Vous êtes sur la page 1sur 10

http://www.paper.edu.

cn

406

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 15, NO. 3, MAY 2007

Gain Scheduled
Control for Air Path Systems of
Diesel Engines Using LPV Techniques
Xiukun Wei, Associate Member, IEEE, and Luigi del Re, Member, IEEE

AbstractThis paper addresses the modeling and control of the


air path system of diesel engines. The underlying issues are critical
for the control of the transient exhaust gas fraction pumped into
the cylinders, which is known to be a dominant factor to reduce
the nitrogen oxides (NOx ) emissions. In this paper, we propose a
new approach, based on a data-based grey-box linear parameter
techvarying (LPV) model as well as on the gain scheduled
nique for the controller design. The modeling step is shown to lead
naturally to a so-called quasi-LPV structure, which also delivers
the scheduling variables to be accounted for. Using this informatechniques allow to design a controller
tion, gain scheduled
which enforces a much better tracking performance than the standard production electronic control unit, while not requiring any
calibration work. The performance of the proposed approach is
demonstrated by experimental results.
Index TermsDiesel engines, engine control, gain scheduling,
linear parameter varying, robust control, system identification.

I. INTRODUCTION
HE popularity of the light-duty diesel vehicles in the European market has grown significantly over the past decade,
with no sign of slowing down. European auto manufacturers expect that overall European diesel penetration will soon climb
to more than 40%. The reason why diesel engines have won
such approval in Europe is that diesels offer several significant
performance advantages over gasoline engines in terms of fuel
efficiency, power, durability, and certain emissions. Light duty
diesels use substantially less fuel, produce more torque at lower
engine speeds, require less maintenance, and have longer recommended service intervals than gasoline engines of similar
power. Because diesels burn less fuel than gasoline vehicles,
they also produce significantly lower emissions of greenhouse
gases such as carbon dioxide. However, diesel engines produce a
substantial amount of highly toxic nitrogen oxides (NO ) which
should be reduced as much as possible in the very near future
to meet the more and more stringent emission legislations from
the European Union and the United States.
One of the most effective means to attenuate raw NO emissions consists in using advanced control strategies to regulate

Manuscript received May 26, 2006; revised October 29, 2006. Manuscript
received in final form January 12, 2007. Recommended by Associate Editor
K. Butts. This work was supported by the Linz Center of Competence in Mechatronics (LCM), Linz, Austria.
X. Wei was with the Institute of Design and Control of Mechatronical Systems, Johannes Kepler University, Linz A-4040, Austria. He is now with the
Delft Center for Systems and Control, Delft University of Technology, 2628
CD Delft, The Netherlands (e-mail: xiukun.wei@tudelft.nl).
L. del Re is with the Institute of Design and Control of Mechatronical Systems, Johannes Kepler University, Linz A-4040 Austria (e-mail: luigi.delre@
jku.at).
Digital Object Identifier 10.1109/TCST.2007.894633

the transient exhaust gas recirculation (EGR) rate. The principles relating EGR to the NO production in the combustion
chamber have been well-studied (see, e.g., [1][4]), which
show, among other, that the transient exhaust gas fraction in the
mixed gas (the mixture of the fresh air from the compressor and
exhaust gas from the exhaust manifold) injected into the cylinders has a dominant impact on the NO generation. However,
up to now, the regulation techniques for EGR are based on gain
scheduled proportional-integral differential (PID) controllers
which depend on some heuristically optimized maps. The calibration of the controller parameters and the maps is annoying
and time consuming work, and the control performance of
the entire system is suboptimal. This motivated some research
efforts in the recent years.
In our former research, a robust output control approach
based on the exhaust gas oxygen (EGO) sensor was presented
in [5]. An early mean value model can be found in [6] where
the diesel engine modeling procedure based on first principles
is discussed as well. The modeling and control issues in turbo
charged diesel models equipped with variable geometry turbine
and exhaust gas recirculation system are reported in [7][9], in
which also plant nonlinearities, complex multiple-input multiple-output (MIMO) interactions, actuators saturation, system
bandwidth limitations, and nonminimum phase characteristics
are discussed. In [10][20], the multivariable control of air path
systems is considered by different control approaches.
Also, the work in this paper considers the transient EGR regulation of the air path of diesel engines. The general aim is also
common to other works, to enforce a precise tracking of references in terms of boost pressure and air fuel ratio in the intake manifold. Our main interest, however, lies in deriving control laws which offer a good tracking performance but can also
be automatically tuned, possibly in an adaptive fashion. To this
end, we need identification techniques instead of first principles modeling and/or manual calibration. Identification, however, does not exclude physical understanding of the plant and,
in particular, turns out to provide a precious guide for the choice
of the model structure.
Engines are known to be nonlinear systems, but a generic
nonlinear model can hardly be used for control and so must be
approximated either by local linear models or by some nonlinear
class able to capture its essential elements but simple enough
to be suitable both for parameter estimation and for control design. This paper takes this second way and uses linear parameter
varying (LPV) models, for which identification techniques have
been developed (see [21][28]), but for which also suitable control design approaches exist (as, e.g., in [29][32]). LPV models
seem a good choice, because they can be treated as to a large

