Anda di halaman 1dari 13

432

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 2, APRIL 2014

Damping and Tracking Control Schemes


for Nanopositioning
Arnfinn A. Eielsen, Marialena Vagia, J. Tommy Gravdahl, and Kristin Y. Pettersen

AbstractFast and accurate tracking of reference trajectories is


highly desirable in many nanopositioning applications, including
scanning probe microscopy. Performance in common positioning
stage designs is limited by the presence of lightly damped resonances and actuator nonlinearities such as hysteresis and creep. To
improve the tracking performance in such systems, several damping and tracking control schemes have been presented in the literature. In this paper, six different control schemes are presented and
applied to a nanopositioning system for experimental comparison.
They include schemes applying damping control in the form of
positive position feedback, integral resonant control, integral force
feedback, and passive shunt-damping. Also, general pole placement in the form of model reference control, as well as a control
scheme requiring only a combination of a low-pass filter and an
integrator, is presented. The control schemes are fixed-structure,
low-order control laws, for which few results exist in the literature with regard to optimal tuning. A practical tuning procedure
for obtaining good tracking performance for five of the control
schemes is, therefore, presented. Experimental results show that
the schemes provide similar performance, and the main differences
are due to the specific implementation of each scheme.
Index TermsMotion control, nanopositioning, output feedback, piezoelectric devices, scanning probe microscopy.

I. INTRODUCTION
ANOPOSITIONING devices are used for high-resolution
positioning, including, but not limited to, scanning probe
microscopy and its many applications for imaging and manipulation. Some tasks require periodic reference trajectory tracking,
typically for imaging, while other tasks require arbitrary reference trajectory tracking, such as for manipulation, fabrication,
and lithography. To improve throughput in such applications,
high-bandwidth control is required [1][3].
Most high-speed nanopositioning devices use piezoelectric
actuators, as they produce large forces and provide friction-less
motion. As such, they are ideal for high-speed, high-resolution
positioning. A positioning device applying piezoelectric actua-

Manuscript received June 12, 2012; revised October 11, 2012; accepted
December 8, 2012. Date of publication February 14, 2013; date of current
version February 20, 2014. Recommended by Technical Editor Y. Sun. This
work was supported by the Norwegian University of Science and Technology
and the Norwegian Research Council. Part of this work was carried out during
the tenure of an ERCIM Alain Bensoussan Fellowship Programme. This Programme is supported by the Marie Curie Co-Funding of Regional, National and
International Programmes of the European Commission.
The authors are with the Department of Engineering Cybernetics, Norwegian
University of Science and Technology, N-7491 Trondheim, Norway (e-mail:
eielsen@itk.ntnu.no; Marialena.Vagia@itk.ntnu.no; Tommy.Gravdahl@itk.
ntnu.no; Kristin.Y.Pettersen@itk.ntnu.no).
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.2013.2242482

tors typically exhibit lightly damped resonances. This is a disadvantage, as it limits the usable bandwidth because reference
signals with high-frequency components will excite the vibration modes, prohibiting accurate positioning. It also makes the
device susceptible to environmental disturbances, such as sound
and floor vibrations. The hysteresis and creep nonlinearities in
piezoelectric actuators is an additional challenge. These are loss
phenomena that prevent the system from having a linear response, introducing bounded input disturbances dependent on
the driving voltage signal [3]. There also exist several sources
of dynamic uncertainty. Hysteresis, in addition to introducing
an input disturbance, change the effective gain of the actuator depending on the amplitude and frequency of the driving
voltage signal. Actuator gain is also dependent on temperature
and reduces over time due to depolarization [4], [5, Ch. 4]. In
addition, users typically need to position payloads of various
masses; thus, resonant frequencies and the effective gain of the
mechanical structure can change as a result.
Tracking control for nanopositioning devices can be achieved
using feed-forward and feedback control techniques. Although
feed-forward techniques can provide very good results [2], [6],
feedback control may be necessary in order to reduce sensitivity to uncertainty and disturbances. Combining feed-forward
and feedback control can improve overall tracking performance
[7], [8]. In order to control lightly damped vibrational modes in
active structures, several control schemes that introduce damping have been developed. These include fixed-structure, loworder control laws, such as positive position feedback (PPF)
[9], integral force feedback (IFF) [10], passive shunt-damping
(PSD) [11], resonant control [12], and integral resonant control
(IRC) [13]. By coupling such schemes with an integral control law, significantly better reference tracking performance can
be achieved. With the exception of PSD, this has been experimentally demonstrated in [14][16]. The main reason for the
increased performance is that a reduction of the dominant resonant peak of the system leads to an increased gain margin,
enabling much higher gain to be used for the reference tracking
integral control law [16].
General model-based control laws can also be used, such as
H -control law synthesis [17], the linear-quadratic-Gaussian
regulator [18], and output feedback control laws such as pole
placement and model reference control (MRC) [18], [19]. H control has seen widespread application on nanopositioning systems [20][23]. Nonlinear control approaches, such as slidingmode control, have also been applied [7], [24].
The advantage of using fixed-structure, low-order control
laws is mostly practical, as such control laws are simple
to implement and have low computational complexity. This

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.

EIELSEN et al.: DAMPING AND TRACKING CONTROL SCHEMES FOR NANOPOSITIONING

Fig. 1. Custom flexure-guided nanopositioning stage (the mechanical design


is described in detail in [29]).

allows for the highest possible sampling frequency when implementing using digital signal processing equipment, which
will reduce the noise floor due to quantization in the analog-todigital converter [25]. The simplicity also makes them feasible
for implementation using analog circuit elements. This can be
beneficial, as it avoids noise due to sampling and quantization
altogether. The disadvantage of using fixed-structure, low-order
control laws, is a lack of methods for optimal tuning, and this
is a long standing and challenging control engineering problem [26], [27].
A. Contributions
Six different damping and tracking control schemes are presented and applied to a nanopositioning system for experimental comparison. All the schemes combine integral action with a
control law that introduces damping of the dominant vibration
mode of the system. The damping control schemes considered
are PPF, IRC, IFF, and PSD.
A control scheme based on the work in [28] is also presented,
and the tuning methodology therein is generalized and also
applied to the presented control schemes based on PPF, IRC,
IFF, and PSD. IRC has been applied in [15], and IFF in [16]. PPF
and polynomial-based control (also known as pole placement)
are applied in [14]. Furthermore, in this paper, pole placement
in the form of MRC is applied, also incorporating integral action
and filtering, in order to reduce sensitivity to disturbances and
uncertainty, and to reduce quantization noise.
II. SYSTEM DESCRIPTION AND MODELING
A. Description of the Experimental System
The experimental setup consists of a dSPACE DS1103
hardware-in-the-loop (HIL) system, an ADE 6810 capacitive
gauge, an ADE 6501 capacitive probe from ADE Technologies, a Piezodrive PDX200 voltage amplifier, two SIM 965
programmable filters from Stanford Research Systems, and
the custom-made long-range serial-kinematic nanopositioner
shown in Fig. 1. The nanopositioner is fitted with a Noliac
SCMAP07-H10 actuator, where one of the stack elements is
used as a force transducer. The transducer current is measured

433

Fig. 2. Simplified schematic of the single-degree-of-freedom flexure-guided


positioning stage.

