You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/260332220

A Waveform Design Method for a Piezo Inkjet Printhead Based on Robust


Feedforward Control

Article in Journal of Microelectromechanical Systems · December 2012


DOI: 10.1109/JMEMS.2012.2205899

CITATIONS READS
32 3,873

6 authors, including:

Amol Khalate Gérard Scorletti


ASML Ecole Centrale de Lyon
23 PUBLICATIONS 199 CITATIONS 167 PUBLICATIONS 2,431 CITATIONS

SEE PROFILE SEE PROFILE

Robert Babuska
Delft University of Technology
474 PUBLICATIONS 17,317 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Control Methods for Walking Robots View project

A piecewise-affine approach to nonlinear performance View project

All content following this page was uploaded by Gérard Scorletti on 11 February 2015.

The user has requested enhancement of the downloaded file.


DRAFT SUBMITTED TO: IEEE JOURNAL OF MICROELECTROMECHANICAL SYSTEMS 1

A Waveform Design Method for a Piezo Inkjet


Printhead based on Robust Feedforward Control
Amol A. Khalate, Xavier Bombois, Gérard Scorletti, Robert Babuška, Sjirk Koekebakker, and Wim de Zeeuw

Abstract—The printing quality delivered by a Drop-on- posed by new applications, these drop properties have to be
Demand (DoD) piezo inkjet printhead is limited mainly due tightly controlled.
to the residual oscillations in the ink channel. The maximal
The performance of the inkjet printhead is mainly limited
jetting frequency of a DoD inkjet printhead can be increased
by quickly damping the residual oscillations and so bringing an due to the residual pressure oscillations. The actuation pulses
ink-channel to rest after jetting an ink drop. In this paper, we are designed to provide an ink drop of a specified volume and
propose a robust optimization-based method to design the input velocity under the assumption that the ink channel is in steady
actuation waveform for the piezo actuator in order to improve the state. Once the ink drop is jetted, the pressure oscillations
damping of the residual oscillations in the presence of parametric
inside the ink channel take several micro-seconds to decay.
uncertainties in the ink-channel model. The proposed method
obtains a robust actuation pulse by minimizing the tracking error If the next ink drop is jetted before the residual pressure
under the parametric uncertainty. Experimental results with a oscillations have settled, the resulting drop properties will be
small droplet DoD inkjet printhead are presented to show the different from the ones of the previous drop. This can degrade
efficacy of the proposed method and the significant improvement the printhead performance, since a printhead has to jet drops
in the ink drop consistency.
with a constant velocity at different frequencies. Given this
fact, an important characteristic is the so-called DoD curve
which represents the ink drop velocity as a function of the
I. I NTRODUCTION
jetting frequency (which is also called the DoD frequency).
The ability of inkjet technology to deposit materials with Ideally, the DoD curve must be flat. However, for the above
diverse chemical and physical properties has made it an reasons, this DoD curve is far from flat in practice. Our goal
important technology for both industry and home use. Apart in this paper is to flatten the DoD curve by redesigning the
from conventional document printing, the inkjet technology piezo actuation pulse.
has been successfully applied in the areas of electronics, In the literature, we can find methods to redesign the
mechanical engineering and life sciences [1]. This is mainly piezo actuation pulse based on exhaustive experimental studies
thanks to the low operational costs of the technology. Recently, [6]–[10] and experimental analysis [11], [12]. In [13], we
inkjet printheads are getting closely associated with the micro- introduced a model-based approach to design the actuation
electromechanical systems (MEMS) technology. Inkjet print- pulse for the DoD inkjet printhead. The model we have used is
heads are now being used to fabricate MEMS devices [2], a discrete-time transfer function H(q) relating the piezo input
[3]. Inkjet printheads are also miniaturized to MEMS level voltage (i.e., the input u) to the velocity of the meniscus1 (i.e.,
[4]. A typical drop-on-demand (DoD) piezo inkjet printhead the output y). We consider this particular model since it is well
consists of several ink channels in parallel. Each channel is known that the velocity of the meniscus is a good measure
provided with a piezo-actuator, which on application of a of the pressure in the ink channel [6], [14]. Consequently,
voltage pulse generates pressure oscillations inside the ink reducing the residual oscillations of the meniscus velocity is
channel. These pressure oscillations push the ink drop out equivalent to reducing the residual pressure oscillations in the
of the nozzle. A detailed description of the droplet jetting ink channel [5].
process can be found in [5]. The print quality delivered by In order to design the piezo actuation pulse with this transfer
an inkjet printhead depends on the properties of the jetted function H(q), we have parameterized the set of actuation
drop, i.e., the drop velocity, the jetting direction and the drop pulses u(k, θ) that the driving electronics can generate with
volume. To meet the challenging performance requirements the pulse parameter vector θ (k is the discrete-time index).
Then a template yref (k) is designed for the desired meniscus
This work has been carried out as part of the Octopus project with Océ
Technologies B.V. under the responsibility of the Embedded Systems Institute. velocity, i.e., a meniscus velocity profile with fast decaying
This project is partially supported by the Netherlands Ministry of Economic residual oscillations. Based on this template yref (k) and the
Affairs under the Bsik program P90600. transfer function H(q) an optimal actuation pulse u(k, θopt ) is
Amol A. Khalate, Xavier Bombois, Robert Babuška are
Delft Center for Systems and Control, Delft University of determined as the one minimizing the tracking error e(k) =
Technology, Mekelweg 2, 2628 CD Delft, The Netherlands. e-mail: yref (k) − y(k). Experimental results have shown [13] that the
({a.a.khalate,x.j.a.bombois,r.babuska}@tudelft.nl). optimal actuation pulse u(k, θopt ) designed in this manner
Gérard Scorletti is with Laboratoire Ampère, Ecole Centrale de Lyon,
Ecully, France. improves the performance of the inkjet printhead compared to
Sjirk Koekebakker and Wim de Zeeuw are with Océ Technologies B.V., the standard actuation pulse (the pulse that is generally used
5900 MA Venlo, The Netherlands.
1 The meniscus is an interface between the ink and air.
DRAFT SUBMITTED TO: IEEE JOURNAL OF MICROELECTROMECHANICAL SYSTEMS 2