1063-6536/$25.00 2007 IEEE


WEI AND DEL RE: GAIN SCHEDULED

http://www.paper.edu.cn
CONTROL FOR AIR PATH SYSTEMS OF DIESEL ENGINES USING LPV TECHNIQUES

407

path dynamics, are in a significantly different time scale and,


thus, can be treated as measured disturbances.
Notation:
intake manifold pressure;
exhaust manifold pressure;
engine speed;
fuel mass flow;
variable geometry vane position;
exhaust gas recirculation valve position;
intake manifold temperature;
exhaust manifold temperature;
compressor air mass flow.

Fig. 1. Air path system of turbocharged diesel engines.

extent as linear ones, while offering the capability to account


for significant nonlinear effects. LPV systems are also known
to approximate well-known first principle models of the air path
system [20]. Still, the obtained LPV model is only the approximation of the real plant, hence, the control system design must
as
offer some robustness, the main reason for the use of
control design method. In addition, under the
framework,
the disturbances such as the fuel mass injection can be incorporated into the controller synthesis explicitly.
The remainder of this paper is organized as follows. In the
Section II the modeling procedure of the air path system of
turbocharged diesel engines by LPV identification techniques
is presented and a quasi-LPV model is derived. In the Section
III, the gain scheduling techniques for LPV systems are briefly
controller synthesis based
reviewed. The gain scheduled
on a nominal parameter set and the closed-loop system performance demonstrated by experimental results are presented in
Section IV. Some conclusions are presented at the end of this
paper.

As we intend to use the LPV framework for modeling, we


also need to recall some essential definitions and properties.
In the state space form, an LPV system can be described in
the form

where
is an external quantity, the scheduling variable,
which is assumed to be measurable in real time. Clearly, an LPV
system reduces to a linear time varying (LTV) system for a speand to an LTI system on constant tracific trajectory
jectories
,
. It follows that for frozen values of
the scheduling variable, LPV systems can be treated as LTI systems.
If the scheduling variable is or depends on system state variables, the LPV becomes a quasi-LPV. In this generic formulation, a quasi-LPV system is simply a generic input affine nonlinear system. Here, we use the name quasi-LPV in its original
meaning (quasi means almost), i.e., to designate systems
which are not really LPV but can be treated as such because the
time scale of the state-dependent scheduling variables is significantly longer than the one of the process to be controlled.
A generic quasi-LPV system model has the form

II. MODELING THE PLANT


The schematic diagram of the BMW M47D turbocharged
diesel engine is shown in Fig. 1. The compressor pumps the
fresh air into the intake manifold to boost the pressure. The
fuel is sprayed directly into the combustion chamber where it is
burnt. Part of the exhaust gas is recirculated into the intake manifold to reduce the NO emission by lowering the oxygen content of the intake gases and the combustion peak temperatures.
Variable geometry turbine (VGT) utilizes the heat energy from
the exhaust gas and propels the compressor. An inter cooler is
used to lower the fresh air temperature and an EGR cooler to
lower the temperature of the recirculated gas, respectively.
The manipulated quantities are the VGT vane position and
EGR valve opening and the outputs are the boost pressure and
compressor air mass flow. Other quantities, like the engine
speed or the temperatures, are indeed also system outputs, but,
for the analysis of fast phenomena like torque production or air

where the scheduling parameter is a vector


,
and
are exogenous scheduling variables.
is part
.
of the system state
It is also possible to define the LPV system in the input-output
framework

where has the same meaning as in the state space case.