Fig. 3.

System diagram.

using a Burr-Brown OPA2111 configured with a 10-k resistor,


thus having a sensitivity of 10 V/mA. The capacitive probe
has a bandwidth of 100 kHz, and the voltage amplifier with
the capacitive load of the actuator has a bandwidth in excess
of 100 kHz. The voltage amplifier is also fitted with a current
monitor with a sensitivity of 1 V/A, which enables the current in
the actuator circuit to be measured. The capacitive measurement
has a sensitivity of 1/5 V/m and the voltage amplifier has a
gain of 20 V/V.
With the DS1103 board, a sampling frequency of fs =
100 kHz was used for all the experiments, as this is the highest
possible sampling frequency attainable in closed loop. This ensures the lowest possible noise floor due to quantization in the
analog-to-digital converter [25, Ch. 13]. For numerical integration, a third-order RungeKutta scheme was used, which was
the highest order integration scheme that could be solved in real
time with the chosen sampling frequency.
A diagram of the system used is shown in Fig. 3. The positioner dynamics is represented by Gp (s), the amplifier and
reconstruction filter dynamics by Wr (s), and the sensor and
antialiasing filter dynamics by Wa (s). The signal u is the input
generated by the digital-to-analog converter (DAC), ym is the
output from the antialiasing filter, n is the sensor noise, and w
is the input disturbance, mostly caused by hysteresis, creep, and
environmental vibration noise.
B. Mechanical Model
The nanopositioning stage used is shown in Fig. 1, and a
simplified schematic is shown in Fig. 2. The serial-kinematic
motion mechanism is designed to make the first vibration mode
dominant and to occur in the actuation direction (piston mode).
More details on the design of this stage can be found in [29].
The displacement is generated using a piezoelectric actuator. Such actuators generate a force proportional to an applied

434

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 2, APRIL 2014

voltage [30]. The applied external force from the piezoelectric


actuator fa (N) can be expressed as
fa = kp (ua + w)

(1)

where kp (N V1 = C N1 ) is the effective gain of the piezoelectric actuator from voltage to force, and ua (V) is the applied
voltage. Piezoelectric actuators introduce hysteresis and creep
when driven by an external voltage signal. These effects occur
in the electrical domain [31], and it is a reasonable assumption
to consider this behavior as a bounded disturbance added to the
input, represented by the term w (V) [32].
The dynamics due to an applied voltage ua or disturbance
w of a point d (m) on the flexible structure, as observed by
a colocated sensor, is adequately described by the following
lumped parameter, truncated linear model [33]:
d

d
i
(s)
+ Dr
2 + 2 s + 2
fa
s
i i
i
i=1

Gd (s) = kp

(2)

where nd is the number of vibration modes included. Here, {i }


(ms2 V1 ) are the control gains, {i } (-) are the damping
coefficients for each mode, and {i } (rads1 ) are the natural
frequencies for the modes. The term Dr (mV1 ) is the residual
mode, which is an approximation of the nonmodeled higher
frequency modes, and can be included to improve prediction of
zero locations. The addition of Dr produces a model that is not
strictly proper, but as the instruments, such as the amplifier and
sensors, have limited bandwidth, Dr can be considered equal to
zero for this system. Equation (2) has a pole-zero interleaving
property [33], which is the origin of positive realness (passivity)
and negative imaginariness for certain inputoutput pairs [34].
The inclusion of instrumentation dynamics and sensor-actuator
pairs that are not perfectly colocated will, in general, invalidate
the pole-zero interleaving property [33].

Fig. 4. Response and uncertainty for the displacement. (a) Frequency response
for the displacement model G d (s). (b) Multiplicative uncertainty weight d (s).

found to be [16], [33]


vf = ks (d ka ua )
where d is the displacement of the mechanical structure, ua is
the applied voltage to the actuator, ka (mV1 ) is the gain of
the feed-through term, and ks (Vm1 ) is the sensor gain. The
transfer function from applied voltage ua to measured sensor
voltage vf can, therefore, be found as
vf
Gf (s) =
(s) = ks (Gd (s) ka ).
(4)
ua

C. Charge

E. Identification and Uncertainty

When applying PSD, the induced charge in the actuator circuit


is utilized. The charge in the actuator circuit can be found as [35]

In order to identify the parameters in (2)(4), the frequency


responses for the displacement, charge, and force are recorded
using an SR780 Dynamic Signal Analyzer from Stanford Research Systems, applying a 150-mV RMS bandwidth-limited
white noise excitation signal. The models are fitted into the
procured data using the MATLAB System Identification and
Optimization Toolboxes. As the noise from the force transducer
is orders of magnitude lower than the noise from the displacement sensor [16], the frequency response for the displacement
is inferred from (4). The frequency response obtained using the
displacement sensor is used to find the parameters in (3) and
(4). The responses for Gd (s), Gq (s), and Gf (s) are displayed
in Figs. 4(a), 5(a), and 6(a), respectively. The identified parameter values are presented in Table I. For the displacement model
(2), three vibration modes, nd = 3, are included. By inspection
of Fig. 4(a), it can be seen that the second mode at 1660 Hz is
the dominant piston mode.
The uncertainty of the models can be quantified as unstructured multiplicative perturbations. Since the control schemes
considered are either single-input single-output (SISO) or

q = kp d + Cp ua = Cp (ua + d)
where Cp (F) is the capacitance of the piezoelectric stack actuator, and (Vm1 ) is a constant determining the amount of
charge generated by the direct piezoelectric effect due to the displacement d of the mechanical structure. The transfer function
from applied voltage ua to induced charge q is therefore
Gq (s) =

q
(s) = Cp (1 + Gd (s)).
ua

(3)

D. Force Transducer
The IFF scheme utilizes a colocated piezoelectric force transducer. The force transducer generates a charge, depending on
the applied force. The current or charge produced by the force
transducer is typically converted to a voltage signal using a simple op-amp circuit with a high input impedance. The output
voltage from such a sensor when measuring the charge can be

EIELSEN et al.: DAMPING AND TRACKING CONTROL SCHEMES FOR NANOPOSITIONING

435

TABLE I
IDENTIFIED MODEL PARAMETERS

Fig. 7.
Fig. 5. Response and uncertainty for the charge. (a) Frequency response for
the charge model G q (s). (b) Multiplicative uncertainty weight q (s).

