You are on page 1of 12

S

r
J. David Powell, N. P. Fekete, and Chen-Fang Chang
here has been a continuing effort to improve the efficiency and exhaust emissions of the automobile engine for the last three decades. A key development in the evolution was the introduction of closed loop control in the 70s [36]. Initially, the closed loop control algorithms were very similar to what was previously implemented mechanically; however, as the electronics transitioned to increasingly capable microprocessors through the 90s, it became practical to entertain the use of modern control and estimation algorithms in a production vehicle. These algorithms were proposed for automotive engine control in 1978 by Athans [3], in 1980 by Cassidy et al., [9], in 1983 by Powers et al., [3S], and in 1987 by Kiencke [29]; however, during those times it was not realistic to expect that such complexity could be put in a production engine and engine designers did not embrace the methods. In the 90s, there has been increasing interest in the methods by many researchers with attention given to the practical implementation in an engine microprocessor and experimental evaluations. Specific applications of observers for AFR control based on measurements in the intake manifold (airflow or manifold pressure sensors) include the research by Bosch [7], a team led by Hendricks at the Technical University of Denmark [22], 1231, [26], the University of Wisconsin, [30], and Siemens, [ 11. Another approach is to base the observer on measurements of the exhaust using 0, sensors and on the throttle position. This approach has been researched by ETH in Zurich, [34], l391,byStanfordUniversity[21, [lo], I l l ] , [121,[161, [17],and by Daimler-Benz, 181, [17], [181, 1191usinglinearobservertheory. In addition, the University of California at Berkeley, [14], [ 151, has addressed the same problem using a nonlinear, sliding mode approach. This paper discusses the application of linear observer theory to engine control with a specific focus on observers based on exhaust measurements. The motivation for this architecture is that a direct measurement of the relevant quantity (the exhaust chem-

istry entering the catalyst) will typically lead to the most accurate control. A secondary motivation is the need to reduce costs. If sufficient accuracy can be obtained with the exhaust sensor providing the primary information, the cost associated with intake manifold sensors can be reduced. The basic theory is discussed and experimental results are reported that illuminate the benefits and superior performance of linear observer-based control.

Observer Concepts
The basic theory behind observerbased control follows from Kalmans original work [28] The key ideas in his work were to include a model of the process as part of a state estimator and to compute the gains of the estimator based on disturba nc e noise st atistics Luenberger [311 suggested a simphfica$ tion where the estimator gams were made constant and the gains were selected to achieve arbitrary estimator dynamics He referred to the result as arl observer The application of these ideas has become standard for many control system designers and the methods are described in most current day control system textbooks, for example [21]. In automotive engine control, these methods are often referred to as model-based control The interesting aspects of the application of observer-based control to engine AFR control are twofold 1) there is a pure delay between the plant and the sensor due to the engine cycle and the exhaust transport time, and 2) the primary disturbance is the throttle, which can be accurately measured if a drive-by-wile throttle is incorporated. These two features of engine AFR control create opportunities for significant improvements to the control performance by utilizing observer methods

Systems with Delay Using classical control, the existence of a pure delay in the system introduces a limitation on the system bandwidth which, in turn, limits the transient accuracy of the system response. Engineers designing classical control of AFR in an engine recog-

Powell is with the Dept. ofMechanica1 Engineering, Stanford University Stanford, CA. Fekete is with Daimler-Benz AG, Stuttgart, Germany. Chang is with General Motors Research & Development Centel; Warren, MI.

72

0272-1708/98/$10.00019984EEE

IEEE Control Systems

nized this and introduced feedforward to eliminate much of the error. The feedforward terms were tuned, but were difficult to make accurate under all dynamic conditions. However, observer-based control formalizes the process, exposes what the important dynamics are, and shows exactly how to apply the feedforward portion of the control. The use of an observer reconstructs the state of the system, including the state elements before the delay, and allows for the application of control as if no delay were present. Therefore, for a perfect model, no error will result due to the measured throttle motion or AFR command inputs. More detail on general features of observer-based control with a pure sensor delay is contained in Section 8.6.1 of [21]. Drive-by-wire Throttle (DBW) The actuator or control input to an AFR controller is the fuel injection quantity. By far the biggest disturbance to this AFR controller is the throttle, which changes the air flow. The controller is tasked to adjust the fuel flow accordingly so as to follow a prescribed air-fuel ratio. Unlike most control systems, this throttle disturbance can be measured and its effect included in the control algorithm. But, being a digital control system, the throttle is sampled at discrete points. For a throttle that is mechanically connected to the drivers foot, its motion between samples is not known precisely and, therefore, its effect on the airflow is not precisely known. For all the results in this paper, the throttle was actuated electronically (called drive-by-wire) so as to make its motion between samples precisely known. A driver would not perceive this modification to the throttle motion because of the fast sample rates. Identification of the role of the throttle and the importance of its proper control was discussed in the 70s by Stivender [37]. Event-based Sampling The AFR is a discrete event for each cylinder. The contents of the intake manifold flow into the cylinder until the intake valve closes, at which time the AFR for that cylinder is fixed until the next engine cycle. Furthermore, since the fuel injection time is constrained to be at a specified point in the engine cycle for optimum emissions performance, it is highly desirable to have the control system sample rate synchronous with the engine cycle so that the model relating the air flow and throttle motion is as accurate as possible. For these two reasons, all the control systems described here will use a sample rate that is synchronous with the crank shaft motion, commonly referred to as event-based sampling. Because the airflow dynamics into the intake manifold are fundamentally a time-based phenomenon, those dynamics will need to be converted to a discrete model with a time-varying sample rate based on the engine speed ( N ) .Other dynamics have been shown to be less varying as a discrete model with eventbased sampling [ 131. To transform a time domain differential equation into the crank angle domain, note that

dt=-.