LPV has arisen from gain scheduling and offers, with respect
to the use of switched or combined linear models, a continuous
transition and typically a better approximation quality if there
really is one or a few variables which explain the changes in the
dynamics of the system. In the case of engines, indeed, there
are a few basic variables which are known to explain at least
a large part of the changes in the dynamics, essentially speed


408

http://www.paper.edu.cn
IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 15, NO. 3, MAY 2007

(which affects the time available for gas exchange and combustion) and load (in some form, which fixes the average temperatures), so LPV is a natural candidate. Parameters of LPV models
can be estimated from measured data by linear algorithms extended from the classical identification algorithms, both in their
state space form [22] and in their input-output form [26][28].
For the identification of the air path, the approach described in
[26] has been used in this work, which is based on a polynomial
representation

Fig. 2. Inputs and outputs of the air path system.

(1)
is the backward shift operator,
are the
where
denotes the modeling error consisting
scheduling variables,
and unmodeled dynamics
and
of bounded disturbance
,
is the system output at time
, and
is the system input. The polyand
are defined by
nomials

Fig. 3. Intake manifold pressure model structure.

(2)
, and
. The coefficients of the
where ,
polynomials and are generic functions of , which can be
expressed as the weighted sum of some special functions as
in the following structure:
(3)
or
, and
where
are constant values,
are functions of the online measurable variables . The values of these
functions can be calculated directly from , e.g.,
or
, but could also be different powers of .
Define
Fig. 4. Validation result of the boost pressure model.

where

,
,
. Thus, system (1) can be rewritten as

, and

1) intake manifold pressure dynamics;


2) exhaust manifold pressure dynamics;
3) air mass flow dynamics.
To evaluate the model quality, the variance-accounted-for (VAF)
value will be used

(4)

(5)

with
,
, and denotes the Kronecker product. Please refer to [26] for the detailed identification algorithms.
After having presented the tools, the next step in deriving the
model consists in fixing inputs and outputs as shown in Fig. 2:
, the VGT vane position
, the engine speed
the EGR rate
, and the fuel mass flow
are the system inputs. The system
outputs are the intake manifold pressure and the air mass flow
. In this work, it has proven sensible to divide the modeling
of the air path system into the following three subsystems:

1) Intake Manifold Pressure Dynamics: The best model


structure for this subsystem (see Fig. 3) was found to be a firstorder quasi-LPV system with the exhaust manifold pressure as
the input and the intake manifold pressure as the scheduling
variable. The dynamics is slightly dependent on the intake manifold pressure. The model has been validated with EGR operation
(see Fig. 4). The VAF values are over 85% for typical measured
data sets.
2) Exhaust Manifold Pressure Dynamics: The model structure used for the parameter estimation is shown in Fig. 5. A
static nonlinearity and a small time delay from the EGR rate
to the exhaust manifold pressure path were found experimentally. The dynamics from the VGT vane position to the exhaust
manifold pressure include a static nonlinearity which depends

where


WEI AND DEL RE: GAIN SCHEDULED

http://www.paper.edu.cn
CONTROL FOR AIR PATH SYSTEMS OF DIESEL ENGINES USING LPV TECHNIQUES

409

Fig. 5. Exhaust manifold pressure structure with EGR.

Fig. 8. Validation result of the air mass flow model.

Fig. 6. Validation result of the exhaust manifold pressure model when EGR is
active.
Fig. 9. Quasi-LPV model structure for the air path system of diesel engines.

