Vous êtes sur la page 1sur 9

1940

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 21, NO. 5, SEPTEMBER 2013

Adaptive Tracking Control of a Common Rail Injection System for Gasoline


Engines: A Discrete-Time Integral Minimal Control Synthesis Approach
Umberto Montanaro, Alessandro di Gaeta, Member, IEEE, and Veniero Giglio, Member, IEEE

Abstract Over the last decade, gasoline direct injection


engines have proven to be a promising solution to reduce both
emission and fuel consumption. The achievement of superior
performance strongly relies on its fuel injection system based
on the common rail (CR) device. In order to tame the CR
pressure dynamics without any a priori knowledge of the plant
parameters, we design a novel model reference adaptive control strategy that extends the discrete-time minimal control
synthesis algorithm. Indeed, an explicit discrete-time adaptive
integral action is added to improve closed-loop performance.
Experimental results support the analytical proof of stability,
and confirm the effectiveness of the novel algorithm to solve
both the regulation and the tracking control problem in a wide
range of working conditions. The closed-loop performance is
quantitatively evaluated via engineering indices.
Index Terms Automotive control, common
discrete-time model reference adaptive control.

rail

(CR),

I. I NTRODUCTION

HE MAIN objective of a common rail (CR) system is


to feed the electroinjectors with fuel at high pressure
independently from the amount of fuel to be sprayed. In
fact, high-pressure injections allow a fine atomization of the
fuel spray and promote a better air/fuel mixing, resulting
in significant combustion improvements. The CR injection
system, which was originally introduced for diesel engines,
but has been extended to gasoline engines, allows a precise
spray of a given amount of fuel directly into the combustion
chamber. The resulting gasoline direct injection (GDI) engine
technology is now a viable solution to reduce fuel consumption
and pollutant emissions. Quantitatively, by means of the GDI
engines, it is possible to reduce fuel consumption by 20%,
improve power output by about 10% while reducing cold start
unburned HC emissions by approximately 30% (see [1] and
references therein).
Rail pressure is a key variable for achieving this performance with GDI engines. A stable pressure value is fundamental to exactly determine not only the required fuel
mass to inject (depending on engine speed and load), but
also the expected spray features, such as droplet size, spatial
distribution, and penetration, as well as the fluid dynamics
interaction with in-cylinder air motion. This has led to the
desired fuel distribution (stratified or homogeneous) in the
Manuscript received July 25, 2011; revised February 3, 2012 and June 19,
2012; accepted August 25, 2012. Manuscript received in final form September 21, 2012. Date of publication November 21, 2012; date of current version
August 12, 2013. Recommended by Associate Editor U. Christen.
The authors are with the Istituto Motori, National Research Council,
Naples 111-80131, Italy (e-mail: u.montanaro@im.cnr.it; a.digaeta@im.cnr.it;
v.giglio@im.cnr.it).
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/TCST.2012.2220772

combustion chamber [1]. A robust combustion process with a


low variability requires the control of the pressure in the rail in
accordance with some injection strategies, see [2]. This control
objective is not easy since plant parameters strongly vary as a
function of the engine working condition. Furthermore, sudden
pressure variations occur due to the opening and closing of the
injectors.
In the automotive literature, different successful modelbased control approaches have been proposed to tackle the rail
pressure dynamics. For example, just to name a few, in [3], a
time varying internal model-based controller is designed, while
in [4], the quantitative-feedback theory is used to provide a
certain degree of robustness to the closed-loop system. In [5],
a hybrid model of the CR fuel injection system for a four
cylinder multijet engine is presented, and a hybrid controller
is designed and compared to a classical regulator. To face the
parameters variation, a model-based gain-scheduling strategy
has also been recently presented in [6].
Even though model-based control strategies are efficient to
regulate the pressure in CR devices, their performance strongly
relies on the knowledge of the plant model. Hence, their
robustness with respect to disturbances and parameter variations is a critical issue. Furthermore, a satisfactory parameter
identification of a CR device as required by those strategies is
time consuming and customized for just a single product.
In order to deal with the CR plant parameter variations due
to different working conditions, parameter uncertainties, and
disturbances, while reducing in addition the time-to-market
of innovative products, it seems reasonable to design and
implement adaptive control strategies where the control gains
are properly varied, or better adapted, according to the actual
system behavior. Nevertheless, this approach has only been
marginally investigated in the existing literature on the control
of CR injection systems.
The aim of this brief is to fill this gap in the control applications by proposing a novel discrete-time adaptive
strategy that extends the family of the so-called minimal
control synthesis (MCS) algorithms. The MCS algorithm for
continuous time linear plants is a model reference adaptive
control (MRAC) strategy that goes back to the nineties [7],
and it was introduced as an extension of the Landau model
reference adaptive scheme [8]. It has the advantage of relying
on a minimal knowledge of the plant dynamics and being
robust to nonlinear unmodeled terms and disturbances. One
drawback is that unwanted dynamics can emerge in closed
loop when the classical MCS strategy is implemented as a
digital controller by using standard discretization methods [9].
To solve this problem, the MCS algorithm has been recently
extended to discrete-time systems [10] and its applicability to