de 6N

For a general quantity x(t), its rate of change in the crank angle domain lis obtained by substituting the following identity into the time-dolmain equation:
dx dx -=6N--. dt dB

(3)

The AFR Control Model


The dynamics that relate the effect of the control actuator (the amount of fuel injected) to the measured output (the exhaust gas oxygen (EGO) sensor) consist of the intake manifold airflow, fuel flow, cycle delays, and the exhaust mixing and delays. Fig. 1 shows a sketch of the system for the single cylinder case. Air Flow One simple model of the air flow in an intake manifold is the filling and emptying model. The air flow enters the manifold through the throttle and is pumped out of the manifold into the cylinder. Assuming no leaks, the air flow ni,, into the manifold and the flow ni, out of the manifold are identical only in steady state. During throttle transients, the difference between these two flows equals the rate of change of the air mass in the manifold plenum. Assuming that the manifold pressure is uniform and the intake manifold temperature (T,) is constant, the continuity equation and ideal gas law can be applied to the manifold plenum. The air filling dynamics are expressed in terms of the manifold pressure p , [lo], [241,

(4)
where V, is the engine volume and V, is the intake manifold volume. This is not a linear differential equation since the air flow rate through the throttle plate,ni,,, is a nonlinear function of manifold pressnre p,. Volumetric efficiency q v ,which measures the

Drivers Command Engine Controller

EGO

Exhaust

1
Fig. 1. The Engine control system.
Control Signal Sensor Signal

where N is engine speed in RPM and 8 is crank angle in degrees. Therefore, crank angle and time are closely related as

October 1998

73

pumping performance of the cylinder, inlet port and valve, also depends on p,. The flow rate into the cylinder is given by

Equations (4) and ( 5 ) represent a first order dynamic system that describes the mass flow in and out of the manifold ignoring inertial effects. The form of the combined equations is
Fig. 2. Intake manifold ai$ow model.

where map is the mass of air inducted per intake stroke and m ss is the steady-state value ofm,. The time constant, ,,, C I , andm are experimentally determined functions of throttle angle (a) and engine speed (N) in order to allow for the possibility of some inertial influence. Integrating the flow yields the total air charge in the cylinder (mac) during one cycle, since
55

It has been found that for a relatively slow, single-cylinder research engine used for some experimental verifications, the manifold filling time constant was sufficiently fast compared to one engine cycle that the air flow dynamics could be neglected and the manifold filling dynamics were assumed to be instantaneous. In this case, the intake air flow was an algebraic function of throttle angle and engine speed, modified by the airflow bias estimate,
m,,(k) = m , ( a , N ) + ma,(/<)

ha, = K(a,N,At)m,, ,

( 7 )
(11)
In general, however, this approximation cannot be made and non-zero time constants were used for the automotive engine experiments.

where the K's are adjusted to give the appropriate air flow into each cylinder as a function of the time through the engine cycle, At, as determined by experiment. Note that the throttle, a, is the primary disturbance input to the system and is a major input to the value of m is. The m ss function would typically be a table lookup based on prototype testing. The actual values of the airflow in a given vehicle would be different because of production variations, environmental conditions, and aging. Therefore, a correction term to the table is also incorporated into the model and continually estimated based on the actual measurements from the exhaust sensor. For purposes of this discussion, we will assume the correction, mob,is a constant bias; however, in the multicylinder engine evaluations this correction was a table where the entries were estimated separately for different engine operating points as discussed further in the section on parameter identification. The discrete model of (6) and (7) is
m,(k+l) =(e-""""'~)m,(k)+(l - e - T ' 2 - ) m s s ( a , N )

Fuel Flow
Previous research indicated that the fuel jet from the injector mixes with the air stream and also deposits on the walls of the intake system components.The fuel vapor and airborne droplets are transported with the fast-moving air stream. The liquid fuel flows along the walls and also enters the air stream through reentrainmenu evaporation with a time constant on the order of a few engine cycles, although it is much longer during a cold start. At steady state engine operating conditions, the fuel droplet deposition on the wall surface balances the fuel reentrainmenuevaporation into the air stream. Then, the amount of fuel entering the cylinder is the same as the amount of fuel injected. The behavior of the liquid film of a fuel injection system was modeled by Acquino as [4],

(8)

where T is the sample time. For multicylinder engines, the sample time is selected to be on the order of 45" of crank travel to integrate the airflow accurately throughout the intake process. For the experiments with automotive engines, the airflow dynamics were significant. In fact, the time constant was found to be different for filling and emptying, i.e., the sign of map,as well as the functionality for mss and K as indicated above. The airflow model for the multi-cylinder case is depicted in Fig. 2 from [ 171where it was shown that the model fit experimental data quite well.

where
inip: liquid

fuel puddle mass

kfi: metered fuel flow rate

fuel flow rate entering the cylinder X: fraction of the metered fuel that enters the puddle I C ~ time : constant of the fuel evaporation process. A discrete fuel puddle model which describes the same physical phenomenon and better reflects the intermittent nature of the fuel injection is [32]:
m,:

74

IEEE Control Systems

m,,(k) = f,mrp(k) + f,W

(15)

The fuel-injector model assumes a dead-time (to)and a linear relationship between mass of fuel injected ( m p )and injector open time, with proportionality constant K,:
m,, = K,. ( LJ - to 1,

where k: engine cycle index mtt:fuel mass injected m p : fuel mass into the cylinder per cycle f, : fraction of injected fuel that enters the cylinder directly each cycle f s : fraction of fuel puddle that evaporates and enters the cylinder each cycle. Therefore, the fuel-air ratio in the cylinder ism,.. I macand the fuel-air ratio normalized to the stoichiometric value (equivalence ratio) is:

(20)

where ttn,is the injector pulse-width. Therefore, this relationship aIlows the computation of the injector pulse width from the desired value of m p . For purposes of the control model, we will consider mjito be the control input to the process. The Combined Model For the singlecylinder case, (lo), ( l l ) , (14), (15), (16), (17), and (1 8) abovecan be collected to yield a set of difference equations describing the dynamic process:

where ( A / F ) is the stoichiometric air-fuel ratio, typically about 14.7:l for gasoline. It is also common to use the reciprocal, i.e., the air-fuel ratio (AFR) which, normalized to the stoichiometric value, is called h where h = 1I 0.
~

Cycle/transport Delay There is a delay between when the fuel injection occurs and when the corresponding exhaust gas reaches the oxygen sensor. It has been found by experiment that this delay is very close to one full engine cycle. Although it is not necessary to limit this delay to an even cycle, doing so simplifies the model considerably. The equivalence ratio ( O d ) is delayed relative to the in-cylinder equivalence ratio ( by one cycle and is expressed as:
0,(k+1) = 0Jk)

where

(17)

For the mutli-cylinder case, there is a each cylinder.

qd state element for

Exhaust Mixing and AFR Sensor Dynamics For multicylinder engines, the mixing of the exhaust from the different cylinders in the exhaust manifold is modeled as a recursive average of the sequential qdoutputs from each cylinder. The EGO sensor (either the linear UEGO version or the nonlinear, switch-type h-sensor) measures the oxygen concentration of the mixed exhaust gas. The UEGO sensor is modeled as a first-order lag with a time constant(T,). The state element associated with the sensor measurement ( 9 , ) is updated once per engine cycle:

Note thatmacappears in the denominator, causing these to be nonlinear equations. For the case where the airflow dynamics are significant, there are the additional equations, (8) and (9)

where 5 = e-* and T is the sample time. For the nonlinear h-sensor, the device is essentially full-rich or full-lean and responds with a time constant that is sufficiently fast to be ignored. Therefore, the model for control purposes is a relay. In other words,

which would be added to (21). The dynamic relations are again nonlinear; however, for purposes of observer design, it does not pose a difficulty because the changing airflow is typically slow compared to the cycle by cycle variations in the AFR and the designs can be carried out by linearizing about specific airflows. For the case where the sensor is the h-sensor, (21c) is replaced with (19). This is shown in Fig. 3 and will be discussed further in the next section. The standard form of a linear dynamic system is often written as

y( k ) = Hx( k ) .

(23b)

October

1998

75

Driver Input

1
Engine

I
, ?

this will be accomplished if

AFR Process
I I I I

I I I

I I
I I

The mass of fuel in the cylinder is a function of the injected mass and the fuel-puddle mass (15), and therefore the mass of fuel which must be injected can be calculated from:

Observer

Fig. 3. Observer-based control structure.

TheA observer provides the estimates of the $-cylinder airmass (mn,)and of the mass of the fuel-puddle (mfJ which are For the AFR model, the input, U( k ) is the fuel injected, m p ( k ) used to calculate the control input, mii. Note that if there were no and y is the exhaust sensor output, 41~. The state vector (for the observer error, the AFR in the engine would respond instantaneously to a 41r command because m p enters the cylinder mixed single cylinder case with the UEGO sensor) is with the predicted value of the air for that cycle from the observer. Because the model is not perfect, a proportional feedback The control is described of $ is also used to minimize the @error. in more detail in [12].

and the state transition matrix, a, from (21) is

Observer Design For a system as described by (23)-(27), an observer to estimate the state, x, given a measurement, Qm,is [21]
measurement update:

r
The input matrix is

l-fa

0 01

% k ) = x ( k ) + U$$,,,(k)-?m(k))

(32)

time update: F ( k + l ) =@?(k)+rmfi(k).


(33)

r=
and the output matrix is H=[O 0 1 0 1 .

The observer gain matrix, L, was selected in order to provide estimation with time constants on the order of a few engine cycles. Although the system is nonlinear and has time-varying coefficients, it was found that constant values of L yielded acceptable dynamics over the entire engine operating region. Combined Controller The control law and the observer are combined as shown in Fig. 3. Although linear observer theory does not specifically address on-off sensors such as the h-sensor, in fact, one can include such a device by inserting the same functionality in the observer model as occurs in the physical process, as indicated in the figure. A requirement for using this sensor is that +r = 1on average. In use over a long period, sensor degradation may decrease the effectiveness of this scheme; however, the experiments performed (discussed below) showed a remarkable tolerance to sensor characteristics that were quite different from the on-off model assumed in the observer. The experiments show results for both the linear UEGO sensor and the on-off h-sensor. For the linear UEGO sensor, the block diagram of the control structure would not include the relay shown in the h-sensor box in either the process or the observer.

