Vous êtes sur la page 1sur 7

Modeling of an Internal Combustion Engine for Control Analysis

Jeffrey A. Cook and Barry K. Powell


ABSTRACT: The process model for an internal combustion engine with spark ignition is inherently nonlinear. For purposes of control analysis, it is desirable to have a linear model. This paper presents a discussion of recent activity in nonthermodynamic modeling of automotive internal combustion engines and shows how adequate linear models can be developed for control analysis. recirculation (EGR) dynamics were developed by Garofalo [4], Prabhakar et al. [5], and Powell [6]. The Prabhakar model was one of the first attempts at a comprehensive representation in that it contained spark advance, throttle and fuel control variables, and empirically based approximations for exhaust emissions. A linear power plant model, complete with an approximation of emission behavior, was developed by Cassidy et al. [7] and was used for linear quadratic control design. More recently, linear quadratic control was applied by Kamei et al. [8] to a 23rd-order linear perturbation model. Static and dynamic dynamometer test results were used to estimate model parameters using a statistical identification method. Nonlinear dynamic models appropriate to wide speed and load operating ranges were developed by Powell [9], Delosh et al. [IO], and Dobner [l l]. Wu et al. [I21 developed a similar nonlinear model with experimentally based dynamic intake manifold fuel wall-wetting condensation effects. A bibliography and more thorough review of nonthermodynamic engine models is presented by Powell and Cook [13]; a recent comprehensive model review, including input-output models and physically based models, is presented by Powell [14]. Other engine models will be referenced in the following discussion of specific engine elements. isentropic compressible flow of an ideal gas. Air mass flow rate through the throttle body may be thought of as a nonlinear separable function of manifold pressure and throttle angle. The partial derivative of the air mass flow rate equation with respect to throttle angle 0 defines the linearized airflow rate sensitivity KOin the model of Fig. 2.

Intake Manifold Dynamics


The dynamic equations for the manifold system are developed by employing the principles of conservation of mass and thermodynamic energy, and assuming that the vaporized constituents satisfy the equation of state. Conservation of momentum is satisfied by assuming that a uniform pressure exists in the intake manifold between the throttle body and the intake valves. This latter assumption, although valid for a low-frequency dynamic representation, precludes simulation of high-frequency acoustic propagation. Yuen [I51 and Servati [I61 describe the dynamic equations that result from using these principles. A method of developing a manifold model that accounts for concentration variation due to EGR, fuel, and air is presented by Moskwa and Hedrick [17]. For certain applications, it may be assumed that temperature is approximately constant and that the dominant manifold intake mass rate is the airflow rate, lit. Making these assumptions provides a first-order differential equation relating the rate of change of manifold pressure ( P ) to the flow rates into (lit,) and out of ( M ) the manifold:

Introduction
The past several years have witnessed a good deal of engine model building activity in the automotive industry in a category referred to as control-oriented or control design models. These models are generally low-frequency engine representations with uniform pulse, homogeneous charge, and lumped parameter approximations of enginebreathing and rotational dynamics. This paper contains a brief review of the modeling literature and presents a fundamental nonlinear model of a spark ignition, internal combustion engine. A linear control-oriented model is derived from the nonlinear process, techniques for experimental verification are examined, and a practical linear engine example incorporating multirate sampling is illustrated.

Literature Review
Power plant characterizations for four-, six-, and eight-cylinder engines were developed in the early 1970s by Hazell and Flower [1]-[3]. These models and related analysis are significant in that Hazell and Flower were among the first to develop discrete models with sampling commensurate with engine crank-angle events, to develop approximations of the fuel-dependent torque by fulland partial-pulse representations, and to perform comparative parameter design and stability analyses. Eight-cylinder nonlinear representations with air-to-fuel ratio ( M F ) and exhaust gas

Linear Model Development


The foundation for this paper is the fundamental, nonlinear engine model developed by Powell [9], which is illustrated in Fig. 1. The model contains representations of the throttle body, engine pumping phenomena, induction process dynamics, fuel system, engine torque generation, and rotating inertia. The linearized version of the engine model is illustrated in Fig. 2. Subsequent paragraphs will proceed from the general nonlinear model of Fig. 1, briefly reviewing each subsystem and developing the linearized relationships illustrated.