for the printhead). However, the DoD-curve with the optimal used ink is 10 · 10−3 Pa·s and the surface tension of the ink
pulse u(k, θopt ) is not completely flat. is 28·10−3 Nm−1 . The speed of sound in the ink is 1250ms−1 .
The main reason for this is that the model H(q) used for
the design of the optimal pulse only represents the dynamics
of the system when a single drop is jetted. The model H(q) B. Parametrization of the actuation pulse
is therefore not representative of the dynamics of the system
The driving electronics of the printhead limit the range of
when jetting a series of drops at a certain DoD frequency.
the actuation pulses, that can be generated in practice, to
Data-based identification of the system dynamics furthermore
trapezoidal waveforms consisting of one or two trapezoidal
suggests that the dynamics from the piezo input to the menis-
pulses. The standard pulse ustd (k) (i.e. the one that is currently
cus velocity depend on the DoD frequency.
used in printheads) consists of only one positive trapezoidal
In this paper, in order to take into account this model
pulse (called the resonating pulse). This pulse is able to jet
variation, we encompass the set of dynamical models obtained
a drop, but it generates residual oscillations as mentioned in
at various operating DoD frequencies in a compact uncertainty
the introduction. In order to be able to damp these residual
set H(q, ∆) (∆ ∈ ∆) and we design a robust actuation pulse
oscillations, we choose a parametrization where the resonating
which will ensure a minimum performance for all models in
pulse is followed by a negative trapezoidal pulse (called the
the uncertainty set. The robust actuation pulse is obtained by
quenching pulse). See Fig. 2.
minimizing the worst-case norm of the tracking error e(k) with
The parameters that are optimized are the rise time (tr ),
the uncertain inkjet system H(q, ∆). As we will see at the end
the dwell time (tw ), the fall time (tf ) and the amplitude (V )
of this paper experimental results show that this robust pulse
of both the resonating (R) and the quenching pulse (Q). The
generates a DoD-curve which is much flatter than when using
time interval between the resonating pulse and the quenching
the optimal pulse (and the standard pulse).
pulse is tdQ . Thus, an actuation pulse u(k, θ) is defined by the
parameter vector θ = [trR twR tfR VR tdQ trQ twQ tfQ VQ ]T .
II. OVERVIEW OF THE O PTIMIZATION -BASED Note that The amplitude of the pulse is expressed in Volts.
F EEDFORWARD C ONTROL D ESIGN The time parameters in θ are number of samples. However, for
We first present a brief summary of our model-based simplicity, we express these parameters in µ s by multiplying
approach [13] using only the nominal model H(q) in order to the number of samples with the sampling time Ts .
motivate the need for the robust pulse design. Another reason u
for this summary is that some elements will also be necessary
for our robust design. This is for example the case for the twR
parametrization of the actuation pulse.

VR
A. System description
trQ tfQ
Flexible foil Piezo unit trR tfR td Q time
VQ
Reservoir

Substrate

Droplet

Channel plate
twQ
Filter Channel Nozzle Fig. 2. Proposed actuation pulse.

Fig. 1. A cross-sectional view of an ink channel.

We use a small droplet DoD inkjet printhead built by Océ C. Modeling and the desired meniscus velocity
Technologies to demonstrate the proposed actuation pulse As mentioned in the introduction, the results in [13] are
design method. The DoD inkjet printhead under investigation based on a linear model H(q) relating the pulse u(k, θ) to the
is made up of two arrays of 128 ink channels each. A meniscus velocity y(k). The model H(q) is derived directly
cross-sectional view of an ink channel is shown in Figure 1. from the so-called narrow-gap model [15].
The ink channel is carved in the channel plate. A filter is By construction, this model describes, almost perfectly, the
placed before the ink channel to remove impurities from the relation between u(k) and y(k) when one single ink drop is
liquid ink. A metallic nozzle-plate with drilled holes, which jetted from an ink channel which was at rest.
act as nozzles, is attached at the end of the channel plate. In [13], the model was used for two purposes. The first
One wall of the ink channel is formed by a flexible foil to purpose is to determine the desired meniscus velocity template
which a piezo unit is attached. The piezo unit acts as an yref (k) and the second one is to compute the optimal pulse
actuator and on the application of a voltage, it deforms the uopt (k) = u(k, θopt ).
wall of the ink channel. The deformation generates pressure To obtain yref (k), we determine the response of the model
waves inside the ink channel and when specific conditions H(q) when the standard pulse ustd (k) of the considered
are met, a droplet is jetted [5]. Note that the viscosity of the printhead is used as an input. The standard pulse is known
DRAFT SUBMITTED TO: IEEE JOURNAL OF MICROELECTROMECHANICAL SYSTEMS 3