Control Law and Observer Design


Control Law The control law is designed so that the in-cylinder equivalence ratio (qC) tracks a commanded value ( Qr).Therefore, we wish that

at each sample k. Since

76

IEEE Control Systems

Parameter Identification
The parameters in the control were determined experimentally for the experimental results to follow. However, the observer-based structure lends itself to adaptive control or on-line identification which eliminates the need for detailed engine calibration and ensures that the accuracy of the control can be maintained throughout the lifetime of the engine. Since the principal dynamics which influence the AFR in a warm engine are the airflow dynamics, a technique was developed to identify the most important parameters in the airflow model. This identification technique can be used to adapt the model parameters as the engine characteristics change with time. The steady-state airmass table is essentially adapted by the bias estimate,m,, .A particular feature of the throttle controller is that there is a built-in delay between the command and the response, which can vary as a function of actuator temperature, and this delay time can also be identified, although it was experimentally determined once for the following tests. Related system identification techniques have been applied to the engine control problem by Auk, et al., [5], [6], [27] and by Turin and Geering [38],[39]. Steady-State Adaption The steady-state control is adapted using feedback from the exhaust sensor. The bias which must be added to the nominal air-table in order for the measured equivalence ratio ( 9 , ) to be equal to the models estimate ( Qm)is estimated while the engine i; running at steady-state, for each point in the table. This bias (ma b ) is stored in a look-up table as a function of speed and load by reverse bi-linear interpolation. The steady-state air-table adaption ensures that the control is adjusted to compensate for modeling errors and slow disturbances such as changes in ambient pressure or temperature, load sensor biases, or engine aging. These steady-state adaption schemes have been described in the literature [12], [17], [25], [ 3 3 ] ,and also exist in production [20], [29]. Intake-Manifold Time-Constant The intake-manifold time-constant is a strong function of the operating point. For the 4-cylinder engine, it varied from about 300 msec at light load to less than 5 msec at full load. The local time-constant at each operating point was determined by exciting the intake-manifold dynamics with a throttle ramp about the operating point. A linear least-squares method was then used to identify the model parameters which result in the best fit of the model to the input/output data consisting of the throttle position and the smoothed intake-manifold pressure signal. The intake-manifold pressure signal was filtered to remove the cyclic oscillations and the high-frequency noise. A smoother consisting of a forward and a backward filter pass is used to eliminate the phase-lag between the original and the smoothed signals [17l, [161. Fig. 4 shows the intake-manifold time-constant as a function of engine speed and throttle position that was generated by this identification technique. Note the strong dependence of the time-constant on the throttle position. For throttle angles greater than about 40, the throttle is essentially open, the engine is running at full-load and, therefore, the intake-manifold time constant is very short.

Experimental Results
Experiments have been performed using a 0.6L laboratory single-cylinder engine using both a linear UEGO and switchtype h sensor and using a 2.2L automotive 4-cylinder engine using a UEGO sensor that could run at h = 1or lean. Both engines were port fuel-injected and had drive-by-wire (DBW) throttles. The 4-cylinder engine was a Mercedes-Benz production engine that was modified so it would run well while lean (h > 1). The experiments on the single-cylinder engine were all run at constant engine speed because it was difficult to change the speed of that engine quickly. Furthermore, throttle changes in an actual engine happen faster than speed changes, so this simplification to the test procedure was judged to be appropriate. In the subsequent tests on the 4-cylinder engine, speed changes were introduced and the engine was run through the entire operating regime as called for by emissions testing. Single-Cylinder Engine with an UEGO Sensor Using the control structure as shown in Fig. 3 without the h-sensor relay function, the observer was implemented as shown using the model given by (21). The airflow was assumed instantaneous as indicated by (1 l), and them sp function was mapped as previously discussed. The fuel model constants were selected to achieve the best AFR control, the best parameters being fa = 0.8 and f p = 0.7. The experiments had an engine speed, N , of 1200 rpm. The results for a square wave throttle input are shown in Fig. 5 while those for a somewhat arbitrary throttle time history are shown in Fig. 6. Both experiments show AFR control accuracy on the order of 0.5%rms. In Fig. 5 , note the spikes in the injection time, t , , , that are compensating for the dynamics of the fuel puddle. Without this correction from the observer, there would be AFR spikes of several percent. Further detail on these experiments are contained in [ 101 and [12]. Single-Cylinder Engine with a h-Sensor Using thie control structure as shown in Fig. 3 including the h-sensor relay functions, the observer was implemented as shown in the figure using (21), but with (21c) replaced by (19) in order to model the on-off sensor. The airflow was assumed instantaneous as given by (11),the fuel model constants were again fa = 0.8 and fp = 0.7, and N = 1200 rpm. The results for a con-

10

20

30

40

50

60

70

80

90

Throttle [deg]

Fig. 4. Intake-manifold time-constant.