Kp(lit,

ik)

(1)

where Kp is a function of the gas constant, gas molecular weight, specific heats, manifold volume, and nominal assumed manifold temperature. Recalling that the mass airflow rate is, in general, a function of throttle angle and manifold pressure P , and noting that the engine pumping mass flow rate, M, is a function of manifold pressure and speed, N , linearization of Eq. (1) yields the following:

An early version of this paper was presented at the 1987 American Control Conference, Minneapolis. Minnesota, June 10-12, 1987. Jeffrey A. Cook and Barry K . Powell are members of the Research Staff at Ford Motor Company, Dearborn. MI 48121.

A P = Kp(am,/aP

aikiap)Ap K,,K,vAN (2)

Throttle Bod4
The basic throttle body model is developed by assuming one-dimensional, steady,
0272-1 708/88/0800-0020 00 0 1988 IEEE $01

+ KpKOA$

where KN is the pumping feedback defined by ahflaN. Defining the inverse of the coef-

iU

I t Control Syiterns Mnqaiine

FUEL COMMAND

r
I mf
ENGINE POWER INERTIA
I

SPARK COMMAND

CHARGE MASS RATE

THROTTLE ma BODY

MANIFOLD e ENGINE PLENUM PUMP

- SPEED

AF

Gf

A '

G6

'

DISTURBANCE

A6

K O

TP

K P

' AP

rps

+1

G P

e-2ST

K R TR
TRS

AN
b

+1

AIR BYPASS COMMAND

MANIFOLD FILLING

240 IP LAG

ROTATIONAL DYNAMICS

ficient of incremental pressure in Eq. ( 2 ) to be the manifold time constant, rP, results in the linearized expression for manifold dynamics contained in Fig. 2 .

Fuel Systems
Operation of an engine at or near a particular air-to-fuel ratio requires management of both air and fuel flow into the manifold system. Numerous methods exist for sensing airflow in order to provide a proportional fuel command, many of which are described by Bowler [ 181 and Manger [19]. These methods generally fall into the categories of volume rate measurement, mass rate measurement, or indirect estimation methods based on the measurement of related variables. This latter method of airflow determination is referred to as speed density. Stated simply, a

speed density air sensing system is based on the calculation of an estimate, defined here as A M , of the mass flow rate quantity &f. An example of an estimate in terms of pressure and engine speed is shown,
A M = cPN

(3)

where A M is the air mass flow rate (Ibm/ min), P the manifold absolute pressure (psia), N the engine speed (rpm), and c the proportionality constant (Ibm-in.*/lbf-rev). Linearizing Eq. (3) about the nominal engine speed, No, and pressure, P o , yields
~~

AAM = c(PoAN

+ NoAP)

(4)

For control to a specific air-to-fuel ratio, the engine fuel flow rate is proportional to the airflow rate. The actual amount of fuel in-

jected at any one event is proportional to the airflow rate divided by engine speed (which is assumed to be proportional to the air charge). The other aspect of fuel management that must be addressed for electronically controlled, port fuel-injected engines is injection timing. Two injection timing strategies will be discussed here. Sequential electronic fuel injection meters fuel individually to each cylinder during the appropriate portion of the engine cycle. For example, fuel might be injected into each cylinder in turn immediately before the intake valve opens. Thus, each cylinder receives a fuel charge delayed the same amount from the time of injection. Another method of injection timing is referred to as bank-to-bank injection. In one

Augu5t 1988

21

implementation of this strategy for a six-cylinder engine, the amount of fuel to be injected IS calculated once per engine revolution based on the speed density airflow. For purposes of economy and simplicity, the injectors are slaved in groups of three and fired alternately at 360-deg crank-angle increments delayed 240 deg from the fuel metering calculation.

Engine Pumping
An estimate for the air, fuel, and EGR mass flow rates out of the manifold and into the cylinders can be developed by treating the engine as a pump. Mass flow rate would thus equal the product of engine speed, engine displacement, and volumetric efficiency. For constant intake manifold temperature and exhaust gas pressure, the volumetric efficiency may be expressed as a high-order polynomial in speed and manifold pressure. Using such a relation in the product to form manifold mass Row rate egress, M , yields a polynomial function of speed and pressure. An example of a plot of mass flow rate as a function of manifold pressure and speed is shown in Fig. 3. These data are referred to as an induction map. The partial derivative of this function with respect to engine speed at a particular engine operating point defines the pumping feedback gain K,v in the linearized model.