Fig. 7. Air mass flow model structure.

on the boost pressure. The product of the boost pressure and the
engine speed is a special input of the system suggested from the
mean value model experience. A validation result of this model
is shown in Fig. 6, where the VAF value is 82%.
3) Air Mass Flow Dynamics: The model structure used for
the air mass flow dynamical model is shown in Fig. 7. A validation result is shown in Fig. 8, where the VAF value is 90%.
Combining the submodels yields the entire air path system
model shown in Fig. 9. It can be seen that the system dynamics

slightly depends also on the boost pressure, which, therefore,


should be treated as a scheduling variable. As this effect is not
dominant, and we want to use the same variable as output to be
tracked, we neglect this dependency. The engine speed can be
set into the numerator of the corresponding transfer function as
. The impact of the VGT vane position to the
air mass flow dynamics can be handled the same way. In such a
case, the final model is a Hammerstein quasi-LPV model where
the engine speed and the VGT vane position are the scheduling
variables. The entire model of the air path system is validated by
the input steps. The result is shown in Fig. 10. This final model
with its numerical values (for the specific engine used) is shown
in (6) and (7)

(6)
(7)


410

http://www.paper.edu.cn
IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 15, NO. 3, MAY 2007

The scheduling variables are the normalized engine speed and


at engine speed
VGT vane position. We define
2500 r/min and
at VGT position
. The
considered engine speed range is between 1000 and 4000 r/min.
The VGT vane range is from 0% to 100%. In such a case, the
two scheduling variables vary in the following ranges:

The frequency response from the inputs to the outputs are


shown in Fig. 11 for different engine speeds and Fig. 12 for
different VGT vane positions.
III. GAIN SCHEDULED

CONTROL FOR LPV SYSTEMS

An instinctive choice for an LPV system could consist in


looking for a control law which ensures the internal stability and
a guaranteed performance bound for the closed-loop system in
spite of the scheduling parameter varying, which somehow includes a compensation of the scheduling variable. This could be
done, for instance, by taking an LPV form for the control law
as well, i.e., of the form
(8)
(9)
where
is the state of the controller. This essentially means
,
to move the design effort in finding a scheduling law for
,
, and
such that the composition of the quasi-LPV
system and this controller is optimal in some sense. In a
framework, this can be obtained by extending the model description by a performance variable and a disturbance input
, obtaining
(10)
(11)
(12)
where
, and, usually,
, where
is a
polytope in
. Assuming the system matrices to be affine in
means that they belong to the polytope defined by

Fig. 10. Validation results of the air path system by the input steps.

(13)
where

where

denote the values of the matrices


at the vertices of the parameter polydenotes the convex hull of a finite number of

tope .
matrices.
,
,
, and
Assume that the system matrices
are parameter independent. In addition, the disturbance
does not affect the performance output and there is no feed


WEI AND DEL RE: GAIN SCHEDULED

http://www.paper.edu.cn
CONTROL FOR AIR PATH SYSTEMS OF DIESEL ENGINES USING LPV TECHNIQUES

411

is any solution to the convex decompowhere


.
are the corners of the
sition problem
parameter vector in its box of . In other words, the controller
are obtained by
state space matrices at the operating point
convex interpolation of the LTI vertex controllers

Fig. 11. Frequency response from the inputs to the outputs for different speeds.
The upper left subplot is for  to W ; upper right is for  to p ; bottom
to W ; bottom right is for 
to p .
left is for 

For the detailed procedure, please refer to [29].


As pointed out in [31], such approaches are potentially conservative because they admit arbitrary rates of variation in the
scheduling variables. More dramatically, it has been shown that
some systems would not be even quadratically stabilizable. A
significant improvement of such techniques can be obtained by
exploiting the concept of parameter-dependent Lyapunov functions which allow the incorporation of knowledge on the rate
of variation in the analysis or synthesis technique, and, therefore, lead to less conservative controllers. To cope with this, in
[30], an advanced gain scheduling technique has been proposed
which takes in account the variation rate of the scheduling variable restating the control as
(15)
(16)
which is shown to ensure internal stability and a guaranteed
gain bound from the disturbance signal to the performance
signal , that is
(17)