October 1998

77

stant throttle input are shown in Fig. 7 while those for the arbitrary throttle time history are shown in Fig. 8. For purposes of control evaluation, an UEGO sensor was installed in the engine so the AFR performance could be monitored, but was not used for control purposes. Its output is also plotted in the figures and labeled as @uego. Again, both experiments show AFR control accuracy on the order of 0.5% rms even with the on-off sensor that only provided the binary information of "too rich" or "too lean." Note in the figures that the control sensor output, $ ,, cycles between rich and lean while the actual AFR in the engine stayed very close to the desired value of @ = 1as measured by the UEGO sensor. It is well known that control systems using an on-off sensor typically exhibit a limit cycle; in fact, all production automobiles have a limit cycle in the AFR at about 2-5 Hz because of their on-off A-sensor. It is remarkable that the use of an observer can be used to eliminate the limit cycle. An explanation of why it performs so well is that the observer is reconstructing the state by performing a time-average of the on-off signal from the sensor and the controller is utilizing the state estimate for the control rather than the raw on-off signal. Whether it is desirable to eliminate the limit cycle behavior for automotive AFR control is another matter. In fact, it has been determined that it is indeed beneficial for catalyst efficiency for the AFR to oscillate around 2-5 Hz. As a result, designers have purposely picked the control parameters to yield that optimum

limit cycle frequency from the standpoint of catalyst efficiency. The difficulty is that an engine's operating point affects the limit cycle frequency and amplitude; therefore, obtaining the best limit cycle characteristics for all conditions is not easy and could degrade the AFR control precision. With an observer-based de-

20

40

60

80 100 120 140 160 180 200 Engine Cycle

7 1 0

20

"

40

60

"

80

100 120 140 160 180 200 Engine Cycle

"

"

'

1.I
1.05

rms: 0.46%

2
00

1
0.95 -

35;

20

40

60

80 IO0 I20 Engine Cycle

140 160 I80 2d0

Fig.6. I-Cylinder AFR control using an UEGO sensor during an


arbitrary throttle history.
'0
20 40

60

80

100 120 140 160 180 200 1.1 .. .

Engine Cycle

1.05 -

,051
1

rms: 0.43%

1
0.95 -

8 k 8

O"O

10

20

30

40

50

60

70

80

90

100

Engine Cycle O.OI5

5
20

1.1

1.05

0.005

2 '
0.95 40 60
80 100 120 140 160 180 200 Engine Cycle

'0

0.9;

10 '

20 '

30 '

40 '

50 ' 60 ' Engine Cycle

70 '

80 '

90 '

100 I

Fig. 5. I-Cylinder AFR control using an UEGO sensor during throttle square waves.

Fig. 7. I -CylinderAFR control using a h sensorfor constant throttle.

78

IEEE Control Systems

sign, it is possible to command a "dither" oscillation at any desired frequency and amplitude. Designers then have the freedom to pick independently the oscillation frequency and amplitude on the one hand, and control parameters on the other hand. Therefore, it is possible to optimize both catalyst performance and control accuracy. As an example of this, Fig. 9 shows the AFR performance with a sinusoidal 9, command at one frequency while the throttle is undergoing a square wave time history at a different frequency. Note that the performance of the system as indicated by the monitored value, Quego, follows the command, Qr,to 0.6% rms. Further detail on these experiments are contained in [2], [I21 and [16]. Four-Cylinder Engine with an UEGO Sensor Using the control structure as shown in Fig. 3 without the h-sensor relay function, the observer was implemented as shown using the model given by (21) and (22). The m ss function was mapped as discussed earlier. The fuel model constants were selected to be f a = 0.85 and f o = 0.75.

Results with StoichiometricAFR Control


The performance of the controller at stoichiometric AFR ( h = 1) is shown by Figs. 10 and 11. Fig. 10 shows a throttle square wave with load transients from 23% to 45% of full-load at 2000 RPM. The AFR control with and without the observerbased controller is compared. In the uncompensated case, the

control calculation is based purely on the steady-state air-map, with no dynamic compensation and no feedback. There is a steady-state error in h relative to the command (A, = = 1.00) because there is a bias in the air-table, and there are large excursions in h (15%-I 8%) during the transients because the dynamics have not been compensated. The second trace shows the control with the dynamic compensation included. The steady-state air-table estimation (mab) eliminates any steady-state errors. When the model for the intake-manifold filling and emptying dynamics is included in the control calculation, the h excursions during the transients are reduced by almost 90%. The remaining h excursions are due to the fuel-puddle dynamics, and when the fuel-puddle compensation is also added, the AFR is regulated to 0.4% rms. These results show that the air-flow dynamics have a much more significant influence on the AFR than do the fuel-flow dynamics, and the time-constant of the intakemanifold filling and emptying is the most important dynamic parameter. In Fig. 11, the throttle executed a square-wave and the engine speed was varied by commanding the dynamometer controller. The fuel-injector pulse-width (and hence the air-flow) changed by 88%, and the h excursions were 0.75% rms. This shows that varying engine speed degraded the performance to some extent, but still provided excellent control of the AFR. More details are available in [17].
$7

20

40

60

80 100 120 140 160 180 200 Engine Cycle

L
3s0

io i o io i o
Engine Cycle

liol;o

2b0

-.b
1.1
1.05

20

40

60

80 I00 120 I40 160 180 2b0


Engine Cycle

rms: 0.58%
.