1063-6536 2012 IEEE

MONTANARO et al.: ADAPTIVE TRACKING CONTROL OF CR INJECTION SYSTEM

the continuous time case has been proven when both the plant
and reference model dynamics are discretized by means of the
forward Euler discretization method [10], [11].
In this brief, we design a novel MRAC algorithm named
discrete-time MCS algorithm with integral action (DTMCSI),
that extends the class of the discrete-time MCS controllers by
adding an explicit discrete-time adaptive integral term to the
control action proposed in [10]. As in the continuous time case
[12], the advantage of introducing this further control action
is the compensation of nonzero mean bias terms of the plant
not only in steady state conditions, but also during transients.
Furthermore, if a locking method is adopted to avoid a possible
drift of the adaptive gains, the resulting control strategy has
a structure comparable with a conventional, but auto-tuned,
discrete-time proportionalintegral controller when the locking
is activated.
To prove experimentally that the novel algorithm is a viable
and effective solution to tame the complex CR dynamics, we
use the DTMCSI strategy to control the rail pressure averaged
over a combustion cycle in a commercial injection system
for GDI engines. Unlike other control methods for the CR
system proposed in the literature, no a priori knowledge of
the plant parameters is required, nor are their variation laws
as a function of the engine speed.
The experimental analysis has been carried out on a wide
range of working conditions by using a rapid control prototyping (RCP) station based on dSPACE hardware [6]. This system
allows the variation of the injection duration as a function of
an emulated intake manifold pressure, engine speed, and CR
pressure by exploiting the static model of the GDI injectors.
To measure quantitatively the effectiveness of the novel MCS
algorithm, we compute different performance indices for a
wide set of experimental data.
II. CR S YSTEM
The CR system for spark ignition engines, schematized in
Fig. 1, is fundamentally formed by two separated sections: 1) a
low-pressure (LP) circuit consisting of a fuel tank, an LP pipe,
and a fuel filter; and 2) a high-pressure (HP) circuit, including
an HP pump, an HP line with a pressure sensor, a pressure
regulator valve, and the injectors. The LP electropump (2)
forces the fuel, through the filter (3), from the tank (1) to the
HP pump (4) with a pressure of 5 bar. The HP mechanical
pump increases the pressure of the fuel (up to 140 bar), and
pushes it into the common manifold (5) (named CR) equipped
with the electroinjectors (6). This pump is driven by the engine
through the camshaft, and it does not require phasing since
both the start and the duration of the injection are controlled
by the engine control unit (ECU). Finally, the pressure in the
manifold is controlled in closed loop by using the pressure
sensor (7) and the solenoid electrovalve (8) that allows the
excess fuel to return to the tank.

1941

Fig. 1. CR injection system. 1. Fuel tank. 2. Low-pressure pump. 3. Fuel


filter. 4. High-pressure pump. 5. Fuel rail. 6. Electroinjector. 7. Pressure sensor.
8. Pressure regulation electrovalve.

form when the plant parameters are unknown. Namely, given


a plant of the form
x(k + 1) = Ax(k) + Bu(k)

where x
and u R are the state and the input of the
system, respectively, with n N being the dimension of the
state space, and the system matrices A Rnn , B Rn1 are
in control canonical form

0
1 0 ... 0
0
0 1 ... 0

..
..
.. . .
..
(2a)
A= .
. .
.
.

0
0 0 ... 1
a1 a2 . . . . . . an

T
B = 0 0 ... 0 b
(2b)
where only the sign of b is supposed to be known.
The control objective is to impose to (1), the dynamics of
a given asymptotically stable reference model
x m (k + 1) = Am x m (k) + Bm r (k)

The DTMCSI algorithm solves the model reference control


problem for discrete-time linear systems in control canonical

(3)

with x m Rn , r (k) R being some desired reference signal


and the matrices Am and Bm given in the same canonical form
as that of the plant.
The control action provided by the DTMCSI algorithm is
given as follows:
u(k) = u MC S (k) + u I (k)

(4)

with
u MC S (k) = L(k)x(k) + L R (k)r (k)
u I (k) = L I (k)x I (k)

(5a)
(5b)

and the adaptive control gains computed as


L(k) =
L R (k) =

III. DTMCSI A LGORITHM

(1)

Rn

i=0
k

ye (i + 1)x T (i ) + ye (k + 1)x T (k) (6a)


ye (i + 1)r (i ) + ye (k + 1)r (k)

i=0
k

L I (k) = I

i=0

ye (i +1)x IT (i )+ I ye (k +1)x IT (k)

(6b)

(6c)

1942

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 21, NO. 5, SEPTEMBER 2013

where , , I , and I are scalar adaptation weights with the


same sign of b in (2b)
x I (k) =

x e (i )

(7)

i=0

is the integral tracking error, and


x e (k) = x m (k) x(k)
(8)
ye (k) = BeT Px e (k) R

T
where Be = 0 . . . 0 1 Rn and P is a symmetric positive
definite matrix so that
T
P Am P = Q,
Am

Q = Q T > 0.