Fig. 12. Frequency response from the inputs to the outputs for different values
of  . The upper left subplot is for  to W ; upper right is for  to p ;
to W ; bottom right is for 
to p .
bottom left is for 

through from the input to the measured output, that is,


and
, respectively. It was found (see [29]) under
these assumptions that the values of the matrices of the LPV
controller can be given by
(14)

for all admissible parameter trajectories


.
Sufficient solvability conditions for this problem can be derived using a suitable extension of the bounded real lemma [31]
and by confining the search of (Lyapunov) variables to some finite-dimensional subspace of functions of . For details see [30].
From the practical point of view, it is important to remark
that the solution of the problem includes the inverse of matrix
obtained through an (online) factorization, which might be very
ill-conditioned, and that in principle, an infinite number of LMIs
should be solved, as the solution depends on the scheduling variables. In practice, a sequential gridding approach can be used,
but it still computationally remains very intensive and numerically unsafe.
In the following, we shall denote the controller synthesized
from the first approach as basic LPV controller and the one
based on the second approach as advanced LPV controller,
respectively.
IV. CONTROLLER SYNTHESIS AND EXPERIMENTAL RESULTS
LPV
The closed-loop performance of gain scheduled
control can be specified by the choice of the weighting functions. The closed-loop system with weighting functions is
shown in Fig. 13.

http://www.paper.edu.cn

412

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 15, NO. 3, MAY 2007

Fig. 13. Interconnection structure for the closed-loop system with weighting functions.

Fig. 14. Closed-loop frequency response from reference inputs to outputs. The
upper left subplot is for air mass reference
to
; upper right is for
to ; bottom left is for intake manifold pressure reference
to
; bottom
right is for
to .

Fig. 15. Air mass flow step response at engine speed

Ne = 1700 r/min.

Fig. 16. Air mass flow step response at engine speed

Ne = 2000 r/min.

The weighting function set used for the controller synthesis


is

Using the first approach (with a linear combination of the


vertex controllers) synthesized with the LMI control toolbox
[33], leads to a controller with eight states with a polytopic description. The frequency response of the closed-loop system of
one vertex is shown in Fig. 14. It can be seen that the closed-loop
system bandwidth of the nominal system is increased. At low
frequency, the system can work like two independent singleinput single-output (SISO) systems due to the decoupling effects on the closed system that can be seen from the frequency
response plots.
Testing this basic LPV controller in the test cell using the
dSPACE environment by C-coded S-functions yields the results
shown in the next figures. In the boost pressure and air mass flow

subplots of all these figures, the solid line and dashed-dotted


line are utilized to express the reference signal and the measured
output response, respectively.
In Figs. 15 and 16, the step responses of the air mass flow
are shown at three different operation points with engine speed
1700 and 2000 r/min, respectively. In most of the time,
the rise time is less than 1.5 s. The overshoot is less than 10%.


WEI AND DEL RE: GAIN SCHEDULED

http://www.paper.edu.cn
CONTROL FOR AIR PATH SYSTEMS OF DIESEL ENGINES USING LPV TECHNIQUES

Fig. 17. Intake manifold pressure step response at engine speed


1700 r/min.

Fig. 18. Intake manifold pressure step response at engine speed


2000 r/min.

Ne =

Ne =

The settling time is also less than 2 s, while the fluctuations of


the boost pressure are very small and decrease very fast. The
boost pressure step response at the same engine speed points
are shown in Figs. 17 and 18, respectively. The overshoot of the
step responses are all less than 15%. The rise time is less than
1.5 s. The settling time is also less than 2.5 s. At all these operation points, the expected fast transient tracking is achieved. At
the same time, the fluctuations of the air mass flow are regulated
to set points fast enough, even if the magnitude of the fluctuation
has some difference due to the system nonlinearity and uncertainties. The controller has also been tested with the set points
directly from the maps in the ECU. The result corresponding to
a sinusoidal variation of the fuel mass flow is shown in Fig. 19.
It can be seen that an acceptable tracking performance for the
air mass flow is achieved. The boost pressure can not track the
reference signal since the set point is unfeasible.
It may be interesting to compare the performance with the
advanced controller. Again, the LMI toolbox can be used,
the different LPV controllers can be computed while using

413

Fig. 19. Fuel sinusoid disturbance response at engine speed

Fig. 20. Air mass flow step response at engine speed


advanced controller.