6
and the optimal actuation pulse uopt (k) = u(k, θopt ) is shown
Response to standard pulse ( y ) in Fig. 6. Simulation results show that the optimal pulse
5
Reference meniscus velocity trajectory ( y
ref
)
uopt (k) tracks the reference yref (k) and damps the residual
4
oscillations almost perfectly (see [13]).
Meniscus velocity (m/s)

2
E. Experimental results with the optimal pulse
1 A B
To validate our approach, the performance of the optimal
pulse uopt (k) has been compared with the one of the standard
0
pulse ustd (k) on an experimental setup. The experimental test
−1 consists of jetting 10 ink drops from the inkjet channel at
−2
a fixed DoD frequency. This experiment was repeated for
different DoD frequencies ranging from 20kHz to 70kHz.
−3
0 10 20 30 40 50
Time (µs)
60 70 80 90 100 For evaluating the performance we can plot the drop veloc-
ities of the 10 drops as a function of the DoD frequency, also
called as the DoD curve (see Fig. 13 and 14 in [13]). In [13],
Fig. 3. Reference meniscus velocity trajectory.
we found that the optimal pulse leads to a DoD-curve which
is much flatter than the DoD curve with the standard pulse.
For the standard pulse, the maximum drop velocity variation
to jet a droplet with the desired properties when the ink over the DoD frequency range is 12ms−1 while with uopt (k)
channel is at rest. The standard pulse for the inkjet print- it is reduced to 4.5ms−1 . The optimal pulse had also reduced
head considered in this paper is parameterized by the vector the drop velocity variation at a given DoD frequency.
θstd = [1.5 2.5 1.5 25 0 0 0 0 0]T (i.e. ustd (k) = u(k, θstd )) Besides the drop velocity, another important property is the
and is represented in Fig. 6. The response y(k) of H(q) to drop volume. The relationship between velocity and volume
this pulse has two parts as shown in Fig. 3 (dashed line). Part can be altered for different actuation waveforms. However,
A of the response allows the drop to be jetted at the desired this relationship remains approximately linear for the printhead
drop velocity: the template yref (k) for the meniscus velocity considered in this work (see Fig.45 of [15]). In other words, if
should therefore be the same as y(k) in Part A. Part B of the the actuation pulse flattens the drop velocity DoD curve then
response y(k) represents the residual oscillations. This is an it will also flatten the drop volume DoD curve. Therefore, to
undesired behavior since the residual oscillations perturb the evaluate the performance of the actuation pulse, plotting the
subsequent drops. Therefore, in Part B, we force the desired drop velocity as a function of the DoD frequency is sufficient.
meniscus velocity yref (k) to zero. Remark:
If the actuation pulse is designed in such a way that the By adding an extra quenching pulse, the duration of the
meniscus velocity y(k) follows the reference trajectory yref (k), optimal pulse will always be longer than the standard pulse.
then the channel will come to rest very quickly after jetting the The duration of the standard pulse is 5.5 µs while the one of
ink drop. This will create the conditions to jet the ink drops the optimal pulse (u(k, θopt )) is 19.5 µs. A consequence of the
at higher DoD frequencies. longer duration of the pulse is that, within the range [0 70kHz]
for the DoD frequency, the optimal pulses will overlap from
D. Optimal Actuation Pulse Design a DoD frequency (1/19.5 µs)= 51 kHz. This does not happen
In [13], we define the optimal actuation pulse as the trape- for the standard pulse. For fDoD > 51 kHz, we have decided
zoidal input u(k, θ) which minimizes the difference between to superimpose the overlapping pulses. This means that, if we
the reference trajectory yref (k) and the meniscus velocity want to jet a series of ND drops at a DoD frequency fDoD
y(k) = H(q)u(k, θ). More precisely, the parameter vector θopt larger than 51 kHz, we in fact apply the following voltage
describing the optimal pulse is the one solving the following uoverlap (k) to the piezo unit
(nonlinear) optimization problem NX
D −1  
uoverlap (k) = uopt k − iDk
θopt = arg min Jnom (θ), subject to θLB ≤ θ ≤ θU B , (1)
θ i=0

with where uopt (k) = u(k, θopt ) the optimal pulse of duration
N  2
X 19.5 µs, and Dk is the number of samples between two
Jnom (θ) = yref (k) − H(q)u(k, θ)
actuation pulses2 .
k=0
T
where N = Ts is the sampling time, T is chosen equal to
Ts ,
III. I NKJET C HANNEL AS AN U NCERTAIN S YSTEM
100 µs, q is the forward shift operator and θLB and θU B are The experimental DoD curve is not completely flat even
vectors containing the lower and the upper bounds on each though the simulation results predicted an almost complete
element of the parameter vector θ.
2 If T is the sampling time and round(x) rounds a real number x to
The optimal actuation pulse vector θopt obtained by solving s
the nearest integer, Dk is then equal to round( T f1 ). Indeed, at a DoD
the above problem is s DoD
frequency of fDoD , the time separating two successive actuation pulses is f 1
DoD
θopt = [2.0 2.5 1.3 22.5 7.6 1.3 0.4 4.4 − 13.2]T , second and Dk must be an integer since uopt (k) is a discrete-time signal.
DRAFT SUBMITTED TO: IEEE JOURNAL OF MICROELECTROMECHANICAL SYSTEMS 4