(a)

(b)

Fig. 2. Regression results for the CR plant parameters (a) c(N ) and (b) d(N )
as third-order polynomial functions of the HP pump speed.

(9)

Remarks:
1) Typically all the adaptive gains are initialized to zero,
i.e., L(0) = L 0 = 0, L R (0) = L R0 = 0, and L I (0) =
L I0 = 0.
2) The term u MCS (k) is that proposed in [10]. Hence, the
DTMCSI algorithm improves the discrete-time MCS
strategy by adding an explicit integral control action
(5b). (As shown in the experimental results, this further
control action plays a fundamental role for controlling
the complex dynamics of the CR system.)
3) Equation (9) is always feasible since the reference
model (3) is assumed to be asymptotically stable.
4) The adaptive mechanism (6) at the time k needs the
knowledge of the output tracking error at the time
k + 1, namely ye (k + 1). This problem, known in the
MRAC literature as one delay problem, can be solved by
estimating the sample ye (k + 1) with effective methods
as those proposed in [8] and [10].
5) As the discrete-time MCS in [10] and [11], also the
DTMCSI can be applied to continuous time systems
when the forward Euler method is used for their discretization. We note that, even though the forward Euler
method does not guarantee the stability of the discretized
plant, it is not restrictive since the DTMCSI strategy can
also be applied to unstable plants and the stability of the
discretized reference model can always be assured as it
is chosen by the designer.
The proof of asymptotic stability of the closed-loop error
system when the novel adaptive algorithm is applied, is given,
in the Appendix. Its applicability to continuous time systems
discretized with the forward Euler method can be carried out
as in [10] and [11], while the robustness of the proposed MCS
strategy with respect to bounded disturbances can be proven
as in [10]. Therefore, they are not presented here for the sake
of brevity.

necessary for control synthesis and its digital implementation.


Nevertheless, for the sake of clarity and to emphasize the
time-varying nature of the plant parameters and the nonlinear
disturbances acting on the CR dynamics, here we present
briefly the CR model proposed in [13]. According to this
choice and after some algebraic manipulations, the CR is
modeled as a first-order system whose dynamics are given
by




Vb
c N R
p+c(N)

u + t; pa , N, Tinj
(10)
p =
N c L
100L
where p (bar) is the pressure in the rail (state of the system,
x := p), u (%) is the duty cycle of the pulse width modulation
signal used to actuate the electrovalve (control input), Vb (V)
is the battery voltage, L (H), and R () are the inductance
and the electrical resistance of the coil of the electrovalve,
respectively, N (r/min) is the rotational speed of the highpressure pump, that is equal to N = Ne /2 with Ne (r/min)
being the engine speed, and c(N) (bar/A) is the third-order
polynomial of Fig. 2(a). The term  is a nonlinear function
of the mean pressure, pa (bar), the duration of injection, Tinj
(ms), and the pump speed. It is modeled as


c N R
d
(d(N)+(t))+
(t) =
N + (t)

(11)
N c L
N
where d(N) (bar)
polynomial of Fig. 2(b),

is the third-order
and the term t; pa , N, Tinj (bar) models the disturbance due
to the operation of the HP pump and the pressure variations
caused by the intermittent functioning of the injectors. In this
brief, (t) is considered as an unknown disturbance acting on
the linear part of the plant.
See [13] for further details on the model (10) and its experimental validation. Here, we emphasize that the knowledge of
the plant parameters, their dependence on the engine speed,
and the knowledge of the disturbance (t) are not required to
successfully implement the DTMCSI strategy.

IV. M ATHEMATICAL M ODEL OF


THE CR S YSTEM

V. S YNTHESIS OF THE DTMCSI C ONTROLLER FOR THE


CR P RESSURE C ONTROL P ROBLEM

The adaptive law discussed in Section III extends the


discrete-time MCS approach and, thus, it relies on a minimal
knowledge of the plant. Hence, it can be implemented easily
without requiring time-consuming experiments for the precise
characterization of the system dynamics. For this reason,
an accurate mathematical model of the plant is not strictly

A. Fulfillment of the Control Canonical Form Assumption


The first step for the design of the DTMCSI adaptive controller is to verify if the plant dynamics match the hypothesis
of control canonical form. It is apparent from (10) that the
CR model satisfies this assumption, hence the novel adaptive
control strategy is applicable.