General control structure.

are determined experimentally, for each of the outputs, and are


presented in Figs. 4(b), 5(b), and 6(b). Overbounding weights
were also found to introduce more conservativeness.
The conservativeness of the weights should reflect the uncertainty in the actuator gain due to, e.g., hysteresis, temperature, and depolarization, depending on the specific application
characteristics of the system. Control law performance will deteriorate if there are large perturbations of the dynamics of the
system. A fixed-structure, low-order control law can only accommodate for a limited amount of uncertainty, and trying to
tune the system, while considering large uncertainty weights
can result in bad performance, and might not yield any robustly
stable solution for the control law parameters.
III. CONTROL DESIGN
The control schemes presented will be analyzed with regard
to the general control structure shown in Fig. 7.
A. Performance Measures
The control schemes considered are either SISO or SIMO.
Considering the general SIMO case, it can be seen that for the
control structure in Fig. 7, the control input is given as

Fig. 6. Response and uncertainty for the force. (a) Frequency response for the
force model G f (s). (b) Multiplicative uncertainty weight f (s).

single-input multiple-output (SIMO), the uncertainty description of the models from the scalar input up to the output vector
yp has the form [36]
yp i = Gi (s)(1 + i (s)i (s))up ; i (s) 1

up = C(s)(r F (s)yp )

(6)

where C(s) is a one-row feed-forward transfer matrix, and F (s)


is a diagonal feedback transfer matrix.
Breaking the loop at the error e of the one-column plant
transfer matrix Gp (s), the loop transfer matrix is

(5)

where i denotes the index into the output vector yp , such that
Gi (s) corresponds to the transfer function from the input up
to the output yp i , and i (s) is the corresponding frequency
dependent uncertainty weight. The uncertainty weights {i (s)}

L(s) = F (s)Gp (s)C(s)


which defines the output sensitivity transfer matrix SO (s) as
e = SO (s)r = (I + L(s))1 r

(7)

436

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 2, APRIL 2014

where e = r F (s)yp . The complementary sensitivity transfer


matrix T (s) becomes
yp = T (s)r = Gp (s)C(s)SO (s)r.

(8)

In addition, the transfer matrix N (s) from the additive sensor


noise n to the output yp is
yp = N (s)n = T (s)F (s)n

(9)

and the transfer matrix E(s), measuring the deviation of the


plant output yp from the reference trajectory r,  = r yp , is
 = E(s)r = (I T (s))r.

(10)

Note that  = e, if F (s) = I.


Breaking the loop at the input up of the plant, the loop transfer
matrix is
LI (s) = C(s)F (s)Gp (s)
and the input sensitivity transfer matrix SI (s) from the disturbance w to the input up is therefore
up = SI (s)w = (I + LI (s))1 w

(11)

which provides the transfer matrix D(s) from the disturbance


w to the output yp as
yp = D(s)w = Gp (s)SI (s)w.

(12)

The performance will be evaluated with regard to the flatness


of the response of T (s), the bandwidth of E(s), the attenuation of the input disturbance w to the displacement d, and the
amplification of sensor noise n to the displacement d. The bandwidth of E(s) is defined as the frequency where |E(j)| first
crosses the line of 3 dB from below at the frequency response
diagram. At this frequency, the tracking error  is 50% of the
reference r; thus, there is effectively no tracking of frequency
components above the bandwidth. The attenuation of the input disturbance is measured by the H -norm, D(s) , which
corresponds to the peak magnitude of D(s). The added displacement noise is measured by the H2 -norm, N (s)2 , which
provides the root-mean-square displacement noise response if n
is taken to be equal to unity variance Gaussian white noise [37].
That is, the displacement variance due to sensor noise can be
found as d 2 = N (s)22 n 2 .
B. Robust Stability Measure
The SIMO robust stability criterion described in [36], for
multiplicative uncertainty on the form (5), can straightforwardly
be adapted to the control structure in Fig. 7. Robust stability can
be ensured if
sup

ny


|SI (j)Gi (j)Ci (j)Fi (j)i (j)| = 1 (13)

i=1

where the matrix elements Ci (s) and Fi (s) correspond to the


output yp i , and ny is the number of outputs. The inverse value of
the norm, 1/, provides a measure of the minimum amount of
additional multiplicative uncertainty that the system can tolerate
before it becomes unstable, for the given frequency weights, i .

C. Tuning
Control design for fixed-structure, low-order control laws using output feedback is a long-standing and challenging problem
in control engineering. A common approach to output feedback
problems is to use H -synthesis. If the control law is allowed to
have any order and every matrix of the control law is freely tunable, H -synthesis guarantees a solution to the control design
problem by convex optimization.
For a control law with a fixed structure and with a lower order
than the plant, this approach cannot be applied. There exist some
results for fixed low-order control problems, solved with the use
of linear matrix inequalities, but these methods do not allow for
the use of unstructured uncertainty, do not guarantee global,
and in many cases not even local, convergence, and might not
accommodate for control laws where the structure is fixed in
addition to the order [26], [27]. In other words, there does not
exist any general control design method for output feedback
using fixed-structure, low-order control laws.
In this paper, a practical optimization procedure is, therefore,
proposed in order to obtain good tracking performance.
Control design is often a tradeoff between conflicting goals.
For nanopositioning systems, it is desirable to have a high bandwidth for E(s) in order to have good reference tracking. Also,
the system is required to be well damped in order to avoid excessive vibrations. This translates to an absence of peaks in T (s).
To counter hysteresis and creep, as well as environmental disturbances, D(s) must provide a high degree of attenuation. In
addition, the amplification of sensor noise should be as small
as possible, meaning that N (s) should have the smallest bandwidth possible. Due to the restriction imposed by the Bode
sensitivity integral [18], it is impossible to meet these criteria
simultaneously.
As the purpose of damping control is to reduce peaks in the
closed-loop response due to lightly damped vibration modes,
and since ideal tracking performance is achieved when
d = r T (s) = 1
it appears that a good overall performance criterion is the flatness
of |T (j)|. Let c be the vector of control law parameters. It is
here proposed that the flatness criterion can be expressed using
the cost function
J(c ) = 1 |T (c ; j)|2

(14)

where  2 is the L2 -norm. The expression 1 |T (c ; j)| is


typically not square-integrable with respect to ; thus, the L2 norm is truncated as needed in order to produce a finite value,
i.e., integrating over the domain [0, ], where > 0 is well
above the bandwidth of the mechanical system.
For the control schemes presented in the following sections, a
practical and straightforward method to find control law parameters that provide good tracking performance for a particular
scheme is then to solve
c = arg min J(c ) s.t. Re{i } R < 1

(15)

where {i } is the set of eigenvalues for the closed-loop system. The optimization problem can be solved either by using

EIELSEN et al.: DAMPING AND TRACKING CONTROL SCHEMES FOR NANOPOSITIONING

Fig. 8.

437

Control structure when using PPF and IRC.


TABLE II
OPTIMAL PARAMETERS FOR (16) AND (17)

an exhaustive grid search over a domain of reasonable control


law parameter values, or by using an unstructured optimization
algorithm, such as the NelderMead method [38].
IV. CONTROL SCHEMES
A. Positive Position Feedback
Damping and tracking control using PPF [9], [33] combined
with an integral control law can be implemented using the control structure in Fig. 8. This is equivalent to the control scheme
in [14]. The damping control law consists of a low-pass filter
with negative gain
kd
Cd (s) = 2
s + 2d d s + d 2

(16)

where kd > 0 is the control law gain, d is the damping coefficient, and d is the cutoff frequency. The tracking control law
is an integral control law with a negative gain, which will be
inverted by the negative gain of the filter (16)
kt
.
(17)
s
Here, kt > 0 is the gain of the integral term.
To analyze the nominal performance of the control scheme,
the control structure in Fig. 8 can be put on the equivalent
formulation adhering to the control structure in Fig. 7. The
feed-forward filter is found as
Ct (s) =

C(s) = Wr (s)Cd (s)Ct (s)

(18)

and the feedback filter is found as


F (s) = Wa (s)(1 + Ct1 (s)).