1
I

*FWMPING LOOP- WORK (NEGATIVE)

TDC

CYLINDER VOLUME

I
I

Fig. 4.

Cylinder pressure versus volume diagram.

Induction-Po~3erStroke Delay
As in a reciprocating pump, individual samples of the mass flow rate for each cylinder are taken into the engine in a continuous speed-dependent sequence. These samples eventually produce torque via the combustion process after being delayed by

the combustion to power stroke lag included in Fig. 2 as a transport delay. Successive 180-deg increments of crankshaft rotation delineate the basic phenomena of the four-stroke cycle, spark ignition engine. These fundamental events are the ingestion of a combustible aidfuel mixture into the cylinder through the open intake valve as the piston traverses from top-deadcenter (TDC) to bottom-dead-center (BDC) of the intake stroke, compression of the mixture as the piston returns to TDC, ignition and rapid expansion during the subsequent power stroke driving the piston downward in the cylinder and imparting torque to the crankshaft, and, finally, elimination of the products of combustion from the cylinder

through the open exhaust port as the piston returns to T D C during the exhaust stroke. These events are illustrated in the cylinder pressure versus volume diagram of Fig. 4. It is clear that a delay exists between the ingestion of the aidfuel mixture and the torque production related to this mixture. That is, the torque developed by the engine at any particular time is a function of the flow and pressure characteristics extant during the previous induction event. Hence, the minimum induction-to-power (IP) stroke lag is 180 deg of crankshaft rotation. It is to be expected that this IP lag has significant control implications. In particular, at low engine speeds, where the IP lag is the longest, the delay in torque can have an adverse effect on engine stability.

Engine Brake and Combustion Torque

1200

72.0

a
\

=
?

1000
800

m
l -

LL

a
3 600

5
LL
ln ln

400

1 2

200
O

L
0

IO

12

1 4

MANIFOLD PRESSURE-PSI

Fig. 3. Example of an internal combustion engine induction map.

Torque is generated from the combustion process, which is dependent on the ignition of a cylinder charge of air, fuel, and residual gas, as well as other variables and parameters that influence combustion efficiency (such as the cylinder head geometry, for example). Defining the engine torque in terms of measurable or physically meaningful independent variables yields a quasistatic relation upon which dynamic elements reflecting friction effects and breathing delays may be superimposed. The structure of the torque equation provides a foundation for experimental determination of appropriate numerical values. An estimate for characterization of the engine torque is obtainable by employing analytical curve-fitting techniques to dynamometer-obtained experimental data as,

22

ItControl

Systrms Moqoiine

for example, in [20]. For the linear model, the following functional dependence is assumed for the engine brake or output torque:

1-1
INTAKE

.
Si\hIPLE INTERVAL = 120 DEGREES
CYL 1

(5)
where M,, is the mass charge delayed by the IP lag (Ibm). F,, the fuel delayed by the IP lag (Ibm), and 6 the ignition timing (degrees before TDC). As previously developed, mass charge is a function of manifold pressure and engine speed. Because pressure is the dominant variable. engine output torque can be considered to be an implicit function of delayed pressure, P,,. The linearized relationship is then
COMPRESSION POWER EXHAUST

INTAKE

COMPRESSION

POWER

Fig. 5. Six-cylinder engine sampling times.

A T,

G,, A P d

+ Gf A Fd + G6A6 + FNAN

moment of inertia, angular acceleration, and the difference between the net torque generated by the engine and the load torque of the shaft. Crankshaft acceleration is given by Newtons second law:

(6)

J e N = ( 3 0 / a ) T , - (30/a)TL

(7)