8
43

1
2

1.1

rms: 0.60%

1 0.95 0.9

0.95

0.9 0

20

40

60

80 100 120 140 160 180 200 Engine Cycle

1.05

1.05
&

l
0.95
0.9;

8 0.95

20

"

40

60

"

' I 80 100 120 140 160 180 200 Engine Cycle


" "

0.9'

20

"

40

60

"

80 100 120 140 160 180 200


Engine Cycle

"

"

'

Fig. 8. 1-Cylinder AFR control using a h sensor during an arbitrary throttle time history.

Fig. 9. I-Cylinder AFR control using a h sensor with a ditheredqr (0.5 Hz) during a square wave throttle time history.

October 1998

79

20

Results with Lean AFR control


The engine was also run at h , = 1.48 ( $ r = 0.676), which is near the lean limit for the conditions tested. Modifications to the engine for these tests consisted of changing the fuel injection timng, the spark timng, and re-adapting them s.l table to account for the slight difference in air-flow which is due to the cooler engine operating temperature when running lean. All of the dynamic parameters in the engine model remained the same. In fact, the intake-manifold time-constant was found to be essentially independent of AFR. Fig. 12 shows the performance during a throttle square wave with load transient from 29% to 58% of full-load at N = 2000 rpm. The AFR was controlled to 0.69% rms and there were no misfires, even though the engine was running near the lean limit. The figure also shows the Indicated Mean Effective Pressure (IMEP) for each cylinder, which is a standard measure of the output of each cylinder that can be made by the use of a cylinder pressure sensor. The consistency of these traces indicate that the distnbution of the AFR between the cylinders was quite good in addition to the overall AFR as measured by the UEGO sensor. More details are available in [17].

18

E
a , -

16

14

12
RPM = 200s BMEP = 2.1 bar to 4.1 bar
10

10 Time [sec]

15

1.2

1.I

7
Y

It is desirable to vary the AFR during operation for the best combination of efficiency and exhaust emissions. A lean AFR 1 improves the fuel efficiency of an engine, but the NO, emissions worsen unless a NO, storage catalyst is used. It has been found 1: that it is possible to trap the NO, in an adsorbent catalyst during lean periods and then to reduce the NO, by momentarily switchI ing to a richer mixture, [SI, [18]. However, during the AFR i 0.8 switch, it is important to transition quickly because the NO, 0 5 10 15 emissions are maximum at h = 1.1+ 1.2.Observer-based control Time [sec] has been found to allow much faster AFR switching and has, Fig. 10. 4-Cylinder AFR control with h, = 1, throttle square wave, therefore, significantly reduced the NO, emissions. Fig. 13 shows the results of tests performed at Daimler-Benz on the 4and N = 2000 rpm. cylinder lean burn engine. The curves on the left side of the figure show the performance with the observer-based controller as described above. The engine was initially running with h = 1.48, then it switched to h = 0.86 for 10 seconds. The curves on the right are for a standard mechanical throttle with conventional feedback from an UEGO sensor. Companng the quality of the h control on the left and right, we see a I IO' I 1000; much tighter control about the operating 5 10 15 20 0 5 10 15 20 points and a much sharper transition with Time [sec] Time [sec] the observer-based control. This superior h control produced a much lower level of 1.1 RMS = 0.75% NO, before and after the catalyst as shown by the upper traces. More details are avail1.05 able in [8]. In order to incorporate AFR switches into an engine controller so that the driver i does not feel engine output jumps, it is necessary to move the throttle simultaneously 0.95 I I I so that the engine torque output is constant 0.9 0 5 10 15 20 0 5 10 15 20 through the AFR switch. With the use of a Time [sec] Time [sec] DBW throttle, this is readily accomplished, as demonstrated in Fig. 14. The Fig. 11. 4-Cylinder AFR control with h, = 1, throttle square wave, and varying N .
X

Results with Mixed AFR Command Values


I

- 1

t'

80

IEEE Control Systems

figure shows a time history of the controlled value of AFR with steps between h = 15and 1 . The DBW controller also included a feedforward control so that the Brake Mean Effective Pressure (BMEP) was maintained constant, no matter what the AFR. As

t :

10

20

30

40

50

60

Time [sec]

RPM = 2001

- Cylinder

- Cylinder 2 . - Cylinder 3
. . . Cylinder 4

10

20

30 Time [sec]

40

50

60

2
OO

10

20

30

40

50

60

Time [sec]

7 1 -

Fig. 14. AFR steps with throttle control demonstrating constant engine output.
100 200 Engine Cycle 300

1.6

Lambda Command = 1.48 RMS = 0.69%

can be seen, the controller adjusted the throttle so that the output on the bottom trace was constant. More details on the throttle control scheme are contained in [19].

Conclusions
Designing AFR engine control with state observers (or model-based control) aids in providing improved control. It provides a framework that makes it possible to essentially elimiTI m nate the deleterious effects of the pure delay in the system and a 1.5 5 has been found to provide improved accuracy and faster re1 sponse. The improved results shown in the paper were based on experiments on one and four cylinder engines that included a 1.45 drive-by-wire throttle and a sampling period that was synchronous with the crank angle (or event-based) as opposed to the conventional time synchronous sampling scheme. The improved 1.4 results are the result of these features and the observer-based de0 100 200 300 sign method. Engine Cycle Experimental evaluations were performed with laboratory Fig. 12.4-CylinderAFRcontrol with h = 1.48, throttle square wave port fuel injected ngle and multi-cylinder engines using Exand N = 2000 rpm. haust Gas Oxygen (EGO) sensors. Both a linear EGO sensor (UEGO), and a nonlinear, switch-type EGO sensor (h-sensor), as Conventional . Model Based typically used in automobiles today, were & . AFR Control AFR Control 0 . 3000demonstrated and evaluated. For (Not Optimized) 8 stoichiometric control, the AFR followed the desired value within 0.5% during throttle transients at various operating points. The observer-based method was shown to eliminate the limit cycle that normally results from the use of an on-off sensor, thereby enabling designers to optimize the AFR dither for catalyst efficiency. The scheme was also applied to a four-cylinder lean-burn engine where accurate AFR control is particularly beneficial because of the need to avoid overlean conditions that may cause misfire and its associated efficiency Fig. 13. Comparison of observer-based and conventional control during AFR step losses and high emissions. The tests with the 4 cylinder engine showed that it was commands.
1.55