(19)

Using the above expressions and assuming


Gp (s) = Gd (s)
it is straightforward to find the transfer functions for the sensitivity (7), the complementary sensitivity (8), the noise attenuation
(9), the error attenuation (10), and the disturbance rejection (12).
Here, Wr (s) = Wa (s) and are second-order Butterworth filters
with a cutoff frequency at 20 kHz.
The robust stability with regard to the multiplicative model
uncertainty can be evaluated using the stability criterion (13),
using (8), (18), and (19).
There are four tunable control law parameters
c = [gf , f , f , kt ]T
the feedback filter gain kd , the damping ratio d , the cutoff
frequency d , and the tracking integral control law gain kt .
The optimal control law parameters for (16) and (17) found
when solving (15) are found in Table II. The resulting nominal

frequency responses for T (s), E(s), and D(s) are shown in


Fig. 12(a).
B. Integral Resonant Control
Damping and tracking control applying IRC [13] to introduce
damping can also be implemented using the control structure in
Fig. 8. In this control scheme [15], the damping control law is
Cd (s) =

kd
.
s kd Df

(20)

Equation (20) is the result of rearranging the IRC scheme to


a form suitable for tracking control [15]. Here, kd > 0 is the
called the integral damping gain, while Df is a feed-through
term. The tracking control law is
kt
(21)
s
where kt > 0 is the gain of the integral term.
As this scheme uses the same control structure as the one
using PPF in Section IV-A, the scheme can be analyzed using the
same equivalent formulation with regard to the general control
structure in Fig. 7, i.e., with
Ct (s) =

C(s) = Wr (s)Cd (s)Ct (s)

(22)

F (s) = Wa (s)(1 + Ct1 (s)).

(23)

and

Here, Wr (s) = Wa (s) and are second-order Butterworth filters


with a cutoff frequency at 20 kHz.
There are three tunable control law parameters
c = [Df , kd , kt ]T
the feed-through term Df , the integral damping gain kd , and
the tracking integral control law gain kt . The optimal control
law parameters for (20) and (21) found when solving (15) are
found in Table III. The resulting nominal frequency responses
for T (s), E(s), and D(s) are shown in Fig. 12(b).
C. Integral Force Feedback
The dual-sensor damping and tracking control scheme proposed in [16] is based on IFF [10], [33] and can be implemented

438

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 2, APRIL 2014

TABLE III
OPTIMAL PARAMETERS FOR (20) AND (21)

TABLE IV
OPTIMAL PARAMETER FOR (24)

This is a SIMO system, and the measurement vector is given


as
yp = [d, vf ]T

(28)

while the input is the applied voltage ua . With regard to Fig. 7,


the plant transfer matrix is
Gp (s) = [Gd (s), Gf (s)]T
Fig. 9.

the feedforward transfer matrix is given as

Control structure when using IFF.

C(s) = [Wr (s)Ci (s), Wr (s)Ci (s)]


using the control structure shown in Fig. 9, where Gf (s) is as
described in (4).
The advantage of using this scheme is that the piezoelectric
force transducer has a noise density which is orders of magnitude
lower than a capacitive probe, thus allowing high bandwidth, but
with substantially lower displacement noise due to feedback.
The drawback is reduced range, as the force sensor replaces
parts of the actuator, and additional instrumentation to amplify
the charge generated by the transducer. The force sensor also
requires good calibration, and the response is slightly nonlinear,
but this is an insignificant source of error at higher frequencies.
The control scheme requires an integral control law
Ci (s) =

ki
s

(24)

to be implemented, where ki > 0 is the gain. In addition, two


splitting filters, a low-pass and a high-pass filter, must be implemented. The low-pass filter is given as
Wlp (s) =

(29)

f
s + f

(25)

while the high-pass filter is given as


s
Whp (s) =
s + f

(26)

(31)

Here, Wr (s) = Wa (s) and are second-order Butterworth filters


with a cutoff frequency at 20 kHz.
There is one tunable control law parameter
c = ki
where ki is the integral gain. The optimal control law parameter
for (24) found when solving (15) is found in Table IV. It can
also be found using a root-locus plot. The resulting nominal
frequency responses for T (s), E(s), and D(s) are shown in
Fig. 12(c).
Remark 1: Neglecting the instrumentation dynamics, and using a sufficiently low cutoff frequency f , the complementary
sensitivity function from r to d can be approximated as
T (s) =

where the gain is found as


= |Gd (0)/Gf (0)| .

and the feedback transfer matrix is given as




Wlp (s)Wa (s)
0
F (s) =
.
0
Whp (s)Wa (s)

(30)

Ci (s)Gd (s)
1 Ci (s)Gf (s) + Wlp (s)Ci (s) [Gd (s) + Gf (s)]
Ci (s)Gd (s)
1 Ci (s)Gf (s)

(32)

which is an unconditionally stable transfer function [39].


D. Passive Shunt-Damping

(27)

The cutoff frequency f determines the split between the


frequency range for where to use displacement feedback, and
where to use force feedback. For the experimental implementation, this was chosen to be f = 2 50. Better noise properties
can be achieved by reducing f , but this is limited to some
extent by the need to high-pass filter the force measurement.
The high-pass filter is used both to allow the use of the capacitive probe measurement at low frequencies, and to remove
bias components in the charge measurement. As the force transducer response is slightly nonlinear, sufficient bandwidth for the
capacitive probe measurement is required to improve linearity.

PSD [11] can introduce damping by adding an inductor and a


resistor in series with the piezoelectric actuator, which acts as a
capacitor, due to the large dielectric constant of the piezoelectric
material. Tuning the resulting LCR circuit for maximal damping
creates a resonant LCR circuit that works analogously to a tuned
mechanical absorber. Adding an integral control law for tracking
results in the control structure shown in Fig. 10, where Gq (s)
is as given in (3). As discussed below, this configuration does
not result in the same tuning of the LCR circuit as would be the
case when optimizing for maximum damping only.
The transfer function for the added shunt is [35]
Z(s) = sL + R

(33)

EIELSEN et al.: DAMPING AND TRACKING CONTROL SCHEMES FOR NANOPOSITIONING

439

Fig. 11. Control structure when introducing damping by applying a low-pass


filter in the signal chain, in this case by utilizing the filter W r (s)W a (s).

Fig. 10.

Control structure when using PSD.


TABLE V
OPTIMAL PARAMETERS FOR (33) AND (34)

where L (H) is the inductance and R () is the resistance. The


integral control law is
Ci (s) =

ki
s

(34)

where ki > 0 is the gain.