attenuation of the residual oscillations. This discrepancy can the above nominal parameters is given by the solid line in
be explained by the difference between the actual dynamics Fig. 4.
of the inkjet channel and the one described by H(q). Indeed,
H(q) is only a representative of the channel dynamics when
a single drop is jetted in a channel at rest. However, to
construct the DoD curve, we jet a series of drops at a certain
DoD frequency. The dynamics are then different than H(q).
Moreover, since we observe in [13] a different result for
different DoD frequencies, we can expect that the behavior
depends on the DoD frequency.
One possible reason for this frequency-dependent behavior
is the fact that the dynamics of the refilling of the ink channel
after a drop has been jetted is not modeled in the narrow-
gap model which is the basis for H(q). By jetting droplets
at different DoD frequencies, these slow refill dynamics may
change the average meniscus position at the time of droplet
separation. This may in turn change the properties of the linear
model H(q) when obtained at different DoD frequencies.
The above observation is confirmed by system identifica-
tion results obtained at different DoD frequencies, which are Fig. 4. Frequency response of the transfer function model Hlow (q).
presented in Appendix A. In these results, we observe that the
main variation of the dynamics of the system (characterized
by two resonance peaks) is a variation of the first resonance The variation of the first resonance peak has now to be
peak. The variation observed in the second peak is neglected parameterized. For this purpose, let us first consider the
since it has a smaller amplitude and since higher frequency following continuous-time version of Hlow (q):
dynamics are known to be of less influence on the ink drop ! !
properties [14]. g1 s(s + α) s2 + 2ζn3 ωn3 s + ωn3
2

In the sequel, our uncertain model will be described by Hlow (s) = 2 2 .


s2 + 2ζn1 ωn1 s + ωn1 s2 + 2ζn2 ωn2 s + ωn2
systems where the frequency of the first resonance peak (3)
varies from 75kHz to 86kHz and its damping ratio varies The variation in the frequency ωn1 and the damping ratio ζn1
from 0.0366 to 0.1585. Since we cannot directly identify the of the first mode is represented in Fig. 5.A.
transfer between the piezo input and the meniscus velocity, The parametric uncertainty in the coefficients ωn1 and ζn1
the size of the uncertainty is considered not only based on the of the continuous-time transfer function is translated into the
system identification results, but also on physical insights. (see uncertainty ∆ on the coefficients a1 and a2 of the discrete-
Appendix A). With respect to the frequency and the damping time transfer function using standard results [16]:
ratio of the first peak of the nominal model H(q) used in
Section 2, the uncertainty set corresponds to a variation in an
a1 = −2r cos θ, a2 = r2 (4)
interval of [−7% +7%] for the frequency and [−70% +30%]
for the damping ratio. The nominal model is thus not in the p
2 and T is the
center of our uncertainty region. where r = e−ζn1 ωn1 Ts , θ = Ts ωn1 1 − ζn1 s
sampling time.
Let us now describe our uncertainty set in more detail. Con-
sider for this purpose the following model structure (sampling Using these relations, the box-type uncertainty in the param-
time Ts = 1µs): eters ωn1 and ζn1 is mapped onto a set ∆ in the parameter
! ! space of the coefficients a1 and a2 as shown in Fig. 5.B. This
q 2 + b 1 q + b2 q 2 + b3 q + b4 is a polytopic uncertainty, which will used to design the robust
Hlow (q) = g 2 . (2) actuation pulse. The four vertices of the box-type uncertainty
q + a1 q + a2 q 2 + a3 q + a4
on ωn1 and ζn1 are mapped on the parameter space of a1
When the parameters are chosen as follows: and a2 . These four vertices, when expressed relative to the
nominal values a1,nom = −1.6480 and a2,nom = 0.8839, are
b1 = −1.2194, b2 = 0.2194, b3 = −1.7170, b4 = 7.0670, ∆1 = [5.93 8.71]T × 10−2 , ∆2 = [1.96 8.25]T × 10−2 ,
a1 = −1.6480, a2 = 0.8839, a3 = −1.0040, a4 = 0.8971, ∆3 = [−3.85 −4.32]T ×10−2 , ∆4 = [0.55 −2.52]T ×10−2 .
g = −0.0074, The convex combination of these four vertices forms the set
∆ =conv(∆1 , ∆2 , ∆3 , ∆4 ).
this model structure allows to represent perfectly the low- In other words, any perturbation in the parameters ωn1
frequency dynamics of the nominal model H(q) used in and ζn1 will bePmapped in perturbation/uncertainty ∆ =
4 P4
Section II-C. Note that the higher order dynamics of H(q) [∆(1) ∆(2) ]T = i=1 αi ∆i with i=1 αi = 1 (i.e. ∆ ∈ ∆)
are neglected since they do not contribute at all in the drop on the coefficients a1 and a2 . These perturbation on the
velocity properties [14]. The Bode diagram of Hlow (q) with coefficients a1 and a2 can be represented in the following
DRAFT SUBMITTED TO: IEEE JOURNAL OF MICROELECTROMECHANICAL SYSTEMS 5

(A)
40
the robust performance index Jrob (θ) as the worst-case sum
of squared tracking errors:

Relative deviation in ζn1 (×10−2 )


20
N 
X 2
0 Jrob (θ) = max yref (k) − Hlow (q, ∆)u(k, θ) . (7)
∆∈∆
k=0
−20
Remark: In (7), the template yref (k) is the same as in (1).
−40 Indeed yref (k) was constructed with the model H(q) in the
−60
condition where H(q) appropriately describes the system
(one single drop jetted at rest) and is thus an appropriate
−80
−8 −6 −4 −2 0 2 4 6 8
description of the desired meniscus velocity.
Relative deviation in ωn1 (×10−2 )
In order to design the robust actuation pulse, it is important
(B) to be able to compute Jrob (θ) for any arbitrary value of θ or at
10
least to be able to compute a good approximation for Jrob (θ).
Relative deviation in a2 (×10−2 )

Obtaining the exact solution of the optimization problem (7)


LB
is nontrivial, but a good approximation Jrob (θ) for Jrob (θ)
5
can be simply determined by griding the uncertainty space ∆.
In our case, as the parameter space of ∆ is only of dimen-
sion two, gridding can be easily and effectively performed.
0
Let the set S be a fine grid on the parametric uncertainty ∆,
defined as
−5 S = {∆i , i = 1, ..., m, | ∆i ∈ ∆}
−4 −2 0 2 4 6
LB
Relative deviation in a1 (×10−2 ) Then a good lower bound approximation Jrob (θ) of the
performance index Jrob (θ) can be obtained as the maximum
Fig. 5. Parametric uncertainty relative to the nominal value. sum squared error over the m grid points, i.e.:
N 
X 2
LB
manner Jrob (θ) = max
i
yref (k) − Hlow (q, ∆i )u(k, θ) . (8)
∆ ∈S
k=0
a1 (∆) = a1,nom (1 + ∆(1) ) (5) Now, the constrained robust actuation pulse parameter is
(2) thus the solution θrobust of the following optimization problem
a2 (∆) = a2,nom (1 + ∆ ), (6)
LB
with a1,nom and a2,nom defined just above. θrobust = arg min Jrob (θ) subject to θLB ≤ θ ≤ θU B , (9)
θ
Now, the set of dynamical models obtained at various DoD
where, θLB and θU B are vectors containing the lower and the
frequencies can thus be represented by the uncertain inkjet
upper bounds on each element of the parameter vector θ.
system Hlow (q, ∆), ∆ ∈ ∆. The frequency response of the
This is a nonlinear optimization problem which can be
uncertain inkjet system, which is represented by Hlow (q, ∆),
solved offline using standard optimization algorithms. Since,
is shown by shaded area in Fig. 4. LB
we can compute Jrob (θ) for each value of θ, we can use
We have seen in Section II-E that uopt (k) does not com-
gradient-based optimization to find θrobust . Gradient-based
pletely flatten the DoD-curve, because it uses only the nominal LB
optimization is an iterative method. The gradient of Jrob (θ)
model H(q) for the actuation pulse design which does not take
is computed numerically around the current value of θ and
into account the multiple dynamical models at various DoD
then the parameter θ is updated in the gradient direction.
frequencies. In the next section, we will see that the robustness
We use the MATLAB function fmincon for this purpose.
of the actuation pulse can be improved if we consider this set
This function implements a range of optimization techniques.
of multiple dynamical models Hlow (q, ∆) with ∆ ∈ ∆ for the
In our experiments we used the default option which is
actuation pulse design.
sequential quadratic programming.

IV. ROBUST F EEDFORWARD C ONTROL D ESIGN Remarks:


In the previous section, we have seen that the multiple mod- • The griding approach to obtain a tight approximation
LB
els obtained at different DoD frequencies can be represented Jrob (θ) for Jrob (θ) may pose numerical problem when
by uncertain inkjet system Hlow (q, ∆), ∆ ∈ ∆. This uncertain the parameter space of ∆ becomes large or when the
set will be used to design a robust actuation pulse, which will dependency of Jrob (θ) on θ become highly nonlinear. In
ensure a minimum performance for all systems in the polytopic such scenario, it is important to obtain some confident
uncertainty rather than obtaining an optimal actuation pulse bounds on Jrob (θ) without increasing numerical com-
whose performance is only good for one single element of plexity. In [17], we have shown that for any arbitrary
UB
this set (as we did in Section II) For this purpose, we define value of θ an upper bound Jrob (θ) for Jrob (θ), i.e.
DRAFT SUBMITTED TO: IEEE JOURNAL OF MICROELECTROMECHANICAL SYSTEMS 6

UB
Jrob (θ) ≤ Jrob (θ), can be obtained as the solution
of a convex optimization problem. However, the de-
sign of the robust constrained actuation pulse remains 7

a nonlinear optimization problem due to the nonlinear


parameterization of u(k, θ) in θ. We have further shown 6

that if we relax the pulse shape constraint, the design


5
of an unconstrained robust actuation pulse is a convex
optimization [17].

Drop Velocity (m/s)


4
• Due to manufacturing tolerances the dynamical behavior
of ink channels inside a printhead may differ. Generally, 3

this is the reason for differences in the performance Drop−1


Drop−2
of inkjet printheads manufactured in a batch. The ink 2 Drop−3
Drop−4
viscosity change caused by temperature change or by Drop−5
Drop−6

slightly different ink can affect the dynamical behavior 1


Drop−7
Drop−8
Drop−9
of a printhead. The proposed method to design the robust Drop−10

pulse is useful to tackle the performance degradation due 0


20 25 30 35 40 45 50
DoD Frequency (kHz)
55 60 65 70

to these manufacturing tolerances.