October 1998

81

possible to follow various commanded AFR values (lean and stoichiometric) to within 0.8% during throttle transients at aiOuS operating points. These m d t s are Significantly better than those achievable with conventional Control Structures and have demonstrated significant improvements in emissions for lean burn engines with NO, storage catalysts by allowing fast switching between lean and rich operation.

[14] S. B. Choi, Design ofaRobust ControllerforAutomotive Engines: Theory and Experiment, Ph.D. dissertation. U. of Calif., Berkeley, Dec. 1993. [15] S. B. Choi and J. K. Hedrick, An Observer-Based Controller Design Method for Improving AirEuel Characteristics of Spark Ignition Engines, IEEE Trans. Contr syst; Technol., May 1998. 1161 N. P. Fekete, Model-Based Air-Fuel Ratio Control of a Multicylinder Leanbun Engine, PhD thesis, Stanford University, Stanford, CA, Dec. 1994. [17] N.P. Fekete, U. Nester, I. Gruden, and J. D. Powell, Model-Based AirFuel Ratio Control of a Lean Multicylinder Engine, SAE Paper No. 950846, 1995.

Acknowledgements
The authors would like to thank all their colleagues at Stanford University and Daimler-Benz AG who helped contribute to these research projects. They helped make this work possible. Special thanks go to Alois Amstutz, V.K. Jones, Brian Ault, Gene Franklin, U. Nester, I. Gruden, andD. Voigtlaender who directly worked on various aspects of observer-based control with the authors. In addition, the authors would like to acknowledge help with facilities as well as many fruitful discussions with Ron Patrick and AI Eaton at Stanford. Helpful discussions with Elbert Hendricks at the Technical University of Denmark are also gratefully acknowledged. Last but not least, the work would not have been possible without financial support from Nissan, GM, Ford, and Daimler-Benz.

[I81 N.P. Fekete, et al, Evaluation of NO, Storage Catalysts for Leanburn Gasoline Fueled Passenger Cars, SAE Paper No. 970746, 1997.
[19] N. P. Fekete, et al, Advanced Engine Control and Exhaust Gas Aftertreatment of a Leanburn SI Engine, SAE Paper No. 972873, 1997. [20] G. Felger, 0. Gloeckler, H. Kauff, U. Kiencke, H. Knapp, and H. Stocker, Regulating Device for a Fuel Metering System, US Patent No. 4440131, 1982. [21] G. F. Franklin, J. D. Powell, and M. L. Workman, Digital Control of Dynamic Systems, Addison-Wesley Longman, 3rd ed., Menlo Park, CA. 1998. [22] E. Hendricks, and S. C. Sorenson, SI Engine Controls and Mean Value Engine Modeling, SAE Paper No. 910258, 1991. [23] E. Hendricks, T. Vesterholm, and S. C. Sorenson, Nonlinear, Closed Loop, SI Engine Control Observers, SAE Paper No. 920237, 1992. [24] J. B. Heywood, Internal Combustion Engine Fundamentals. McGrawHill, New York, 1988.
[25]J. Ishii, N. Kurihara, M. Shida, H. Miwa, and T. Sekido, An Automatic Parameter Matching for Engine Fuel Injection Control, SAE Paper No. 920239, 1992.

References
[ 11 E. Achleitner, W. Hosp, A. Koch, and Schurz Electronic Engine Control

System for Gasoline Engines for LEV and ULEV Standards, SAE Paper No. 950479, 1995. [2] A. Amstutz, N. P. Fekete, and J. D. Powell, Model-Based Air-Fuel Ratio Control in SI Engines with a Switch-Type EGO Sensor, SAE Paper No. 940972, 1994. [3] M. Athans, The role of Modern Control Theory for Automotive Engine Control, SAE Paper No. 780852, 1978.

[4] C. F. Aquino, Transient A & Control Characteristics of the 5 liter Central


Fuel Injection Engine, SAE Paper No. 810494, 1981. [5] B. Auk, System Identification and Air-Fuel Ratio Control of a SparkIgnition Engine. PhD thesis, Stanford University, Stanford,CA, June 1994. [6] B. Ault, V. K. Jones, J. D. Powell, and G. E Franklin, Adaptive Air-Fuel Ratio Control of a Spark-Ignition Engine, SAE Paper No. 940373, 1994. [7] N. F. Benninger and G. Plapp, Requirements and Performance of Engine Management Systems under Transient Conditions, SAE Paper No. 9 10083,
1991.