MONTANARO et al.: ADAPTIVE TRACKING CONTROL OF CR INJECTION SYSTEM

1943

B. Control Specification and Reference Model


The control specifications are mainly related to requirements
on the system dynamic response and they are set as follows:
1) the settling time is required to be = 4 cc (where cc
denotes the engine cycles) and 2) no overshoot should be
present in the step response.
With respect to the control specification 1), we remark
that it is common for automotive control systems to specify
control objectives in the crank-angle domain instead of the
time domain. In doing so, the performance of the controlled
plant is the same independently of the duration of the engine
cycle. In particular, a settling time of 4 cc guarantees a short
off-design operation of the injectors with a maximum of 3 cc
for a step variation of the pressure set point. This is the worst
case if we consider that in commercial ECUs, the reference
pressure is scheduled on the basis of the engine speed that
never varies as an ideal step function. The second control
objective, on the other hand, is required to guarantee a smooth
combustion process by avoiding unwanted phenomena like
impingement of the piston top and cylinder wall, which imply
reductions of the engine performance in terms of emission and
engine torque [3].
To achieve the control specifications we choose as continuous time reference model, the following first-order system
xm = F x m + Gr , with F = 4.6 6Ne /(720 ) and G = F.

Fig. 3. Schematic diagram for the generation of the injection duration used
for the experimental tests.

In doing so, only the parameter has to be selected for


the tuning of the entire adaptive control strategy. Here,
we have chosen = 1 from a finite set of values so
that the mean value of the tracking error is minimized.
We point out that the adaptive weights are all negatives
in accordance with the sign of b of the input matrix (2b),
indeed b := c(N)(Vb /100L) < 0 N due to the negative
sign of c(N) as shown in Fig. 2(a).
4) A first-order filter is introduced to limit the tracking error
during rapid variations of the reference signal. Moreover,
to limit noise on the feedback signal, a further first-order
filter, whose bandwidth is adapted as a function of the
engine speed, is applied on the instantaneous pressure
p(t).

C. Details on the Implementation of the DTMCSI Strategy


In implementing the control action (4), the following issues
have to be considered.
1) The reference model is discretized with a sampling
time Ts of 1 ms, which guarantees stable discrete-time
reference dynamics.
2) The one-delay problem is solved following the approach
in [8] and [10]. Hence, we set
ye (k +1)

ye (k)


1+pnn b(+)Ts x (k)x(k)+r 2(k)

(12)

with pnn being the entry of the matrix P in (9) in


position (n, n). This choice is claimed to be a good
prediction if the sampling period is small enough [8].
Notice that the computation of ye (k + 1) according
to (12) also requires an estimate of the uncertain plant
parameter b. Again in accordance with [8] and [10], b
should be replaced by a value that falls within its range
of variation. Experimental results reported in Section VII
will show that, despite its simplicity, this choice yields
a successful implementation of the control strategy.
3) As usually happens when implementing the MCS strategy [7], the scalar quantities, which modulate the adaptive gains in (6), have to be chosen heuristically as a
tradeoff between convergence time and reactivity of the
control action. According to the MCS empirical rule [7],
we set / = 0.1. For further reducing the parameters to
be tuned, we use the same rule for the integral weights,
i.e., I / I = 0.1. Furthermore, we select I / = 0.08
so that the adaptive integral action improves the closedloop performance without being the dominating one.

VI. C ALCULATION OF THE I NJECTION D URATION


As shown in Section IV, the disturbance (t) acting on the
plant depends on the engine speed (or HP pump speed) and the
duration of the injection. In order to generate a realistic disturbance for the testing of the control strategy, we implemented
experimentally the injection duration management sketched in
Fig. 3.
Briefly, the injection duration is scheduled so that the airto-fuel ratio is kept at the stoichiometric value, S (14.56 for
a commercial gasoline). To this aim, the mass m d to inject is
m ac /(S Ncyl ), with m ac being the total inlet air mass of the
engine and Ncyl the number of cylinders. The speed-density
function f sd (Ne , Pi ) proposed in [14] for a 2.0 liter fourcylinder GDI engine is used for the inlet mass estimation
based on the engine speed Ne and intake manifold pressure Pi
(mbar). Once the fuel mass to be injected has been computed,
it is mapped to an injection duration by exploiting the inverse
model of the injector g(m d , pa ) as indicated in Fig. 3. (Further
details on the steady flow model for GDI injectors can be
found in [6].)
We note that in our RCP environment, Pi is a variable that
can be set to different values in order to emulate different
engine loads, while Ne is twice the HP pump speed as in real
vehicles.
The feedback of the mean value pressure pa to compute
the injection duration introduces a nonlinear coupling between
the variable to be controlled and one of the source of the
disturbance. This coupling can generate further unwanted
dynamics that must be rejected by the controller.

1944

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 21, NO. 5, SEPTEMBER 2013

(a)

(b)

Fig. 4. Experimental results. (a) Intake manifold pressure. (b) Engine speed.

(a)

(b)

(c)

(d)

Fig. 5. Experimental results. (a) Model reference pressure (red dashed line)
and mean value pressure (black solid line). (b) Percentage pressure tracking
error. (c) Control action. (d) Adaptive gains, L (blue dashed line), L R (red
dotteddashed line), and L I (black solid line).