V. E XPERIMENTAL R ESULTS Fig. 7. Experimental DoD curve with the robust pulse urobust (k).

In this section, we present experimental results to show


the improvements in the drop consistency with the robust
actuation pulse. The robust actuation pulse parameter vector
θrobust obtained after solving the optimization problem (9) is Paper
location

T drop 1,2,3&4
θrobust = [1.5 2.5 1.5 23.0 6.93 2.74 0.4 3.75 −13.22] merged

Note that the time parameters (tr , tw , tf , tdQ ) in the parameter


Position

vector θrobust are expressed in µs.


The robust pulse urobust (k) = u(k, θrobust ) designed in such Satellite drop

a manner is shown in Fig. 6 by the solid blue line.


Nozzle
location

30
Time
Standard pulse ustd (A) Standard pulse
Optimal pulse uopt
25
Robust pulse u
robust
Paper
location
20
drop 1,2&3
merged
15
Input (Volt)

10
Position

0
Satellite drop

−5

−10 Nozzle
location
−15 Time
0 2 4 6 8 10 12 14 16 18 20
Time (µ s) (B) Optimal pulse
Paper
location

Fig. 6. Actuation pulses.


drop 1&2
Position

merged

We have done similar experiments as mentioned in Section


II-E with the robust pulse urobust (k) and obtained the DoD-
curve as shown in Fig. 7. Since the robust pulse urobust (k) is
designed to give a good average performance over the set of Nozzle
location
models represented by H(q, ∆), it delivers better results than Time
the optimal pulse. Now, the drop velocity variation is even less (C) Robust pulse
than 2 ms−1 .
Let us now focus on a single DoD frequency of 48 kHz Fig. 8. Experimental result: 10 drops jetted at DoD frequency 48 kHz.
and let us compare the performance of the three pulses (i.e.,
the standard pulse, the optimal pulse and the robust pulse) at
DRAFT SUBMITTED TO: IEEE JOURNAL OF MICROELECTROMECHANICAL SYSTEMS 7

that frequency. Consider for this purpose the flight profile3 of opinion, introducing the control aspects (such as positioning
ten drops jetted at that frequency. This flight profile, from the of actuators and sensors) as an extra tuning variable in the
nozzle to the paper, is represented in Fig. 8. The vertical axis design process could lead to even better performance and a
represents the position of the ink drop. The starting position is cheaper design.
the nozzle level and the paper is placed at the end position. The
distance between the nozzle and the paper is approximately VII. ACKNOWLEDGEMENT
2mm. The authors gratefully acknowledge fruitful discussions
Ideally, it is required that the 10 drops should be placed at an with H. Wijshoff, R. Waarsing, P. Klerken, and experimental
equal distance on the paper. For the standard pulse (Fig. 8.A), support from J. Simons.
the first four drops are slower and they are merged together
and finally, only seven drops reach the paper. Most of the A PPENDIX A
jetted drops have different velocities. Also, a small and slower VALIDATION USING S YSTEM I DENTIFICATION
satellite drop is generated after the 10th drop. For the optimal It is very difficult to experimentally measure the meniscus
pulse (Fig. 8.B), the first two drops are slower and they are position and the meniscus velocity while jetting an ink drop.
caught by the faster third drop. The remaining seven drops Fortunately, the piezoelectric crystal can be used simultane-
have similar velocities. The satellite drop generated after the ously as an actuator and as a sensor. The measured piezo
10th drop is faster and gets merged in 10th drop. With the sensor signal is proportional to the rate of change of the
robust pulse (Fig. 8.C), except for the first two drops, the drop ink-channel pressure. (See Appendix B for more details on
velocities of the remaining drops are very similar to each other. the measurement of the sensor signal.) Therefore, the model
There is also no satellite drop generated after the 10th drop. G(q) identified using the piezo sensor as an output is different
The overall improvement in the velocity consistency from the model Hlow (q) which relates the pulse u(k) to the
achieved using the optimal piezo actuation pulse has far- meniscus velocity. However, the model G(q) from the piezo
reaching consequences on the print quality. This is because input to the piezo sensor obtained at different fixed DoD
the printing paper is placed at approximately 2mm from the frequencies, give us sufficient insight into the dependency of
inkjet printhead and the relative speed between the paper and Hlow (q) on the DoD frequency, since the meniscus velocity
the printhead is around 1ms−1 . The proposed method reduces depends on the channel pressure.
the drop velocity variation to 2.0ms−1 compared to a drop In order to obtain a model from the piezo input to the
velocity variation equal to 4.5ms−1 in our previous work [13] piezo sensor we have jetted several ink drops at a given
(see Section II-E). This reduction in the drop velocity variation DoD frequency. The actuation pulse used in the experiment
can reduce the drop placement error by 40% compared to [13]. is the standard actuation pulse. In order to extract sufficient
information from the experimental data, the input signal should
VI. C ONCLUSION be persistently exciting [19]. Therefore, to improve the order
In this paper we have proposed a robust feedforward control of excitation of the input signal, a white noise signal with a
by extending the optimization-based technique in [13]. The small variance is added in between two consecutive actuation
actuation pulse design in [13] can provide best results if pulses. Fig.9 shows the piezo input and the piezo sensor output
the actual system dynamics do not deviate from the model measured during an experiment when ink droplets are jetted
used in the design. We have observed that the inkjet system at a fixed DoD frequency of 20 kHz.
dynamics at different DoD frequencies will not be the same.
Therefore, we have proposed to represent this set of dynamical 150
Piezo sensor output (mV)