This can be interpreted as a SIMO system, where the measurement vector is given as
yp = [d, vf ]T

(35)

and the input is the applied voltage ua . With regard to Fig. 7,


the plant transfer matrix is
Gp (s) = [Gd (s), Gq (s)]T

(36)

the feed-forward transfer matrix is


C(s) = [Wr (s)Ci (s), sZ(s)]
and the feedback transfer matrix is given by


Wa (s) 0
F (s) =
.
0
1

(37)

(38)

Here, Wr (s) = Wa (s) and are second-order Butterworth filters


with a cutoff frequency at 20 kHz.
There are three tunable control law parameters
c = [L, R, ki ]T
the shunt inductance L, the shunt resistance R, as well as the
tracking integral control law gain ki . The optimal control law
parameters for (33) and (34) found when solving (15) are found
in Table V. The resulting nominal frequency responses for T (s),
E(s), and D(s) are shown in Fig. 12(d).
It should be noted that the values for L and R found using (15)
are not the same as when optimizing for maximum damping.
In that case, L = 46.6 mH and R = 165 , which produces a
resonant LCR circuit response due to the much smaller resistance
value. Since
Gd (s)
d (s) = d (s) =
= Ws (s)Gd (s)
G
us
1 + sZ(s)Gq (s)
when the shunt is present, the source voltage us is filtered by
Ws (s) before it is applied to the actuator. The response ob-

tained for Ws (s) when using the values of L and R in Table V


is almost identical
 to that of a Butterworth filter with cutoff frequency c = 1/LCp , although there is still some additional
damping introduced due to the Gd (s) term in (3). This means
that the shunt can be approximated by a second-order low-pass
Butterworth filter, which is done in Section IV-E. It might also
be noted that when using PSD, the low-pass filter Ws (s) makes
the use of the antialiasing filter Wa (s) unnecessary.
The shunt was implemented using an inductor constructed
using a closed ferrite core and approximately 45 turns of copper
wire. A potentiometer was used to implement the required resistance. The inductor and resistor were tuned to their required
values using an Agilent U1733C LCR meter.
E. Damping Integral Control
As noted in Section IV-D, the optimal values for the resistance and inductance for the shunt result in a low-pass filter when
connected to the capacitance of the actuator, with approximately
the same response as a second-order low-pass Butterworth filter. Implementing a control scheme on a microcontroller or a
computer, there must be reconstruction and antialiasing filters
present in order to avoid aliasing and to reduce quantization
noise. Instead of applying a shunt circuit, the reconstruction and
antialiasing filters that are already present as part of the instrumentation can be used. The resulting control structure is shown
in Fig. 11 and corresponds to the scheme presented in [28].
As when using PSD, only an integral control law needs to be
implemented, i.e.,
Ci (s) =

ki
s

(39)

where ki > 0 is the gain. The cutoff frequency, c , for the


filters Wr (s) and Wa (s) must be tuned as well. Here, it is
assumed that Wr (s) = Wa (s) for simplicity. The filters used in
the experimental setup are second-order Butterworth filters:
Wr (s) = Wa (s) =

2
c
.
s2 + 2c s + c 2

(40)

The combined filter Wr (s)Wa (s) is of fourth order, but the


closed-loop response of the system is almost identical to the
case when using a passive shunt. The added benefit is that the
shunt is no longer needed.
Formulating the control scheme in terms of the general control
structure in Fig. 7, the feed-forward filter is
C(s) = Wr (s)Ci (s)

(41)

and the feedback filter is


F (s) = Wa (s).

(42)

440

Fig. 12.

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 2, APRIL 2014

Nominal frequency responses. (a) PPF. (b) IRC. (c) IFF. (d) PSD. (e) DI. (f) MRC.
TABLE VI
OPTIMAL PARAMETERS FOR (39) AND (40)

There are two tunable control law parameters


c = [c , ki ]T
the filter cutoff frequency c , and the tracking integral control
law gain ki . The optimal control law parameters for (39) and (40)
found when solving (15) are found in Table VI. The resulting
nominal frequency responses for T (s), E(s), and D(s) are
shown in Fig. 12(e).
F. Model Reference Control
The MRC [19] objective is to make the plant output yp track
the output of a reference model ym . Similar to polynomial-based
control, or pole placement [18], [19], MRC provides a control
law for arbitrary closed-loop pole placement, but also allows
for arbitrary minimum phase zero placement. A second-order
pole-placement control law was applied to a nanopositioning

stage in [40]. The synthesis equations for the MRC scheme are
summarized in the Appendix.
For the MRC design, the displacement model of the system
Gd (s) is truncated to only include the dominant piston mode at
1660 Hz, the second mode of the positioning stage,
2
d (s) = Gd (0)
G
.
2 /2 2 s2 + 22 2 s + 2 2

(43)

By truncating the model, the order of the resulting control law is


reduced by eight states. When including the additional modes,
the control law could no longer be solved in real time with the
chosen sampling frequency and numerical integration scheme.
The aforementioned model is augmented with an integrator
in order to reduce the sensitivity to plant uncertainty at lower
frequencies. In addition, the reconstruction and antialiasing filters Wr (s) and Wa (s) are incorporated into the control design.
The filters provide an additional degree of freedom when tuning the control law and can be used to attenuate nonmodeled
high-frequency dynamics, as well as to attenuate quantization
and sensor noise. The complete plant model is thus taken to be
d (s)Wr (s)Wa (s)
p (s) = 1 G
G
s

(44)

p (s),
and is of seventh order. In addition to the plant model G
there are two additional design choices with regards to the

EIELSEN et al.: DAMPING AND TRACKING CONTROL SCHEMES FOR NANOPOSITIONING

TABLE VII
CONTROL LAW PARAMETERS FOR MRC

TABLE VIII
BANDWIDTH OF E(s), N (s)2 FROM n d TO d, D(s) FROM w TO d, 1/,
RMSE, AND ME OBTAINED FOR THE CONTROL SCHEMES

control law: the reference model Wm (s) and the output filter
1/(s). The main limiting factor in determining these filters is
the uncertainty of the plant model, which for the system at hand
is mostly due to nonmodeled high-frequency dynamics.
For simplicity, the reference model Wm (s) is chosen to be a
seventh-order Butterworth filter with cutoff frequency m . Since
p (s) does not have any zeros, (s) should be a
the plant model G
polynomial of sixth order. The zeros of (s) were chosen to have
a Butterworth pattern with radius l . The reconstruction and
antialiasing filters have a user-programmable cutoff frequency
c , which can be tuned, given that c is below the Nyquist
frequency.
The design problem is as such reduced to three tunable control
law parameters
c = [c , m , l ]T
the cutoff frequencies c , m , l . The reference model is a
Butterworth filter, which is maximally flat by definition. The
optimality criterion described in (14) is, therefore, satisfied for
all m , l , and c that render a stable closed-loop system. The
cutoff frequencies were, therefore, tuned manually, attempting
to obtain the highest bandwidth for E(s) while still having a
robustly stable closed-loop system. The control law parameters
used are found in Table VII.
The actual implementation of the scheme in the Appendix is
done by augmenting the filter (49) by an integrator, i.e.,
1

C(s)
= C(s)
s

(45)

and using the filter F (s) in (50) as it is. The parameters for the

filters C(s)
and F (s) are found using (51) and (52), as given in
the Appendix.
With regard to the general control structure in Fig. 7, the analysis in terms of the sensitivity (7), complementary sensitivity
(8), noise attenuation (9), error attenuation (10), and disturbance
rejection (12) is done using
Gp (s) = Gd (s)

441

(46)

C(s) = C(s)W
r (s)

(47)

F (s) = F (s)Wa (s).