VII. E XPERIMENTAL R ESULTS


In this section, the controller designed in Section V is
experimentally tested to show its effectiveness in steering the
mean value of CR pressure despite the motion of the HP
pump and the opening and closing of the injectors. Note that
in what follows the average value, pa , of the rail pressure
is computed by averaging the instantaneous rail pressure,
p(t), over one engine cycle with a sampling step of 18
crank-angle degrees. Moreover, all the experimental activities
have been carried out on a high-pressure injection system
(Bosch) for GDI engines. The HP mechanical pump is actuated
by a three-phase induction motor of 1.5 kW whose speed
has been electronically controlled by an AC drive (Model
COMMANDER SE 23400220 by Control Techniques). (The
reader is referred to [6] for further details on the experimental
test rig.)
First, we show the closed-loop performance when the intake
manifold pressure and the engine speed are those in Fig. 4(a)
and (b), respectively, and the duration of injection is generated
according to the strategy described in Section VI. We note
that, when the intake manifold pressure increases (air load),

(a)

(b)

Fig. 6. Experimental results. Model reference pressure (red dashed line)


and mean value pressure (black solid line) (a) regulation performance and
(b) tracking performance.

the engine speed increases up to about 6000 r/min for a load of


1000 mbar. Analogously, when the intake pressure decreases
the engine speed decreases as well.
The demanded injection pressure and the closed-loop performance when the DTMCSI is inserted in the loop are shown in
Fig. 5(a). Notice that, to emulate a real working condition, the
demanded injection pressure increases when the engine speed
increases. The maximum demanded rail pressure is 100 bar
for engine speeds around 6000 r/min.
To better appreciate the effectiveness of the proposed adaptive control to tame the CR dynamics when plant parameters
persistently vary due to engine speed variations, Fig. 6(a) and
(b) shows a zoom of regulation and tracking performance,
while Fig. 5(b) shows the percentage tracking error over the
entire manoeuvre. As clearly pointed out in Fig. 5(b), the
DTMCSI provides a percent tracking error always below 3.5%.
This is a remarkable result taking into account that neither
plant parameters nor a detailed plant model have been used
for the design of the control action. Hence, we expect that such
performance can be achieved independently from the particular
CR device.
For the sake of completeness, we report the control action
and the evolution of the adaptive gains in Fig. 5(c) and (d),
respectively. More in detail [Fig. 5(c)], the control action is
always in the admissible range (0%100%) and saturations
never occur. It is worth pointing out here that the control gains
evolve so that tracking is achieved despite the presence of
unmodeled terms and unknown disturbances that persistently
act on the plant. The variations of the control gains are therefore the key not only for imposing the tracking of reference
dynamics but also for rejecting these disturbances and the plant
parameter variations especially when the engine speed varies
from low to high speed and vice versa (t [80, 130]) as
shown in Fig. 4(b). The drift observed in Fig. 5(d) is mainly
due to the persistent excitation of the reference input and
the plant parameter variations. Ad hoc locking strategies or
parameter projection methods available in the control literature
can be implemented to keep the control gains in a preassigned
range.
To investigate in depth the tracking and regulation capabilities of the DTMCSI algorithm for the control of CR
dynamics, in the following we test the closed-loop performance under different working conditions. In more detail,

MONTANARO et al.: ADAPTIVE TRACKING CONTROL OF CR INJECTION SYSTEM

1945

TABLE I
L IST OF S IGNALS T HAT C OMPOSE THE R EFERENCE
I NPUT TO THE C ONTROL S YSTEM
Signal

Period

Amplitude

Bias

bar

bar

Constant

25

S2

Square wave

15

45

28

S3

Sinusoidal wave

20

60

21

S4

Sequence of steps

S5

Square wave

15

45

28

S6

Sinusoidal wave

20

60

21

S7

Sinusoidal wave

40

60

16

S1

Type

Duration

we test the control algorithm when the engine speed Ne


Ne  {1000, 2000, 3000, 4000, 5000, 6000} r/min and intake
pressure Pi Pi  {500, 750, 1000} mbar in order to mimic
low, medium, and high load conditions. (Hence, 18 experiments have been carried out.) For each experiment, we chose,
as reference input, a signal formed by concatenating signals
listed in Table I. In particular, the sequence of steps (signal
S4) is composed of eight rising/falling steps of amplitude 20
bar starting from 20 bar and ranging to a maximum of 100 bar
and back. Hence, this piecewise constant signal takes values
s = {20, 40, 60, 80, 100} bar.
in Pinj
To measure quantitatively the control performance, for each
operating point belonging to Pi Ne the following performance indices have been computed.
1) To evaluate the steady state accuracy when the reference
pressure is a constant, we chose as performance index
the root of the mean squared error, I2 , computed in
steady state regime as