models by a compact parametric uncertainty ∆ ∈ ∆ on the 100

50
nominal model of the inkjet system, i.e. Hlow (q, ∆). System 0
identification results are presented to support the choice of the −50

parametric uncertainty ∆ ∈ ∆. In order to damp the residual −100

oscillations in the presence of this parametric uncertainty, the −150


0 10 20 30 40 50 60 70 80 90 100

optimization-based method is extended to design the robust ac-


25
tuation pulse which minimizes the worst-case squared tracking
20
error for Hlow (q, ∆) with ∆ ∈ ∆. Experimental results have
Piezo Input (V)

15
demonstrated that a considerable improvement in the ink drop 10

consistency can be achieved with the proposed robust pulse. 5

In this paper we have shown how control techniques can im- 0

−5
prove the performance of an inkjet printhead. Note that control 0 10 20 30 40 50
Time (µs)
60 70 80 90 100

is here applied to a printhead which is already designed. In our


3 In order to capture the flight profile of jetted drops, images of the jetted
Fig. 9. Experimental data obtained while jetting at a fixed DoD frequency
drops are captured with a CCD camera at an interval of 10 µs. All the images of 20kHz.
are placed adjacent to each other in the order of the time instant when they
are taken. The composite image constructed in this manner is shown in Fig. 8.
The details about the experimental setup, such as the camera, the microscopic We have used the Identification toolbox of Matlab, in
lens, etc. can be found in [18]. particular we have used Prediction-Error (PE) identification to
DRAFT SUBMITTED TO: IEEE JOURNAL OF MICROELECTROMECHANICAL SYSTEMS 8

p1
30
piezo
20
input u output p
10

p2
ink
Magnitude (dB)

−10

Fig. 11. The basic principle to measure the acoustic sensor signal
−20

−30
400

−40 1 2
10 10
Freq. (kHz) 200

0
Fig. 10. Magnitude plot of the transfer function from the piezo input to the

Piezo output (mV)


piezo sensor signal.
−200

obtain a discrete-time transfer function for the experimental −400

data. The transfer function is identified in the Box-Jenkins


model structure. The order of the plant is chosen to be 6 and −600 pfill

the order of the noise model is chosen to be 2. The order of pempty


sensor signal: pfill − pempty
the transfer function is chosen to ensure that the true inkjet −800
0 10 20 30 40 50 60 70 80 90 100
system dynamics are contained in the model structure. This Time (µs)

can be validated by analyzing the statistical properties of the


residual signal and ensuring that this residual signal is not
correlated with the input signal. Fig. 12. Experimental data obtained while jetting at a DoD frequency of
We have done several such experiments at various fixed 20kHz.
DoD frequencies in the operating range of the inkjet printhead,
i.e., from 20 kHz to 70 kHz.
has to be extracted from the measured signal p. Typically,
Fig. 10 shows the magnitude plot for the transfer function
the contribution of the indirect-path (i.e. the sensor signal
from the piezo input voltage (V) to the piezo sensor signal
p2 ) is very small compared to the direct-path. Consequently,
(mV) obtained at those DoD frequencies. These results confirm
it is difficult to measure the sensor signal (indirect-path)
that the main variation of the dynamics of the system (made
simultaneously while using the piezo as an actuator.
up of different resonance peaks) is a variation of the first
In [5], a hardware compensation method is proposed to
resonance peak. The observed variation in the second peak
extract the sensor signal p2 . In this method, two ink channels
is neglected since it has a smaller amplitude and since higher
with a similar piezo capacitance are connected in the bridge
order peaks are known to be of less influence for the ink drop
circuit. Out of these two ink channel one ink channel is
properties [14].
filled with the ink and the other ink channel is empty. Thus,
the sensor signal p2 is obtained by subtracting the measured
A PPENDIX B signals of the two ink channels. However, there are several
M EASUREMENT OF THE PIEZO SENSOR SIGNAL drawbacks of this method. A major difficulty is to make the
Self-sensing of a piezo unit has been employed in various hardware changes in the ink printhead to create an empty
applications for simultaneous actuation and sensing. Generally, ink channel. Another drawback is that even though the ink
the self-sensing unit uses a bridge circuit with a piezo actuator channel is empty, a small contribution due to the deformation
in one arm and a passive capacitor of equivalent capacitance in of the structure is present in the indirect-path. Also, there
the other arm. This ensures that one can measure the dynamic will be always a small difference in the piezo capacitance.
behavior of the piezo structure. However, in case of an inkjet In order to overcome these drawbacks, we use the offline
channel, the measured signal p from this bridge circuit consists approach proposed in [20] to reconstruct the sensor signal
of two contributions (see Fig. 11). The first contribution p1 is p2 . We have carried out two experiments and measure the
the charge generated proportional to the piezo deformation piezo output signal p from the ink channel by applying the
due to the applied actuation voltage and is referred to as the same actuation pulse at a fixed DoD frequency. The first
direct-path. The second contribution p2 originates from the experiment is carried out with the ink inside the channel and
force exerted by the ink in the channel and is referred to as the measured piezo output signal pfill = p1 + p2 is stored.
the indirect-path. As the second contribution has information During the second experiment the channel is kept empty
about the jetting process, it is the required sensor signal and and the measured piezo output signal pempty = p1 is stored.
DRAFT SUBMITTED TO: IEEE JOURNAL OF MICROELECTROMECHANICAL SYSTEMS 9