Ne = 2000 r/min.

Ne = 2500 r/min of the

different combination of the Lyapunov variables and [30].


are gridded with ten points
The scheduling variables and
for each variable. In this case, around 1200 LMIs should be
solved. For each case, at least half an hour is needed to obtain
the solutions with Pentum-4 CPU 3G with 1-GB memory.
The advanced LPV controller was also tested in the test cell.
Some results are shown in the plots from Figs. 2022. Most
of the time, the controller can reach a little better performance
than the basic one. The rise time is smaller than the basic LPV
controller. The control performance achieved from the two approaches are shown in Table I, from which it can be concluded
(as also from the measurements) that the closed-loop performance is improved by the advanced LPV synthesis approach,
however, the performance difference is not very large. A detailed comparison shows that the advanced LPV controller can
achieve less overshoot and faster regulation of the intake manifold pressure, however, there is a slightly larger fluctuation of
the boost pressure when the air mass flow tracks its reference
signal.

http://www.paper.edu.cn

414

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 15, NO. 3, MAY 2007

Fig. 21. Intake manifold pressure step response at engine speed


r/min of the advanced controller.

Ne = 1800

models and obtaining these models from first principles can be


extremely time consuming and still imprecise. Furthermore,
generic nonlinear models are usually not suitable for control
design.
Against this background, this paper has proposed a new approach based on LPV identification and control design techniques which uses physical background knowledge to define the
model structure, but recurs to numerical techniques to reduce the
tuning work (which, at the end, consists solely in defining the
weighting functions for performance and robustness).
The modeling procedure consists in choosing physically motivated subsystems and determining simple and sensible models
for each of them, then using LPV identification techniques to
estimate their parameters. A final quasi-LPV model is the combination of the three sub-models, and has been used for the control synthesis. To this end, robust techniques, two approaches
, have been used and tested in real time
of gain scheduled
on a production engine, yielding a very good tracking behavior.
It seems, therefore, that the use of LPV modeling and control
techniques for engine control is a promising approach also in
the industrial context.
The limited difference in performance between the basic control approach as from (14) and the advanced version proposed
by [30] can be interpreted as that the model information on the
variation rates of the scheduling variables, engine speed, and the
VGT position, is not really significant.
A next step could, therefore, consist in trying to optimize
the structure of the LPV models using statistical or nonlinear
parametrically varying [27] approaches to determine the optimal
choice of the scheduling variables.
ACKNOWLEDGMENT
The authors would like to thank Prof. L. Glielmo, who attracted their attention to the use of LPV models a few years ago.
They would also like to thank the three anonymous reviewers for
their comments and suggestions.
REFERENCES

Fig. 22. Speed disturbance response of the advanced controller.

TABLE I
CONTROL PERFORMANCE OF THE TWO APPROACHES

V. CONCLUSION
There exists a wide consensus that the actual industrial
standard for engine control, essentially a huge collection of
very simple, mostly heuristically tuned control loops, has very
many drawbacks, both in terms of performance and of calibration workload. On the other side, model-based control needs