1 t +T
I2 =
( pinj pa (t))2 dt
(13)
T t
where the initial time t and the time interval T are
chosen so that the closed-loop system is in steady state
for t [t , t + T ] and pinj is the selected pressure set
point.
2) To measure the closed-loop settling time, the following
percentage index is introduced:
exp mod
(14)
s = 100
mod
where exp and mod are the settling time observed
experimentally and that of the reference model, respectively, both expressed in engine cycles.
3) The tracking capability of the DTMCSI controller is
tested by means of the mean percentage tracking error
defined as
 #
100 t +Tcs | p(t) pm (t),|
dt
(15)
E=
Tcs t #
pm (t)
with t # being the initial time instant of the generic signal
with duration Tcs (see also Table I) and pm (t) being the
pressure of the reference model.
4) Even if an in-depth comparison of the novel adaptive
controller with those proposed in the literature for the

(a)

(b)

(c)

(d)

Fig. 7. Experimental results. Performance indices as a function of the engine


speed (a) I2 -index, (b) s -index, (c) E-index, and (d) R-index.

CR control is out of the scope of this brief, for the sake


of completeness, we introduce an index to compare the
pressure ripple obtained with the DTMCSI strategy and
that observed with the model-based gain scheduling integral control (MBIC) strategy proposed in [6]. Precisely,
we compute the experimental percentage ripple variation
defined as
rDTMCSI rMBIC
(16)
R = 100
pinj
with rDTMCSI and rMBIC being the experimental residual
pressures provided by the DTMCSI algorithm and the
model-based strategy, respectively, while pinj is the
required pressure set point.
Notice that the indices R, s , and I2 are computed for each
working point and for each level of the signal S4, namely for
s P N (162 points),
each point belonging to  Pinj
i
e
while the E-index is evaluated for each working condition in
Pi Ne and for each signal of Table I (108 points).
The performance indices computed on the experimental data
are shown in Fig. 7 as a function of engine speed. In particular,
we note the following.
1) For any operating condition in , the I2 -index is always
below 0.8 bar, and for engine speed less than 6000 r/min
it never exceeds 0.35 bar [see Fig. 7(a)]. Hence, the
adaptive algorithm is effective to impose a given reference set point in a wide range of speeds. Notice that
the reference set points are achieved without overshoots
since the controller imposes to the plant the dynamics
of a first-order system. In Fig. 8, some step responses
are shown for pointing out this closed-loop feature.
2) Fig. 7(b) shows that the performance index s is less
than 20%, hence the closed-loop settling time is always
below 4.8 cc independently from the working condition.
This value is still acceptable taking into account both
the complexity of the plant dynamics and the simplicity

1946

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 21, NO. 5, SEPTEMBER 2013

engine. Finally, we point out that the turnaround time (i.e., the
execution time of the control task) observed for the DTMCSI
algorithm is similar to that of the MBIC controller, i.e., about
10 s on a 600-MHz RISC processor.
VIII. C ONCLUSION

(a)

(b)

Fig. 8. Experimental results. Steady state performance. Mean value CR


pressure (black solid line) pa , model reference pressure (red dashed line) pm ,
input to the reference model (blue dashed line) pinj for two cases belonging
to .
TABLE II
M EAN VALUE AND S TANDARD D EVIATION OF THE E -I NDEX C OMPUTED
FOR THE

S IGNALS IN TABLE I AND W ORKING C ONDITIONS IN Pi Ne


S2-S5

S3-S6

S4

S7

Mean Value [%]

2.56

2.23

0.91

4.15

Standard Deviation [%]

0.88

0.44

0.26

1.37

of the control structure. Furthermore, there are many


working points having a settling time smaller than that
required by the reference model (working points in
with s < 0). The mean value of the s index is 0.4%,
hence the mean settling time is about 4 cc.
3) As is apparent in Fig. 7(c), the E-index takes small
values independently of the working condition and the
signal to be tracked and it never exceeds 7%. Furthermore, the average E-index is 2.44% with a standard
deviation of 1.24%. Table II reports the mean value
and the standard deviation of the E-index computed
for each signal of Table I. It clearly shows that the
strategy provides good tracking performance over the
entire range of engine speeds and loads with a maximum
error of 4.15% for the tracking of a sinusoidal wave
with amplitude 40 bar, bias 60 bar, and period 16 s
(signal S7).
4) As shown in Fig. 7(d), both the DTMCSI algorithm and
the MBIC strategy provide comparable ripples. More
precisely, except at 5000 r/min, the adaptive algorithm
reduces the ripple in almost all operating conditions
(negative values of the R-index) by up to 10%. At
5000 r/min, the adaptive method amplifies the residual
pressure by a maximum of 6%.
We remark that model-based gain scheduling controllers are
still widely used in automotive engineering and implemented
in commercial ECUs for their low computational cost. Nevertheless, an effective tuning requires the knowledge of the plant
parameters and their performance strongly rely on the model
accuracy. We point out that, unlike the MBIC controller, the
DTMCSI algorithm does not need any a priori identification
of the plant parameters nor a feedforward action based on
the plant model and its parameters. This is fundamental
when controlling the CR device since plant parameters can
change from one car to another during the lifespan of the

