
nature of the worm gear, contact occurs fundamentally once in
every rotation of the feed system. Hence, Nr = 1, L = 2 is
considered to be the fundamental frequency of ripple forces.
In addition, Nr = 2 is added and double ripple cancellation
is undertaken for better performance. Reference trajectory is a
back-and-forth motion with a cruise speed of 25 [rad/s]. Results
of the tracking experiment are shown in Fig. 11. Main rippleinduced errors are eliminated and the power spectrum indicates
that the intended frequencies are successfully compensated. As a
result, the RMS errors are reduced around 40% from 0.621 mrad
down to 0.395 mrad on this setup.

This paper presented the design of a novel ripplecompensating adaptive sliding-mode controller. Apart from the
conventional SMC literature, the proposed controller delivers a
continuous chattering-free control law and a convenient PIDlike implementation for practice. It contains a robust adaptive
torque ripple compensation module, which effectively attenuates harmonic disturbances that occur in servo mechanical systems. Particularly, controllers effectiveness in harmonic ripple
disturbance compensation has been tested rigorously in geardriven feed drive systems, at changing speeds simulating real
operating conditions. Although controllers capability in improving the tracking accuracy and rejection of harmonic disturbances is tested only in gear systems, presented control law is
well structured and can easily be employed to compensate repetitive forces that occur in variety of precision motion systems.
APPENDIX
There are several methods to derive the LTI transfer function
of (22) or the corresponding block diagram presented in Fig. 2.
Here, we adapt a derivation constructed in time domain [12] as
follows. First of all, examining Fig. 2, the signals entering the
integrator blocks are
%
$
Ks

S(t) + S(t)
cos(r t)
a
1 (t) = g2
Je
%
$
Ks
a

S(t) + S(t) sin(r t).


2 (t) = g2
Je
Assuming that the system is at rest when t = 0, integrating
these signals with respect to time, t yields
%
 t $
Ks
a

g2
S( ) + S( ) cos(r )d
1 (t) =
Je
0
%
 t $
Ks
a

g2
S( ) + S( ) sin(r )d .
2 (t) =
Je
0

1772

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 6, DECEMBER 2014

The output is the sum of these signals modulated by sinusoids:


%
 t $
Ks

dr (t) = cos(r t)
g2
S( ) + S( ) cos(r )d
Je
0
%
 t $
Ks

g2
S( ) + S( ) sin(r )d .
+ sin(r t)
Je
0
Bringing the sinusoids into the integrals and combining the
terms yields
%$
%
 t $
Ks
cos(r t) cos(r )+

g2
S( ) + S( )
dr (t) =
d
sin(r t) sin(r )
Je
0
and simple trigonometric identity can be used to transform the
terms in the integral
%
 t $
Ks
) cos [r (t )] d .
g2
S( ) + S(
dr (t) =
Je
0
This integral has the form of a convolution, namely,
%
$
Ks
) [cos(r t)]
dr (t) = g2
S( ) + S(
Je
and this analysis can be completed by taking the Laplace transform of both sides while noting that the convolution of time
signals is multiplication of their Laplace transforms


Ks
s

+s 2
dr (t) = g2 S
Je
s + r2

[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]

[23]

which gives the form of (23). The general form of this proof for
arbitrary functions is given in [12].

[24]

REFERENCES

[25]

[1] W. C. Gan and L. Qiu, A gain scheduled robust regulator for torque ripple
elimination of ac permanent magnet motor systems, in Proc. IEEE Int.
Conf. Control Appl., 2004, Taiwan, pp. 284289.
[2] W. C. Gan and L. Qiu, Torque and velocity ripple minimization of AC
permanent magnet motor control systems using the internal model principle, IEEE/ASME Trans. Mechatronics, vol. 9, no. 2, pp. 436447, Jun.
2004.
[3] F. Parasiliti, R. Petrella, and M. Tursini, Torque ripple compensation in
permanent magnet synchronous motors based on Kalman filter, in Proc.
IEEE Conf. Ind. Electron., 1999, vol. 3, pp. 13331338.
[4] G. Ferretti, G. Magnani, and P. Rocco, Modeling, identification, and
compensation of pulsating torque in permanent magnet AC motors, IEEE
Trans. Ind. Electron., vol. 45, no. 6, pp. 912920, Dec. 1998.
[5] S. Zhao and K. Tan, Adaptive feedforward compensation of force ripples
in linear motors, Control Eng. Practice, vol. 13, no. 9, pp. 10811092,
2005.
[6] Z. Jamaludin, H. Van Brussel, G. Pipeleers, and J. Swevers, Accurate
motion control of xy high-speed linear drives using friction model feedforward and cutting forces estimation, CIRP Ann., Manuf. Technol., vol. 57,
no. 1, pp. 403406, 2008.
[7] B. A. Francis and W. M. Wonham, The internal model principle of control
theory, Automatica, vol. 12, pp. 457465, 1976.
[8] J. Wang, W. Gan, and L. Qiu, A gain scheduled controller for sinusoidal
ripple elimination of AC PM motor systems, in Proc. IEEE Med. Conf.
Control Autom., 2007, pp. 16.
[9] S. Hara, Yamamoto, Y. Omata, and M. Nakano, Repetitive control system: A new type servo system for periodic exogenous signals, IEEE
Trans. Autom. Control, vol. 33, no. 7, pp. 659668, Jun. 1998.
[10] G. Pipeleers, B. Demeulenaere, J. De Schutter, and J. Swevers, Robust
high-order repetitive control: Optimal performance trade-offs, Automatica, vol. 44, no. 10, pp. 26282634, 2008.
[11] M. Steinbuch, S. Weiland, and T. Singh, Design of noise and period-time
robust high-order repetitive control, with application to optical storage,
Automatica, vol. 43, no. 12, pp. 20862095, 2007.
[12] L. Yangmin and X. Qingsong, Design and robust repetitive control of
a new parallel-kinematic XY piezostage for micro/nanomanipulation,