(48)

and

The resulting nominal frequency responses for T (s), E(s), and


D(s) are shown in Fig. 12(f).
V. EXPERIMENTAL RESULTS AND DISCUSSION
The six different control schemes were implemented on the
HIL system, described in Section II, and the tracking perfor-

mance when using a triangle wave reference signal with a fundamental frequency of 80 Hz and an amplitude of 1 m was
recorded for each scheme. The fundamental frequency of the
reference signal was chosen in order for the 21st harmonic of
the signal to be close to the dominant vibration mode. The
displacement for all the schemes was measured on a separate
channel using an antialiasing filter with a 35-kHz cutoff frequency. The generated current from the force transducer was
measured, and integrated numerically. The cutoff frequency for
the antialiasing filter for this measurement was always 20 kHz
for all the experiments.
A. Results and Discussion
Nominal frequency responses for the various schemes are
found in Fig. 12(a)(f). The measures from Section III are summarized in Table VIII. Note that the values for 1/ are not
directly comparable between SISO and SIMO systems.
The results when tracking a triangle-wave reference signal
are presented in Fig. 13. The maximum error (ME) ranges from
15% to 24%, and the root-mean-square error (RMSE) ranges
from 0.11 to 0.20 m. The error values are also summarized
in Table VIII. Note that tracking performance can be increased
by adding feed-forward, but this is not done in order for the
error signals to be significantly larger than the noise in the
measured displacement signal to avoid obfuscating the actual
results achieved due to feedback.
The best performing control schemes in terms of the error
are the scheme using IFF and the MRC scheme. The worst performance is obtained when using PPF and the damping integral
control (DI) scheme, while when using IRC and PSD, errors in
the middle of the range are obtained.
The error figures in terms of ME and RMSE can be changed
by the control law tuning, but a reduction in RMSE typically
leads to an increase in ME, due to a more oscillatory response.
The MRC scheme is the most complex scheme. It requires
the implementation of a sixth-order and a seventh-order filter,
a total of 13 integrators. By comparison, the IFF-based scheme
only requires three integrators, but with the disadvantage of
reduced range due to the force transducer, and it requires more
instrumentation and good calibration. On the other hand, the
noise performance is superior, due to the extremely low noise
density of the force transducer, although this benefit it lost for
a digital implementation, due to quantization noise and DAC
artifacts, as discussed below.

442

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 2, APRIL 2014

Fig. 14. Time derivative of force measurement, when having a filter with high
cutoff frequency (PPF, IRC, and IFF) versus low cutoff frequency (PSD, DI,
and MRC) in the signal chain before the voltage is applied to the piezoelectric
actuator.

thus, the nonideal DAC behavior is much more noticeable. This


beneficial effect can also be achieved when using PPF and IRC
schemes by implementing the damping control law Cd (s) using
analog components, as it takes the form of a low-pass filter in
either case, but implementing the whole scheme using analog
components by adding an analog integrator might then be a
better option.
Overall, the performance is fairly similar among the five
schemes, but the excellent nominal noise performance of the
IFF-based scheme and the simplicity of the DI scheme is noteworthy.
VI. CONCLUSION
Fig. 13. Steady-state tracking performance when applying a 1-m amplitude,
80-Hz triangle-wave reference signal. (a) Displacement measurement signals
and reference signal. (b) Error signals.

The simplest control schemes to implement on a digital platform are the PSD based and DI schemes, as they only require
one integrator. The DI scheme is the simplest with regard to extra instrumentation, as it is not necessary to add a shunt circuit,
although for the PSD-based control law, the antialiasing filter is
not necessary and can be omitted. For an analog implementation,
the DI scheme and the schemes based on PPF, IRC, and PSD
are almost equivalent in terms of complexity. The MRC scheme
is likely too complex for an efficient analog implementation.
As quantization noise is the dominant noise source in the
experimental system, it is not possible to obtain reliable closedloop noise measurements. However, due to the low noise and
high sensitivity of the force transducer, the effect of quantization noise and DAC artifacts can be measured. An example of
this is shown in Fig. 14, where the time derivative of the force
measurement is shown when using the IFF-based scheme and
the MRC scheme. The MRC scheme, as well as the PSD and
DI scheme, has a low-pass filter with a low cutoff frequency
before the voltage is applied to the piezoelectric actuator, and
so the noise and disturbances coming from the DAC are effectively attenuated. For the PPF-, IRC-, and IFF-based schemes,
the reconstruction filter has a cutoff frequency of 20 kHz, and

Six fixed-structure, low-order control schemes for damping


and tracking control for a nanopositioning device have been
presented, experimentally assessed, and compared with regard
to performance. Investigated schemes were based on PPF, IRC,
IFF, and PSD, in addition to DI and MRC. This paper furthermore presented a practical and systematic tuning method for the
DI scheme, and the schemes based on PPF, IRC, IFF, and PSD.
Overall, the performance was fairly similar among the
schemes, but features of notice is the noise performance of the
IFF-based scheme and the simplicity of the DI scheme. It was
also demonstrated that when implementing control schemes on
a digital platform, it is beneficial to use control schemes that
reduce the effect of quantization noise and DAC artifacts by using low-pass filers with low cutoff frequencies before the input
to the actuator. Of the schemes investigated, this is most easily
done using the MRC and DI scheme, as well as the PSD-based
scheme. It was also demonstrated that the noise benefits when
using the IFF-based scheme are lost for a digital implementation, due to quantization noise and DAC artifacts.
APPENDIX
MRC [19] is summarized as follows.
The plant model can be expressed as
yp
Zp (s)
.
(s) = Gp (s) = kp
up
Rp (s)

EIELSEN et al.: DAMPING AND TRACKING CONTROL SCHEMES FOR NANOPOSITIONING

It is assumed that Rp (s) and Zp (s) are monic polynomials.


The polynomial Zp (s) is also Hurwitz and is of degree mp .
In addition, the degree np of Rp (s), the relative degree n =
np mp of Gp (s), and the sign of the high-frequency gain kp
are known.
The reference model
Zm (s)
ym
(s) = Wm (s) = km
r
Rm (s)
consists of the monic Hurwitz polynomials Zm (s) and Rm (s)
of degrees qm and pm , respectively, where pm = np , and the
relative degree nm = pm qm = n .
The MRC control law is

F (s)yp )
up = C(s)(r

where C(s)
is a feed-forward filter given as
c0 (s)
(s) 1 T (s)

C(s)
=

(49)

and F (s) is a feedback filter given as


2 T (s) + 3 (s)
F (s) =
.
c0 (s)

(50)

Here,
(s) = [sn p 2 , sn p 3 , . . . , s, 1]T ,

for np 2

(s) = 0,

for np = 1

and
(s) = 0 (s)Zm (s)
is a monic and Hurwitz polynomial of degree np 1. Thus,
0 (s) is a monic and Hurwitz polynomial of degree n0 = np
1 qm . The control law parameter vector is
T

c = [1 T , 2 T , 3 , c0 ]

and should be chosen such that the closed-loop complementary


sensitivity function matches the reference model, i.e.,
yp
ym
(s) = T (s) = Wm (s) =
(s).
r
r
This can be done by choosing
c0 = km /kp

(51)

and solving the Bezout identity


1 T Rp + kp Zp (2 T + 3 ) = Rp Zp 0 Rm
which again can be expressed as
S c = p
T

(52)

where c = [1 T , 2 T , 3 ] and S is a (2np 1) (2np 1)


Sylvester matrix that depends on the coefficients of the polynomials Rp , kp Zp , and , and p is a (2np 1) vector with the
coefficients of the polynomial Rp Zp 0 Rm .
Having learnt the plant parameters of Gp (s), and chosen a
reference model Wm (s) and an output filter 1/(s), the control law parameters c are determined by using the parameter
mapping c defined by (51) and (52).