In this brief, we presented an MRAC of a CR injection


system for GDI engines. The main advantage of the proposed
approach is that it only requires a minimal knowledge of
the plant and is robust to the nonlinear pressure perturbations without any model-based feedforward compensation. An
extensive experimental investigation showed that the proposed
adaptive algorithm is effective in controlling the CR pressure,
and it can be a viable alternative solution with respect to
more standard gain scheduling strategies without increasing
the turnaround time. Hence, it can represent a promising
compromise between simplicity of the control structure and
performance. Future research will be aimed at designing a
discrete-time robust adaptive action, as done for the continuous
time MCS algorithm, to improve further closed-loop performance in attenuating effects of external disturbances.
A PPENDIX
P ROOF OF A SYMPTOTIC S TABILITY
The proof relies on showing that the closed-loop error
dynamics (8) can be rewritten as a feedback hyperstable
system [8], [15], thus guaranteeing the asymptotic convergence
of the tracking error. The proof is then organized as follows:
1) recast the error dynamics as a feedback system; 2) prove
that the feedforward path is strictly positive real; and 3) show
that the feedback block satisfies the Popov inequality. In what
follows each step of the proof is proven.
Step 1: Taking into account the canonical structure of
the system matrices, after some algebraic manipulations the
tracking error dynamics can be recast as
x e (k + 1) = Am x e (k) + Be (k + 1)
ye (k) = BeT Px e (k)

(17a)
(17b)

(k + 1) = (k)w(k)
T

w(k) = x T (k) r (k) x IT (k)
T

(k) =  X (k)  R (k)  I (k)


= 1 (k) . . . n (k) . . . 2n+1 (k)

(18a)

where
(18b)
(18c)

with  X (k) = BeT (Am A) bL(k),  I (k) = bL I (k), and


 R (k) = bm bL R (k), where bm is the last entry of Bm .
System (17a) and (17b) is the negative feedback
interconnection of the LTI block with dynamic matrix Am ,
input matrix Be , and output matrix BeT P, while the feedback
path is the nonlinear time-varying block that maps ye (k) onto
(k) (see Fig. 9 for a schematization of the tracking error
system).
Step 2: The transfer function of the feedforward path is
H (z) = z BeT P(z In Am )1 Be

= BeT P Be + BeT P Am (z In Am )1 Be .

(19)

MONTANARO et al.: ADAPTIVE TRACKING CONTROL OF CR INJECTION SYSTEM

1947

where the pair ( j , j ) is



(, ),
( j , j ) =
( I , I ),

j = 1, . . . , n + 1
j = n + 2, . . . , 2n + 1.

(26)

From (25a) and (25b), we get that the term ye (k +1)w j (k)
can be expressed both as

1 
j (k) j (k 1) (27a)
b j
j (k)
.
(27b)
ye (k + 1)w j (k) =
b j

Fig. 9. Closed-loop error dynamics represented as an equivalent feedback


system.

According to [8, Lemma B.4-2, p. 376], a discrete transfer


1
B, is strictly
function of the form H (z) = D + C z In A

positive real if there exist a symmetric positive matrix, Q,


and matrices K and L of appropriate dimensions such that:
2) B T P A + K T L T = C,
and
1) A T P A P = L L T Q,
T
T
T
3) D + D B P B = K K .
In our case, the strictly positive realness of the transfer
function (19) can be shown straightforwardly by choosing
A = Am , B = Be , C = BeT P Am , D = BeT P Be , Q = Q,
L = 0, and K = P 1/2 Be with P the solution of (9).
Step 3: To complete the proof, it suffices to prove the
fulfillment of the following Popov inequality:
l

ye (k + 1) (k + 1) c2 ,

cR

(20)

k=0

with l being a generic time instant.


given a generic vector
=
T
 In what follows,
m
1 2 m
R we define as  () the m-dimensional
vector:
T

(21)
 () = 0 1 2 m1 .
Notice that, from (21), it is trivial to show that
 ()  

(22)

ye (k + 1) (k + 1) =

2n+1

S j (l)

(23)

j =1

k=0

ye (k + 1) j (k)w j (k).