where G,>is the influence of delayed pressure on torque, Gf the influence of delayed fuel on torque, G6 the spark advance influence on torque, and FN the engine friction defined as the partial derivative of engine torque with respect to speed. The first three terms of Eq. (6) define what is usually referred to as the combustion torque, T,.. For a four-cylinder engine, the fundamental engine events are offset 180 crank-angle degrees per cylinder, such that each cylinder produces power over one-quarter of the two engine revolutions per cycle. Assuming that a uniform torque pulse is produced over the entire I80 deg of the power stroke based on the sample value taken at the end of the previous intake stroke, the combined output of the four cylinders would describe a nonintermpted pulse torque function extending over the entire cycle. For the six-cylinder engine, the basic cycle events are offset 120 deg per cylinder so that the power stroke of any particular cylinder begins 60 deg before the power stroke of the previous cylinder in the rotation ends. If the terminal 60 deg of the power stroke are neglected, a 120-deg sample-rate-and-zero-order hold, beginning at TDC of the power stroke, provides a continuous, nonoverlapping torque output for the combined cylinders with any individual torque component comprised of a 120-deg torque pulse, the magnitude of which is based on the value sampled 60 deg before BDC of the intake stroke. Note that the I P delay extends from 60 deg before BDC of the intake stroke until TDC of the power stroke, an interval of 240 deg or two sample periods at the 120-deg rate. These events are illustrated in Fig. 5.

where J , is the engine inertia (ft-lbf-sec/rad), T, the engine output torque (Ibf-ft), and TL the engine external torque load (Ibf-ft). For a vehicle employing an automatic transmission, the external torque load consists of the load applied by the torque converter plus external torque disturbances, which may arise as a result of auxiliary loads imposed on the engine (engagement of the air conditioner compressor, for example). The torque from the converter is generally specified as the square of the ratio of engine speed to a converter input capacity factor, K,. Linearizing Eq. (7) and substituting the expressions for T, and TL provides a differential equation describing the incremental engine acceleration, N :

(e)

(G,,AP

+ GfAFd
(8)

+ G6A6 - A Td)

If a constant KR is defined to incorporate the polar moment of inertia term and the inverse of the coefficient of A N is defined as the rotating inertia time constant 7 R , then the inclusion of a disturbance torque, A Td, results in the linear transfer-function relationship for engine acceleration illustrated in Fig. 2.

Model Parameter Estimation and Validation


Model parameter estimation and validation experiments might include static and dynamic tests compatible with allowable dynamometer or vehicle measurements. First, static experimental data may be used to calibrate the throttle body, estimate engine positive crankcase ventilation and other leakage,

and generate an engine pumping induction map. This information combined with the predetermined engine torque data may be used to develop an algebraic expression for brake torque as a function of AIF, mass flow rate, speed, and spark advance. Subsequent to static calibration, dynamic tests should be performed using a throttle kicker (at various throttle-angle levels) and spark advance steptype inputs. A representative engine system response to a small throttle-angle step input is shown in Fig. 6 for a four-cylinder, carburetted engine along with modeling results showing relatively good correlation. Adjustment in pressure-sensitive induction map parameters would reduce the steady-state pressure error, and modification of the simulated transmission damping would reduce the transient error exhibited in the illustration. A method for performing a number of simple tests to obtain model parameter values is delineated by Coats and Fruechte [21]. The experiments essentially consist of throttle, spark, and load inputs that give perturbation responses of engine outputs, the corresponding measurement of which allows direct estimates of model parameter values. The linearized structure of the engine model also forms the foundation for the use of identification techniques to determine model parameters. This approach was employed by Morris et al. [22] and Moms and Powell 1231. Generally, in these approaches, the engine model is grouped into dynamic effects associated with the intake manifold and the rotating inertia. Landaus identification technique [24] may then be applied to the multiple-inputisingle-output subsystems. A beneficial aspect of the identification approach is that the signal measurement and control implementation effects are incorporated into the model parameters.

Linear Multiple Sampling Rate Example


A block diagram of a linearized six-cylinder engine model and idle speed controller is illustrated in Fig. 7. This engine representation is the basic linear engine model of

Power-Train Rotational Dynamics


The rotational motion of the engine crankshaft is given in terms of the engine polar

Auyubt

1988

i J

MAN IFOLD
PRESSURE ("Hg)

"1A *
/Vahicle 15
I1

-L
'.I

-.------

\ Mode
I

U 0 1 2 0 1 2 TIME ( S E C )

Fig. 6. Sample transient response for validation.