443

REFERENCES
[1] S. M. Salapaka and M. V. Salapaka, Scanning probe microscopy, IEEE
Control Syst. Mag., vol. 28, no. 2, pp. 6583, Apr. 2008.
[2] G. M. Clayton, S. Tien, K. K. Leang, Q. Zou, and S. Devasia, A review of
feedforward control approaches in nanopositioning for high-speed SPM,
Trans. ASME, J. Dyn. Syst., Meas., Control, vol. 131, no. 6, pp. 0611011061101-19, 2009. DOI: 10.1115/1.4000158.
[3] S. Devasia, E. Eleftheriou, and S. O. R. Moheimani, A survey of control
issues in nanopositioning, IEEE Trans. Control Syst. Technol., vol. 15,
no. 5, pp. 802823, Sep. 2007.
[4] G. Bertotti and I. D. Mayergoyz, The Science of Hysteresi, (ser. Mathematical modeling and applications), vol. 3, 1st ed. New York, NY, USA:
Academic, 2005.
[5] D. A. Apte and R. Ganguli, Influence of temperature and high electric field on power consumption by piezoelectric actuated integrated
structure, Comput. Materials Continua, vol. 10, no. 2, pp. 139161,
2009.
[6] S. R. Viswamurthy and R. Ganguli, Modeling and compensation of piezoceramic actuator hysteresis for helicopter vibration control, Sens Actuators A, Phys., vol. 135, no. 2, pp. 801810, 2007.
[7] S. Bashash and N. Jalili, Robust multiple frequency trajectory tracking control of piezoelectrically driven micro/nanopositioning systems,
IEEE Trans. Control Syst. Technol., vol. 15, no. 5, pp. 867878, Sep.
2007.
[8] J. A. Butterworth, L. Y. Pao, and D. Y. Abramovitch, A comparison of
control architectures for atomic force microscopes, Asian J. Control.,
vol. 11, no. 2, pp. 175181, 2009.
[9] J. L. Fanson and T. Caughey, Positive position feedback-control for large
space structures, AIAA J., vol. 28, no. 4, pp. 717724, 1990.
[10] A. Preumont, J. Dufour, and C. Malekian, Active damping by a local
force feedback with piezoelectric actuators, J. Guid. Control Dynam.,
vol. 15, no. 2, pp. 390395, 1992.
[11] N. W. Hagood and A. von Flotow, Damping of structural vibrations with
piezoelectric materials and passive electrical networks, J. Sound Vib.,
vol. 146, no. 2, pp. 243268, 1991.
[12] H. Pota, S. O. R. Moheimani, and M. Smith, Resonant controllers for
smart structures, Smart Mater. Struct., vol. 11, no. 1, pp. 18, 2002.
[13] S. S. Aphale, A. J. Fleming, and S. O. R. Moheimani, Integral resonant
control of collocated smart structures, Smart Mater. Struct., vol. 16, no. 2,
pp. 439446, 2007.
[14] S. S. Aphale, B. Bhikkaji, and S. O. R. Moheimani, Minimizing scanning
errors in piezoelectric stack-actuated nanopositioning platforms, IEEE
Trans. Nanotechnol., vol. 7, no. 1, pp. 7990, Jan. 2008.
[15] A. J. Fleming, S. Aphale, and S. O. R. Moheimani, A new method for
robust damping and tracking control of scanning probe microscope positioning stages, IEEE Trans. Nanotechnol., vol. 9, no. 4, pp. 438448, Jul.
2010.
[16] A. J. Fleming, Nanopositioning system with force feedback for highperformance tracking and vibration control, IEEE/ASME Trans. Mechatronics, vol. 15, no. 3, pp. 433447, Jun. 2010.
[17] S. Skogestad and I. Postlethwaite, Multivariable Feedback Control: Analysis and Design. New York, NY, USA: Wiley-Interscience, 2005.
[18] G. C. Goodwin, S. F. Graebe, and M. E. Salgado, Control System Design.
Englewood Cliffs, NJ, USA: PrenticeHall, 2000.
[19] P. A. Ioannou and J. Sun, Robust Adaptive Control. Englewood Cliffs,
NJ, USA: PrenticeHall, 1995.
[20] G. Schitter, A. Stemmer, and F. Allgower, Robust two-degree-of-freedom
control of an atomic force microscope, Asian J. Control, vol. 6, pp. 156
163, Jun. 2004.
[21] A. Sebastian and S. M. Salapaka, Design methodologies for robust nanopositioning, IEEE Trans. Control Syst. Technol., vol. 13, no. 6, pp. 868
876, Nov. 2005.
[22] C. Lee and S. M. Salapaka, Robust broadband nanopositioning: fundamental trade-offs, analysis, and design in a two-degree-of-freedom control
framework, Nanotechnology, vol. 20, pp. 035501-1035501-16, 2009.
DOI: 10.1088/0957-4484/20/3/035501.
[23] Y. K. Yong, K. Liu, and S. O. R. Moheimani, Reducing cross-coupling
in a compliant XY nanopositioner for fast and accurate raster scanning,
IEEE Trans. Control Syst. Technol., vol. 18, no. 5, pp. 11721179, Sep.
2010.
[24] H. C. Liaw, B. Shirinzadeh, and J. Smith, Robust motion tracking control
of piezo-driven flexure-based four-bar mechanism for micro/nano manipulation, Mechatronics, vol. 18, no. 2, pp. 111120, 2008.
[25] R. G. Lyons, Understanding Digital Signal Processing, 3rd ed. Englewood Cliffs, NJ, USA: PrenticeHall, 2010.