[26] P. B. Jensen, M.B. Olsen, J. Poulsen, E. Hendricks, M. Fons, and C. Jepsen, A New Family of Nonlinear Observers for SI Engine AFR Control, SAE Paper No. 970615, 1997. [27] V. K. Jones, B. A. Ault, G. F. Franklin, and J. D. Powell, Identification and Air-Fuel Ratio Control of a Spark-Ignition Engine, IEEE Trans. Contr Syst. Technol., March 1995. [28] R. E. Kalman, A New Approach to Linear Filtering and Prediction Problems, J. Basic Engineering, vol. 85, pp. 34-45, 1960. [29] U. Kiencke, The Role of Automatic Control in Automotive Systems, in Proc. IFAC 10th Triennial World Congress, vol. 3, pp. 203-211, 1987. [30] M. Kao and J. J. Moskwa, Nonlinear Cylinder and Intake Manifold Pressure Observers for Engine Control and Diagnostics, SAE Paper No. 940375, 1994. [31] D. G. Luenberger, Observing the State of a Linear System, IEEE Trans. Mil. Electronics, vol. MIL-8, pp. 74-80, 1964. [32] T. Matsumura and Y. Nanyoshi, New Fuel Metering Technique for Compensation Wall Flow in a Transient Condition using the ModelMatching Method, Int. J. Veh. Design, vol. 12, no. 3, pp. 315-323, 1991. [33] S. Nakaniwa, J. Furuya, and N. Tomisawa, Development of NestStructured Learning Control System, SAE Paper No. 910084, 1991. [34] C. H. Onder and H. P. Geering, Model-Based Multivariable Speed and Air-to-Fuel Ratio Control of an SI Engine, SAE Paper No. 930859, 1993. [35] W. E Powers, B. K. Powell, and G. P. Lawson, Applications of Optimal Control and Kalman Filtering to Automotive Systems, Int. J. Veh. Design, Applications of Control Theory in the Automotive Industry, SP4, pp. 39-53, 1983.

[8] M. S. Brogan, N. P. Fekete, et al, Evaluation of NOx Storage Catalysts as an Effective System for NO, Removal from the Exhaust Gas of Leanbum Gasoline Engines. SAE Paper No. 952490, 1995. [9] J. F. Cassidy, M. Athans and W. H. Lee, On the Design of Electronic Automotive Engine Controls using Linear Quadratic Control Theory, IEEE Trans. Contr Syst. Technol., vol AC-25, no. 5, pp 901-912, Oct 1980. [lo] C.-E Chang, Air-Fuel Ratio Control in an IC Engine Using an EventBased Observer. PhD thesis, Mechanical Engineering, Stanford University, Stanford, CA, June 1993.
[ 111C.-F. Chang, N. P. Fekete, and J. D. Powell, Engine Air-Fuel Ratio Con-

trol using an Event-Based Observer, SAE Paper No. 930766, 1993.

[12] C.-E Chang, N.P. Fekete, A. Amstutz, and J. D. Powell, Air-Fuel Ratio Control in Spark-Ignition Engines Using Estimation Theory, IEEE Trans. Contl: Syst. Technol., March 1995.
[13] Y.-K. Chin and F. E. Coats, Engine Dynamics: Time-Based Versus Crank-Angle Based. SAE Paper No. 860412, 1986.

82

IEEE Control Systems

[36] J. G. Rivard, Closed-loop Electronic Fuel Injection Control of the IC Engine, SAE Paper No. 730005, January 1973. [37] D. L. Stivender, Engine Air Control - Basis of a Vehicular Systems Control Hierarchy, SAE Paper No. 780346, 1978. [38] R. C. Turin and H. P. Geering, On-Line Identification of Air-Fuel Ratio Dynamics in a Sequentially Injected SI Engine, SAE Paper No. 930857, 1993. [39] R. C. Turin and H. P. Geering, Model-based Adaptive Fuel Control in an SI Engine, SAE Paper No. 940374, 1994.

N.P. Fekete was born in Montreal, Canada in 1962. He


received the B. Eng. and M. Eng. Degrees from McGill University in 1986 and 1990, respectively. He received the Ph.D. from Stanford University in 1995, with a specialization in automotive controls and electronics. Since 1995 he has been an employee of Daimler-Benz AG, where he developed model-based control strategies for leanburn SI engines. He is currently working on control algorithms for common-rail diesel engines. Dr. Fekete was a recipient of the 1994 SAE Arch T. Colwell Merit Award and the 1995 SAE Vincent Bendix Automotive Electronics Award.

J. David Powell received his B S degree in Mechanical Engineering from MIT in 1960 and his Ph D in
Aero/Astro from Stanford in 1970 He joined the Stanford Faculty in 1971 and is a Professor in the Aeronautics and Astronautics Department and the Mechanical Engineenng Department Much of his career has been directed toward engine control research while his current focus IS primarily applications of GPS to air and ground vehicles He teaches courses in dynamcs, navigation, and control and has co-authored two text books in control systems design

Chen-Fang Chang received the B.S. degree from National Tawan University, the M S degree from Pennsylvania State University, University Park, and the Ph D. degree from Stanford University,all In Mechanical Engineering with Ph D minor in Electncal Engineering. He is currently a semor research engineer at the General Motors Research and Development Center, working on the cylinder-pressure-based engine control system for highefficiency and low-emssions vehcles HISresearch interests are in autgmohve controls and electro-mechanical systems.

October 1998

83

You might also like