Fig. 2 evolving at the fundamental six-cylinder period, T, equal to 120 crank-angle degrees. To this model has been added speed density airflow estimation as described by

Eq. (4) and the bank-to-bank fuel-injection timing previously described. Note that the fuel-injection timing period, f, of 360 deg produces a sampling rate in the fuel subsys-

tern that is one-third the fundamental engine rate. The samplers in Fig. 7 are included to emphasize this sampling rate discrepancy between the discrete-time subsystems. The multiple sampling rates provide a nonstandard but tractable stability analysis problem as described by Powell et al. [ 2 5 ] . Closedloop idle speed feedback control is effected by pure integral control of airflow via a closed throttle bypass valve to provide steady-state accuracy and proportional control of spark timing to enhance speed of response. Typical engine parameters for the six-cylinder engine are enumerated in the Table. Figure 8 illustrates the open- and closed-loop response of the system to a unit torque disturbance. The oscillatory nature of the open-loop response is due to the injection timing delays in combination with the IP lag and the manifold filling dynamics. Although the model illustrated is linearized about a particular idle speed, it should be emphasized that the model is actually accurate within a reasonably large neighborhood of operating points by virtue of the speed-dependent sampling. Such use of state(or feedback-)variable transformation to rep-

(7 =

3T)

FUEL GAIN

AIR BY PASS

MANIFOLO FILLING

KZ
1 GP

I
IP LAG

ROTATIONAL

SPARK GAIN

PUMPING FEEDBACK
KN

Gs

AN

l + T

Fig. 7.

Linearized six-cylinder engine idle speed control model.

24

Table Six-Cylinder Engine Model Parameters at 600 rpm


r,,. sec K,,, (Ibf-h)/(lbm-in.-sec) G,,. ft-lbfipsi T. sec rR. sec K,. rpni/( ft-lbf-sec) K,,+ Ibm/(rpm-min) G,. ft-lbfilbm

0.21 0.776 13.37 0.033 3.98 67.2 0.08 36.6

5 0

RPM

.: 1
-2J
0 5

1
1
1 5 2 2 5

I 5

4 5

resent a nonlinear system by an equivalent linear system is referred to by Kokotovic [26] and illustrated by Cook and Powell [27].

TIME (SECONDS)

Fig. 8. Open-loop (upper) and closed-loop (lower) time re sponse of six-cylinder engine model to unit torque disturbance [7] J . F. Cassidy, M. Athans, and W-H. Lee, On the Design of Electronic Automotive Engine Controls Using Linear Quadratic Control Theory, IEEE Truns. Auto. Conrr., vol. AC-25, no. 5, Oct. 1980. 181 E. Kamei, H . Namba, K. Osaka, and M. Ohba, Application of Reduced Order Model to Automotive Engine Control System, ASME J . Dyn. Syst., Meas., Contr., vol. 109. pp. 232-237, Sept. 1987. 191 B. K. Powell. A Dynamic Model for Automotive Engine Control Analysis, Proc.
18th IEEE Conf Decision and Control. pp.

Conclusions
Thc development of a basic nonlinear representation of an engine dynamic system has been reviewed. The model contains descriptions for the induction process and engine power system a s well as characterization of the fuel system. The general description forms a foundation to which other important transient characterizations, such as exhaust gas recirculation system dynamics o r intake manifold fuel wall-wetting may be added. In addition. a linear model has been developed for a particular six-cylinder engine and a time response of the system is presented.

[I81

[ 191

[20]

References
P. A . Hazell and J. 0. Flower, Sampled

Data Theory Applied to the Modeling and Control Analysis of Compression Ignition Engines-Part I , fnt. J . Conrr., vol. 13. no. 3. pp. 549-562, 1971. P. A. Hazell and J . 0. Flower. Sampled Data Theory Applied to the Modeling and Control Analysis of Compression Ignition Engines-Part 11, Int. J . Contr., vol. 13, no. 4. pp. 609-623, 1971. P. A . Hazell and J . 0. Flower, Discrete Modeling of Spark Ignition Engines for Control Purposes, fnt. J . Contr., vol. 13, no. 4. 1971. F. J . Garofalo, Performance Capabilities of Various Air Measurement Techniques for AiriFuel Ratio Feedforward Control Strategies on Electronic Fuel Metered Engines, University of Michigan. June 30,
1975. R . Prabhakar, S. J. Citron, and R. E. Goodson. Optimization of Automobile Engine

Fuel Economy and Emissions. ASME Paper 75-WAlAut-19, Dec. 1975. B. K . Powell. A Simulation Model of an Internal Combustion Engine-Dynamometer System. Proc. Summer Computer SimuItrrion Cor$, Newport Beach, CA, July 24, 1978.