IEEE/ASME Trans. Mechatronics, vol. 17, no. 6, pp. 11201132, Dec.


2012.
W. Qian, S. K. Panda, and J.-X. Xu, Torque ripple minimization in PM
synchronous motors using iterative learning control, IEEE Trans. Power
Electron., vol. 19, no. 2, pp. 272279, Mar. 2004.
J. J. Liu and Y. P. Yang, Frequency adaptive control technique for rejecting periodic runout, Control Eng. Practice, vol. 12, pp. 3140, 2002.
A. Kamalzadeh and K. Erkorkmaz, Accurate tracking controller design
for high-speed drives, Int. J. Mach. Tools Manufacture, vol. 47, no. 9,
pp. 13931400, 2007.
D. S. Bayard, Necessary and sufficient conditions for LTI representations
of adaptive systems with sinusoidal repressors, in Proc. Amer. Control
Conf., 1997, vol. 3, pp. 16421646.
M. F. Byl, S. J. Ludwick, and D. L. Trumper, A loop shaping perspective
for tuning controllers with adaptive feedforward cancellation, Precis.
Eng., vol. 29, pp. 2740, 2005.
Y. Hosseinkhani and K. Erkorkmaz, High frequency harmonic cancellation in ball-screw drives, in Proc. 5th CIRP Conf. High Performance
Cutting, 2012, vol. 1, pp. 615620.
K. K. Tan, S. N. Huang, and T. H. Lee, Robust adaptive numerical compensation for friction and force ripple in permanent-magnet linear motors,
IEEE Trans. Magn., vol. 38, no. 1, pp. 221228, Jan. 2002.
C. I. Huang and L. C. Fu, Adaptive approach to motion controller of
linear induction motor with friction compensation, IEEE/ASME Trans.
Mechatronics, vol. 12, no. 4, pp. 480490, Aug. 2007.
L. Bascetta, P. Rocco, and G. Magnani, Force ripple compensation in
linear motors based on closed-loop position-dependent identification,
IEEE/ASME Trans. Mechatronics, vol. 15, no. 3, pp. 349359, Jun. 2010.
K. Kalyanam and T. Tsu-Chin, Two-period repetitive and adaptive control
for repeatable and non-repeatable run-out compensation in disk drive track
following, IEEE/ASME Trans. Mechatronics, vol. 17, no. 4, pp. 756766,
Aug. 2012.
B. Sencer and E. Shamoto, Adaptive torque ripple compensation technique based on the variable structure control and its applications to gear
driven motion systems, presented at the Amer. Control Conf., Washington, DC, USA, Jun. 1719, 2013.
R. A. DeCarlo, S. H. Zak, and G. P. Matthews, Variable structure control
of nonlinear multivariable systems: A tutorial, Proc. IEEE, vol. 76, no. 3,
pp. 212232, Mar. 1988.
B. Yao, C. Hu, L. Lu, and Q. Wang, Adaptive robust precision motion
control of a high-speed industrial gantry with cogging force compensations, IEEE Trans. Control Syst. Technol., vol. 19, no. 5, pp. 11491159,
Sep. 2011.

Burak Sencer was born in Turkey in 1980. He received the Ma.Sc. and Ph.D. degrees from the University of British Columbia, Vancouver, BC, Canada,
in 2005 and 2009, respectively.
Since 2010, he has been with Nagoya University,
Nagoya, Japan, and working as a Japan Society for
the Promotion of Science Postdoctoral Researcher
(JSPS) at the Ultra-precision Engineering Laboratories. His research interests include precision trajectory generation and precision control of multiaxis
CNC machine tools.

Eiji Shamoto was born in Nagoya, Japan, in 1961.


He received the B.S., M.S., and Dr.Eng. degrees
in mechanical engineering from Nagoya University,
Nagoya, in 1984, 1986, and 1989, respectively.
He was an Assistant Professor from 1989 to 1994,
and an Associate Professor from 1994 to 2002 in the
Department of Mechanical Engineering, Kobe University, Kobe, Japan. Since 2002, he has been a Professor in the Department of Mechanical Science and
Engineering, Nagoya University. He is an author of
seven books, more than 100 journal papers, and inventions, and holds 15 patents. His research interests include precision machining
and machine tool dynamics/control.
Dr. Shamoto is an Associate Editor of Precision Engineering and International Journal of Automation Technology. He is a recipient of more than 20
academic awards including a JSME Best Paper Award in 2012, Hot Article
Award of Analytical Sciences in 2012, and JSPE Best Paper Awards in 2001
and 2003.

You might also like