[1] J. B. Heywood, Internal Combustion Engine Fundamentals. New


York: McGraw-Hill, 1988.
[2] N. Ladommatos, S. M. Abedlhalim, H. Zhao, and Z. Hu, The dilution,
chemical, and thermal effects of exhaust gas recirculation on diesel engine emissionsPart1: Effects of carbon dioxide, SAE Int., Warrendale, PA, Tech. Rep. 961167, 1996.
[3] , The dilution, chemical, and thermal effects of exhaust gas recirculation on diesel engine emissionsPart2: Effects of water vapour,
SAE Int., Warrendale, PA, Tech. Rep. 971658, 1997.
[4] , The dilution, chemical, and thermal effects of exhaust gas
recirculation on diesel engine emissionsPart3: Effects of carbon
dioxide and water vapour, SAE Int., Warrendale, PA, Tech. Rep.
971659, 1997.
[5] A. Amstutz and L. del Re, EGO sensor based robust output control of
EGR in diesel engines, IEEE Trans. Contr. Syst. Technol., vol. 3, no.
1, pp. 3948, Mar. 1995.
[6] M. Kao and J. J. Moskwa, Turbocharged diesel engine modeling for
nonlinear engine control and state estimation, ASME J. Dyn. Syst.
Meas., Control, vol. 117, pp. 2030, Mar. 1995.
[7] M. Jung, R. G. Ford, K. Glover, N. Collings, U. Christen, and
M. J. Watts, Parameterization and transient validation of a variable geometry turbocharger for mean-value modeling at low and
medium speed-load points, SAE Int., Warrendale, PA, Tech. Rep.
2002-01-2729, 2002.


WEI AND DEL RE: GAIN SCHEDULED

http://www.paper.edu.cn
CONTROL FOR AIR PATH SYSTEMS OF DIESEL ENGINES USING LPV TECHNIQUES

[8] I. V. Kolmanovsky, P. E. Moraal, M. J. van Nieuwstadt, and A. Stefanopoulou, Issues in modeling and control of intake flow in variable
geometry turbocharged engines, in Proc. 18th IFIP Conf. Syst. Model.
Opt., 1997, pp. 436445.
[9] X. Wei and L. del Re, Modeling and Control of the boost pressure
for a diesel engine based on LPV techniques, in Proc. Amer. Control
Conf., 2006, pp. 18921897.
[10] M. Larsen and P. V. Kokotovic, Passivation design for a turbocharged
diesel engine model, in Proc. IEEE Conf. Dec. Control, 1998, pp.
15351540.
[11] M. van Nieustandt, P. E. Moraal, I. Kolmanovsky, A. G. Stefanopoulou,
P. Wood, and M. Criddle, Mean-value modelling and robust control of
the air path of a turbocharged diesel engine, in Proc. IFAC Workshop
Adv. Autom. Control, 1998, pp. 15351540.
[12] A. G. Stefanopoulou, I. V. Kolmanovsky, and J. S. Freudenberg, Control of variable geometry turbocharged diesel engines for reduced emissions, in Proc. Amer. Control Conf., 1998, pp. 191196.
[13] M. van Nieuwstadt, I. V. Kolmanovsky, P. E. Moraal, and A. G. Stepfanopoulou, Experimental comparison of EGR-VGT control schemes
for high speed diesel engine, IEEE Control Syst. Mag., vol. 20, no. 3,
pp. 6379, Mar. 2000.
[14] A. G. Stefanopoulou, I. V. Kolmanovsky, and J. S. Freudenberg, Control of variable geometry turbocharged diesel engines for reduced emissions, IEEE Trans. Control Syst. Technol., vol. 8, no. 4, pp. 733745,
Jul. 2000.
[15] M. Larsen, M. Jankovic, and P. V. Kokotovic, Coordinated passivation designs, Automatica, vol. 39, pp. 335441, 2003.
[16] M. Jankovic, M. Jankovic, and I. V. Kolmanovsky, Robust nonlinear
controller for turbocharged diesel engine, in Proc. Amer. Control
Conf., 1998, pp. 13891394.
[17] V. I. Utkin, H. C. Chang, I. V. Kolmanovsky, and J. A. Cook, Sliding
mode control for variable geometry turbocharged diesel engines, in
Proc. Amer. Control Conf., 2000, pp. 584588.
[18] M. Jankovic, M. Jankovic, and I. V. Kolmanovsky, Constructive Lyapunov control design for turbocharged diesel engines, IEEE Trans.
Control Syst. Technol., vol. 8, no. 2, pp. 288299, Mar. 2000.
[19] M. Jung and K. Glover, Control-oriented linear parameter varying
modelling of a turbocharged diesel engine, in Proc. IEEE Conf. Control Appl., 2003, pp. 155160.
[20] M. Jung, Mean-value modeling and robust control of the airpath of
a turbocharged diesel engine, Ph.D. thesis, Dept. Eng., Univ. Cambridge, Cambridge, U.K., 2003.
[21] V. Verdult, Nonlinear system identification: State-space approach,
Ph.D. thesis, Faculty Appl. Phys., Univ. Twente, Enschede, The
Netherlands, 2001.
[22] V. Verdult and M. Verhaegen, Subspace identification of multivariable linear parameter-varying systems, Automatica, vol. 38, no. 5, pp.
805814, May 2002.
[23] , Kernel methods for subspace identification of multivariable
LPV and bilinear systems, Automatica, vol. 41, no. 9, pp. 15571565,
Sep. 2005.
[24] V. Verdult, M. Lovera, and M. Verhaegen, Identification of linear parameter-varying state space models with application to helicopter rotor
dynamics, Int. J. Control, vol. 77, no. 13, pp. 11491159, 2004.
[25] X. Wei, Advanced LPV techniques for diesel engines, Ph.D. dissertation, Dept. Des. Control Mechatron. Syst., Johannes Kepler Univ. Linz,
Linz, Austria, 2006.