120-126, 1979. [IO] R. G. Delosh, K. J. Brewer, L. H. Buch, T. F. W. Ferguson, and W. E. Tobler, Dynamic Computer Simulation of a Vehicle with Electronic Engine Control, SAE Paper 810447, 1981. [ 1 I] D. 1. Dobner, A Mathematical Model for Development of Dynamic Engine Control, SAE Paper 800054, 1980. 112) H. Wu, C. F. Aquino, andG. L. Chou, A 1.6 Litre Engine and Intake Manifold Dynamic Model, ASME Paper 83-WA/DSL39, 1984. [I31 B. K. Powell and J. A. Cook, Nonlinear Low Frequency Phenomenological Engine Modeling and Analysis, Proc. 1987 Amer. Contr. Con$, vol. I, pp. 332-340, June 1987. [I41 J. D. Powell, A Review of IC Engine Models for Control System Design, International Federation of Automatic Control, Munich, Germany, July 28, 1987. [15] W. W. Yuen, A Mathematical Engine Model Including the Effect of Engine Emissions, Department of Mechanical and Environmental Engineering, University of California, Santa Barbara, CA, Feb. 26, 1982. [I61 H. B. Servati, Investigation of the Behavior of Fuel in the Intake Manifold and Its Relation to S.1. Engines, Ph.D. thesis, University of California, Santa Barbara, CA, Mar. 1984. [I71 J . J . Moskwa and J. K. Hedrick, Auto-

1211

[22]

1231

1241

[25]

[26]

[27)

motive Engine Modeling for Real Time Control Application, Proc. 1987 Anier. Contr. Cor$, vol. 1 , pp. 341-346. June 1987. L. L. Bowler, Electronic Fuel Management Fundamentals, SAE Paper 800539, 1980. H. Manger, Electronic Fuel Injection, SAE Paper 820903, 1982. Z. Mencik and P. N. Blumberg. Representation of Engine Data by Multi-Variate Least-Squares Regression, SAE Paper 780288, 1978. F. E. Coats and R. D. Fmchete, Dynamic Engine Models for Control DevelopmentPart 11, Application of Control Theon in the Automotive Industrs, Geneva. Switzerland: Interscience, 1983. R. L. Moms, R. H. Borcherts, and M. V. Warlick. Parameter Estimation of SparkIgnited Internal Combustion Engines. Sixth IFAC Symp. Identification and System Parameter Estimation, Washington, DC, June 1982. R. L. Moms and B. K. Powell, Modem Control Applications in Idle Speed Control, Proc. 1983 Amer. Contr. Con$, vol. 2, pp. 79-85, June 1983. I. D. Landau, Unbiased Recursive Identification Using Model Reference Adaptive Techniques, IEEE Truns. Auto. Contr., vol. AC-32, no. 2, pp. 194-202, Apr. 1986. B. K. Powell, J. A. Cook, and J . W. Grizzle. Modeling and Analysis of an Inherently Multi-Rate Sampling Fuel Injected Engine Idle Speed Control Loop, ASME J . Dyn. S ) x , Meas., Contr., vol. 109, pp. 405-410, Dec. 1987. P. V. Kokotovic, Control Theory in the 80s: Trends in Feedback Design, Ninth World Congress of IFAC, Budapest, Hungary, July 1984. J. A. Cook and B. K . Powell. Discrete Simplified External Linearization and Analytical Comparison of IC Engine Families, Proc. 1987Amer. Conrr. Con$, vol. 1, pp. 326-330, June 1987.

25

Jeffrey A. Cook has been a Research Engineer in the Control Systems Department, Research Staff, Ford Motor Company, since 1976. He received the B.S. degree in mechanical engineering from The Ohio State University in 1973 and the masters degree in electronics and computer control from Wayne State University in 1985. His research interests are in the areas of air/ fuel ratio and emissions control for internal combustion engines. He is also an Adjunct Faculty Member at Lawrence Institute of Technology, Southfield, Michigan.

Barry K. Powell has been a Research Engineer in the Control Systems Department, Research Staff, Ford Motor Company, since 1976. Prior to that time, he worked at Ford Automotive Safety Research and Bendix Corporation Research Laboratories. His responsibilities have included mathematical modeling and control of aerospace and automotive systems. His current activity is in real-time analysis and control of automotive power-train systems.

Out of Control

This years model comes with a servo amplifier for speed control.

26

I Control Systems Mogozlne

Vous aimerez peut-être aussi