Thus, the sensor signal p2 is obtained by offline subtraction


of the signal pempty from the signal pfill . Fig. 12 shows the
signals pfill , pempty and the reconstructed piezo sensor signal
p2 = pfill − pempty at a DoD frequency of 20kHz.

R EFERENCES
[1] C. Williams, “Ink-jet printers go beyond paper,” Physics World, vol. 19,
pp. 24–29, 2006.
[2] T. Kawase, T. Shimoda, C. Newsome, H. Sirringhaus, and R. H. Friend,
“Inkjet printing of polymer thin film transistors,” Thin Solid Films, vol.
438439, no. 0, pp. 279 – 287, 2003.
[3] V. Pekkanen, M. Mntysalo, K. Kaija, P. Mansikkamki, E. Kunnari,
K. Laine, J. Niittynen, S. Koskinen, E. Halonen, and U. Caglar,
“Utilizing inkjet printing to fabricate electrical interconnections in a
system-in-package,” Microelectronic Engineering, vol. 87, no. 11, pp.
2382 – 2390, 2010.
[4] A. van der Bos, T. Segers, R. Jeurissen, M. van den Berg, H. Reinten,
H. Wijshoff, M. Versluis, and D. Lohse, “Infrared imaging and acoustic
sizing of a bubble inside a micro-electro-mechanical system piezo ink
channel,” Journal of Applied Physics, vol. 110, no. 3, p. 034503, 2011.
[5] M. Wassink, “Inkjet printhead performance enhancement by feedforward
input design based on two-port modeling,” Ph.D. dissertation, Delft
University of Technology, 2007.
[6] D. Bogy and F. Talke, “Experimental and theoretical study of wave
propagation phenomena in drop-on-demand ink jet devices,” IBM J. Res.
Dev., vol. 28, no. 3, pp. 314–321, 1984.
[7] J. Chung, S. Ko, C. Grigoropoulos, N. Bieri, C. Dockendorf, and
D. Poulikakos, “Damage-Free low temperature pulsed laser printing of
gold nanoinks on polymers,” Journal of Heat Transfer, vol. 127, no. 7,
pp. 724–732, Jul. 2005.
[8] MicroFab Technologies Inc., “Drive waveform effects on ink-jet device
performance,” 99-03, Tech. Rep., 1999.
[9] H. Gan, X. Shan, T. Eriksson, B. Lok, and Y. Lam, “Reduction of droplet
volume by controlling actuating waveforms in inkjet printing for micro-
pattern formation,” Journal of Micromechanics and Microengineering,
vol. 19, no. 5, p. 055010, 2009.
[10] H. Dong, W. Carr, and J. Morris, “An experimental study of drop-on-
demand drop formation,” Physics of Fluids, vol. 18, no. 7, p. 072102,
Jul. 2006.
[11] K. Kwon and W. Kim, “A waveform design method for high-speed inkjet
printing based on self-sensing measurement,” Sensors and Actuators A:
Physical, vol. 140, no. 1, pp. 75–83, 2007.
[12] K. Kwon, “Waveform design methods for piezo inkjet dispensers based
on measured meniscus motion,” Journal of Microelectromechanical
Systems, vol. 18 (5), pp. 1118–1125, 2009.
[13] A. Khalate, X. Bombois, R. Babuška, H. Wijshoff, and R. Waarsing,
“Performance improvement of a drop-on-demand inkjet printhead using
an optimization-based feedforward control method,” Control Engineer-
ing Practice, vol. 19, pp. 771–781, 2011.
[14] J. Dijksman, “Hydrodynamics of small tubular pumps,” Journal of Fluid
Mechanics, vol. 139, pp. 173–191, 1984.
[15] H. Wijshoff, “The dynamics of the piezo inkjet printhead operation,”
Physics Reports, vol. 491, pp. 77–177, 2010.
[16] C. Rabbath, “Sensitivity of the discrete- to continuous-time pole trans-
formation at fast sampling rates,” Master’s thesis, McGill University,
Canada, 1995.
[17] A. Khalate, B. Bayon, X. Bombois, G. Scorletti, and R. Babuška, “Drop-
on-demand inkjet printhead performance improvement using robust
feedforward control,” in IEEE Conference on Decision and Control and
European Control Conference, Orlando, 2011.
[18] H. Wijshoff, “Structure and fluid-dynamics in piezo inkjet printheads,”
Ph.D. dissertation, University of Twente, 2008.
[19] L. Ljung, System Identification - Theory For the User. Prentice Hall,
Upper Saddle River, 1999.
[20] K. S. Kwon, “Methods for detecting air bubble in piezo inkjet dis-
pensers,” Sensors and Actuators A: Physical, vol. 153, no. 1, pp. 50 –
56, 2009.

View publication stats

You might also like