444

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 19, NO. 2, APRIL 2014

[26] V. Syrmos, C. Abdallah, P. Dorato, and K. Grigoriadis, Static output


feedbackA survey, Automatica, vol. 33, no. 2, pp. 125137, 1997.
[27] S. Ibaraki and M. Tomizuka, Tuning of a hard disk drive servo controller
using fixed-structure H controller optimization, Trans. ASME, J. Dyn.
Syst. Meas., Control, vol. 123, no. 3, pp. 544560, 2001.
[28] A. A. Eielsen, M. Burger, J. T. Gravdahl, and K. Y. Pettersen, PI2controller applied to a piezoelectric nanopositioner using conditional integrators and optimal tuning, in Proc. 18th IFAC World Congr., Milan,
Italy, 2011, pp. 887892.
[29] K. K. Leang and A. J. Fleming, High-speed serial-kinematic SPM scanner: Design and drive considerations, Asian J. Control, vol. 11, no. 2,
pp. 144153, 2009.
[30] A. Preumont, Mechatronics. New York, NY, USA: Springer-Verlag,
2006.
[31] C. V. Newcomb and I. Flinn, Improving the linearity of piezoelectric
ceramic actuators, Electron. Lett., vol. 18, no. 11, pp. 442444, 1982.
[32] Y. Stepanenko and C.-Y. Su, Intelligent control of piezoelectric actuators, in Proc. 37th IEEE Decis. Control, Tampa, FL, USA, Dec. 1998,
vol. 4, pp. 42344239.
[33] A. Preumont, Vibration Control of Active Structures: An Introduction, 2nd
ed. Boston, MA, USA: Kluwer, 2002.
[34] I. Petersen and A. Lanzon, Feedback control of negative-imaginary systems, IEEE Control Syst. Mag., vol. 30, no. 5, pp. 5472, Oct. 2010.
[35] A. A. Eielsen and A. J. Fleming, Passive shunt damping of a piezoelectric
stack nanopositioner, in Proc. Amer. Control Conf., Baltimore, MD, USA,
2010, pp. 49634968.
[36] W. P. Heath and S. Gayadeen, Simple robustness measures for control of
MISO and SIMO plants, in Proc. 18th IFAC World Congr., Milan, Italy,
2011, pp. 1135611361.
[37] S. Boyd and C. Barratt, Linear Controller Design: Limits of Performance.
Englewood Cliffs, NJ, USA: PrenticeHall, 1991.
[38] J. Lagarias, J. Reeds, M. Wright, and P. Wright, Convergence properties
of the Nelder-Mead simplex method in low dimensions, SIAM J. Optimiz.,
vol. 9, no. 1, pp. 112147, 1998.
[39] A. J. Fleming and K. Leang, Integrated strain and force feedback for high
performance control of piezoelectric actuators, Sens. Actuators A, Phys.,
vol. 161, nos. 1/2, pp. 256265, 2010.
[40] S. S. Aphale, S. Devasia, and S. O. R. Moheimani, High-bandwidth control of a piezoelectric nanopositioning stage in the presence of plant uncertainties, Nanotechnology, vol. 19, pp. 125503-1125503-9, 2008. DOI:
10.1088/0957-4484/19/12/125503.
Arnfinn A. Eielsen received the M.Sc. and Ph.D. degrees in engineering cybernetics from the Norwegian
University of Science and Technology, Trondheim,
Norway, in 2007 and 2012, respectively.
He is currently a Postdoctoral Researcher in the
Department of Engineering Cybernetics, Norwegian
University of Science and Technology. His research
interests include motion control, estimation and identification, physical modeling, and instrumentation.

Marialena Vagia received the B.Sc. and Ph.D. degrees from the University of Patras, Greece, in 2005
and 2009, respectively.
She is currently a Postdoctoral Researcher at the
Norwegian University of Science and Technology
(NTNU), Trondheim, Norway, working in the area
of robotics. Previously, she was awarded with postdoctoral scholarships such as the one funded by the
ERCIM Marie Curie Network completed at NTNU,
the postdoctoral scholarship from the Greek State
Scholarships Foundation (IKY) completed at the
University of Patras, and the Post Doctoral Researchers Scholarships funded
by the EU and General Secretariat Research and Technology, Greece. She has
also served as an Adjunct Lecturer at the University of Patras. Her research
interests include robust and nonlinear control, switching control, electrostatic
microactuators, microcantilever beams, nanostructures, micropiezo actuators,
nanopositioners, and robotic manipulators. She has published ten journal papers and 20 conference papers. She has also served as a reviewer for many
international conferences and journals.
Dr. Vagia received the Best Student Paper Award at the IEEE Mediterranean
Conference on Control and Automation in 2007.

J. Tommy Gravdahl received the Siv.ing and Dr.ing


degrees in engineering cybernetics from the Norwegian University of Science and Technology (NTNU),
Trondheim, Norway, in 1994 and 1998, respectively.
He became an Associate Professor in 2001 and
then Professor in 2005 in the Department of Engineering Cybernetics, NTNU, where he also served as
the Head of the Department in 20082009. In 2007
2008, he was a Visiting Professor at the Centre for
Complex Dynamic Systems and Control, The University of Newcastle, Newcastle, Australia. He has
published more than 100 international conference and journal papers. He is
the author of Compressor Surge and Rotating Stall: Modeling and Control
(Springer, 1999), coauthor of Modeling and Simulation for Automatic Control
(Marine Cybernetics, 2002), coeditor of Group Coordination and Cooperative
Control (Springer, 2006), and coauthor of Snake Robots: Modelling, Mechatronics, and Control (Springer, 2013). He is also a coauthor of Modeling and
Control of Vehicle-Manipulator Systems (Springer, 2013). His current research
interests include mathematical modeling and nonlinear control in general, modeling and control of turbomachinery, and control of vehicles, spacecraft, robots,
and nanopositioning devices.
Prof. Gravdahl received the IEEE TRANSACTIONS ON CONTROL SYSTEMS
TECHNOLOGY Outstanding Paper Award in 2000.

Kristin Y. Pettersen was born in 1969. She received


the M.Sc. and Ph.D. degrees in engineering cybernetics from the Norwegian University of Science and
Technology (NTNU), Trondheim, Norway, in 1992
and 1996, respectively.
She is currently a Professor in the Department of
Engineering Cybernetics, NTNU, where she has been
a Faculty Member since 1996. She is the Head of the
Department, and a Director of the ICT Programme
of Robotics which is one of the universitys strategic
research areas. From 2013 to 2022, she was also a
key Scientist at the CoE Centre for Autonomous Marine Operations and Systems. She has published more than 140 international papers for conferences and
journals. Her research interests include nonlinear control of mechanical systems
with applications to robotics, satellites, AUVs, and ships. She edited Group Coordination and Cooperative Control (Springer-Verlag, 2006) and is a coauthor
of Snake Robots (Springer-Verlag, 2013). She is also a coauthor of Modeling
and Control of Vehicle-Manipulator Systems (Springer-Verlag, 2013). In 2008,
she was a Visiting Professor at the Section for Automation and Control, Aalborg
University, Aalborg, Denmark, and in 1999, she was a Visiting Fellow with the
Department of Mechanical and Aerospace Engineering, Princeton University
Princeton, NJ, USA.
Dr. Pettersen, with her coauthors, received the IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY Outstanding Paper Award for Global Uniform
Asymptotic Stabilization of an Underactuated Surface Vessel: Experimental
Results (K.Y. Pettersen, F. Mazenc, and H. Nijmeijer) in 2006. She has served
as an Associate Editor for several conferences, including IEEE Conference
on Decision and Control, IEEE Conference on Robotics and Automation, and
IEEE/RSJ International Conference on Intelligent Robots and Systems. She has
served as a member of the Editorial Board of Simulation Modeling Practice
and Theory, and is an Associate Editor of the IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY and the IEEE Control Systems Magazine. She is
a member of the Board of Governors of the IEEE Control Systems Society. She
also holds several Board positions in industrial and research companies.

Anda mungkin juga menyukai