415

[26] X. Wei and L. del Re, On persistent excitation for parameter estimation of quasi-LPV Systems and its application in modeling of diesel
engine torque, in Proc. 14th IFAC Symp. Syst. Id., 2006, pp. 517522.
[27] F. Previdi and M. Lovera, Identification of nonlinear parametrically
varying models using separable least squares, Int. J. Control, vol. 77,
no. 16, pp. 13821392, 2004.
[28] B. Bamieh and L. Giarre, Identification of linear parameter varying
models, Int. J. Robust Nonlin. Control, vol. 12, pp. 841853, 2002.
control
[29] P. Apkarian, P. Gahinet, and G. Becker, Self-scheduled
of linear parameter-varying systems: A design example, Automatica,
vol. 31, pp. 12511261, Sep. 1995.
[30] P. Apkarian and R. J. Adams, Advanced gain-scheduling techniques
for uncertain systems, IEEE Trans. Control Syst. Technol., vol. 6, no.
1, pp. 2132, Jan. 1998.
-norm
[31] F. Wu, X. H. Yang, A. Packard, and G. Becker, Induced
control for LPV systems with bounded parameter variation rates, Int.
J. Robust Nonlin. Control, vol. 6, pp. 983998, 1996.
[32] F. Wu and S. Prajna, SOS-based solution approach to polynomial LPV
system analysis and synthesis problems, Int. J. Control, vol. 78, no. 8,
pp. 600611, 2005.
[33] P. Gahinet, A. Nemirovski, A. J. Laub, and M. Chilali, LMI control
toolbox, The MathWorks, Inc., Natick, MA, 1995.

Xiukun Wei (S05A07) received the M.E. degree


from Lanzhou University of Technology, Lanzhou,
China, in 1995, and the Ph.D degree from Johannes
Kepler University, Linz, Austria, in 2006.
Currently he is a PostDoc Researcher at Delft
Center for System and Control, Delft University
of Technology, Delft, The Netherlands. From 1995
to 2002, he was a Lecturer at Beijing Institute of
Machinery, Beijing, China. From 2002 to 2006, he
held a Research Assistant position at the Institute
of Design and Control of Mechatronical Systems,
Johannes Kepler University. His research interests include nonlinear system
identification, robust control theory, fault tolerant control and their applications
in a variety of fields such as Automotive control, offshore wind turbines, etc.

Luigi Del Re (S79M93) received the Dipl.El.Ing.


and the Dr.Sc.Techn. from the ETH Zrich, Zurich,
Switzerland.
From 1984 to 1990, he was with ETH Zurich
as a Group Leader in hydraulic control and, from
1991 to 1994, an Oberassistent at the Institut fr
Automatik. After working as Head of Hybrid Control
at Swatch, Biel, Switzerland, in 1998, he went to the
Johannes Kepler University, Linz, Austria, where
he is currently a Full Professor and Head of the
Institute of Design and Control of Mechatronical
Systems. His research interests include nonlinear control of complex systems
using analytical approximations, with applications mainly in the automotive,
hydraulic, and biomedical field.

Vous aimerez peut-être aussi