(24)

k=0

Now, it is evident that inequality (20) is satisfied if each term


S j can be made greater or equal to some constant c2j , j =
1, . . . , 2n + 1.
To this aim taking advantage of the integral-proportional
form of the adaptive gains (6), the generic component j (k) of
the vector in (18c) can be decomposed as j (k) = j (k) +
j (k), j = 1, . . . , 2n +1, where the integral and proportional
adaptive components are, respectively
j (k) = j (k 1) b j ye (k + 1)w j (k)
j (k) = b j ye (k + 1)w j (k)

S j (l) =

l


1

j (k) j (k) j (k 1)
b j
k=0

l
1
2
j (k).
b j

(28)

k=0

Now defining
j = [j (0) . . . j (l)]T

(29)

and taking into account definition (21) we have






1 
j 2 T  j j (0)j (1)
S j (l) =
j
b j
+

l
1
2
j (k).
b j

(30)

k=0

By using the Schwarz inequality and (22), the following


inequalities hold:
  T      2

T
 j j
 j  j j  j
(31)
j
hence

(32)

Since sgn( j ) = sgn( j ) = sgn(b) for assumption, we have


b j > 0, b j > 0, then from (30) and (32) we obtain

1 
j (0)j (1) := c2j
b
l
2n+1

ye (k +1) (k +1)
c2j := c2 .

S j (l)

with
S j (l) 

Taking into account (27a) and (27b), after algebraic manipulations the generic term S j (l) can be written as

 2  2


T
T
j
 j j j j
 j 0.

where  is the Euclidean norm.


The summation (20) can be decomposed as follows:
l

ye (k + 1)w j (k) =

(25a)
(25b)

(33)

j =1

k=0

Therefore, step 3 holds and then the convergence of the


tracking error to zero remains proven according to the hyperstability theorem [8, Th. C-4, p. 386].
R EFERENCES
[1] F. Zhao, D. L. Harrington, and M. C. D. Lai, Automotive Gasoline
Direct-Injection Engines. Warrendale, PA: SAE, 2002.
[2] Y. J. Park, Injection pressure controlling method of gasoline direct injection engine, U.S. Patent 6 349 698, Dec. 12,
2002.
[3] Z. Zhang and Z. Sun, Rotational angle based pressure control of a
common rail fuel injection system for internal combustion engines, in
Proc. Amer. Control Conf., Jun. 2009, pp. 26902695.

1948

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 21, NO. 5, SEPTEMBER 2013

[4] W. Chatlatanagulchai, T. Aroonsrisopon, and K. Wannatong, Robust


common-rail pressure control for a diesel-dual-fuel engine using QFT
based controller, in Proc. SAE Tech. Pap., 2009, no. 2009-01-1799.
[5] A. Balluchi, A. Bicchi, E. Mazzi, A. L. S. Vincentelli, and G. Serra,
Hybrid modelling and control of the common rail injection system, in
Hybrid Systems: Computation and Control (Lecture Notes in Computer
Science), vol. 3927, J. P. Hespanha and A. Tiwari, Eds. Berlin, Germany:
Springer-Verlag, 2006, pp. 7992.
[6] A. di Gaeta, U. Montanaro, G. Fiengo, A. Palladino, and V. Giglio, A
model-based gain scheduling approach for controlling the common-rail
system for GDI engines, Int. J. Control, vol. 85, no. 4, pp. 419436,
Feb. 2012.
[7] D. P. Stoten and H. Benchoubane, Empirical studies of an MRAC
algorithm with minimal controller synthesis, Int. J. Control, vol. 51,
no. 4, pp. 823849, 1990.
[8] Y. D. Landau, Adaptive Control: The Model Reference Approach. New
York: Marcel Dekker, 1979.
[9] O. S. Bursi, D. P. Stoten, N. Tondini, and L. Vulcan, Stability and
accuracy analysis of a discrete model reference adaptive controller
without and with time delay, Int. J. Numer. Methods Eng., vol. 82,
no. 9, pp. 1158117, May 2010.

[10] M. di Bernardo, F. di Gennaro, J. M. Olm, and S. Santini, Discrete-time


minimal control synthesis adaptive algorithm, Int. J. Control, vol. 83,
no. 12, pp. 26412657, Dec. 2010.
[11] M. di Bernardo, A. di Gaeta, U. Montanaro, J. M. Olm, and S. Santini,
Discrete-time MRAC with minimal controller synthesis of an electronic
throttle body, in Proc. 18th IFAC World Congr., Milan, Italy, Aug. 2011,
pp. 50645069.
[12] D. Stoten, The adaptive minimal control synthesis algorithm with
integral action, in Proc. IEEE 21st Int. Conf. Ind. Electron., Control,
Instrum., vol. 2. Nov. 1995, pp. 16461651.
[13] A. di Gaeta, G. Fiengo, A. Palladino, and V. Giglio, A control oriented
model of a common-rail system for gasoline direct injection engine, in
Proc. 48th Joint IEEE Conf. Decision Control, Dec. 2009, pp. 6614
6619.
[14] A. di Gaeta, U. Montanaro, and V. Giglio, Model-based control of
the air fuel ratio for gasoline direct injection engines via advanced cosimulation: An approach to reduce the development cycle of engine
control systems, J. Dynamic Syst., Meas., Control, vol. 133, no. 6, pp.
061006-1061006-17, Nov. 2011.
[15] V. M. Popov, Hyperstability of Automatic Control Systems. New York:
Springer-Verlag, 1973.

Vous aimerez peut-être aussi