Sie sind auf Seite 1von 57

ME 7236 Module 1 Autumn 2012

1.1



ME 781
Powertrain Dynamics

Giorgio Rizzoni
Krishnaswamy (Cheena) Srinivasan
The Ohio State University
Department of Mechanical Engineering



Autumn Semester 2012




The Ohio State University
Center for Automotive Research
ME 7236 Module 1 Autumn 2012
1.2




Part I
Engine Dynamics


Giorgio Rizzoni
Jonathan Dawson (The Mathworks)
Yong-Wha Kim (Ford Motor Company)
Byungho Lee (General Motors)
Qi Ma (General Motors)
Amr Radwan (Detroit Diesel Corporation)
Devesh Upadhyay (Ford Motor Company)
Inkwang Yoo (Ford Motor Company)
Marcello Canova (The Ohio State University)
Fabio Chiara (The Ohio State University)


ME 7236 Module 1 Autumn 2012
1.3
Module 1
Introduction to Powertrain Dynamics

1.1. COURSE ORGANIZATION ..................................................................... 5
Objective ......................................................................................................... 5
Assignments and examinations ....................................................................... 5
What do we mean by System Dynamics? ................................................... 6
What this course is not .................................................................................... 6
Structure of the course .................................................................................... 6
Required background ...................................................................................... 7
1.2. BACKGROUND AND MOTIVATION ................................................ 8
Fuel efficiency improvement .......................................................................... 8
Emission standards ......................................................................................... 9
Emission controls for spark-ignition engines ............................................... 12
The importance of engine control systems ................................................... 14
1.3. INTRODUCTION TO POWERTRAIN DYNAMICS ............................. 16
1.4. MODELING FOR POWERTRAIN CONTROL ..................................... 20
Fundamental equations for modeling fluid systems dynamics ..................... 25
Model coupling techniques ........................................................................... 29
Modeling guidelines ..................................................................................... 31
1.5. TWO IMPORTANT ENGINE CONTROL PROBLEMS ................... 32
The AFR Control Problem ............................................................................ 32
Open-loop AFR Control ............................................................................... 32
Closed-Loop AFR Control ............................................................................ 37
ME 7236 Module 1 Autumn 2012
1.4
The Idle Speed Control Problem .................................................................. 39
Idle Air Control ............................................................................................. 41
Ignition Timing Control ................................................................................ 47
Summary ....................................................................................................... 50
1.6. REFERENCES ......................................................................................... 51
1.7. EXAMPLES ............................................................................................ 52
Example 1.1: Water tank filling dynamics model ........................................ 52
Example 1.2: Compressible flow through an isentropic nozzle ................... 54
Example 1.3: Torsional system dynamics .................................................... 55

ME 7236 Module 1 Autumn 2012
1.5
1.1. COURSE ORGANIZATION
Objective
The objective of this course the first in a sequence of two - is to introduce
practicing engineers and engineering students (the latter typically in the first or second
year of a M.S. program in electrical or mechanical engineering) to the essential aspects of
modeling and control of automotive powertrains. The course recognizes the significance
of this growing area of engineering and its central role in the automotive industry. The
lectures place emphasis on the integration of many different aspects of powertrain
engineering, including: the dynamics of mechanical, fluid, and thermodynamic systems;
sensor and actuator technology; and feedback controls. Primary emphasis will be given to
dynamics and control of fuel-injected, spark-ignited, internal combustion engines, while
the integration with the complete powertrain (torque converter and transmission) will
also be addressed. The course will present an overview of the major dynamic phenomena
that characterize powertrain behavior: intake and exhaust air flow dynamics; fuel system
dynamics; combustion and emissions; crankshaft dynamics; and air-fuel ratio control.
Emphasis will be placed on explaining the interaction between subsystems, and the
importance of considering the entire vehicle system when assessing the impact of the
performance of a subsystem on overall system performance. Modeling and computer
simulation will be integral part of the course, showing how to build numerical models of
engine/powertrain systems and components and apply them to solve problems pertaining
to powertrain dynamics.

Assignments and examinations
A homework assignment will be handed out each week. Each homework
assignment will include the following components:
a) theoretical: understanding the nature of the problem and identify the path
for solution;
ME 7236 Module 1 Autumn 2012
1.6
b) analytical: development of a mathematical formulation of the problem,
based on a set of equations and a number of physical constants;
c) computational: develop numerical procedures for solving the equations,
formulate criteria for verification and interpretation of the results.
In addition to the homework assignments, the students will complete a modeling,
simulation and system identification project based on data from an actual powertrain.

What do we mean by System Dynamics?
System dynamics is a discipline that studies the mathematical representation of
systems, focusing on the dynamic behavior. Typically, the system-dynamic
representation of a complex system such as an automotive powertrain consists of a set of
coupled, nonlinear, ordinary differential equations. Linear and nonlinear system analysis
and computer simulation methods are used to analyze the properties and characteristics of
the describing equations. System dynamics is a discipline that is often presented in the
junior or senior year of the undergraduate mechanical engineering curriculum, focusing
on the dynamics of electrical, mechanical, electro-mechanical, hydraulic, pneumatic, and
thermal systems [1]. In this course it is expected that students have a good working
knowledge of modeling and analysis of these families of systems, and of linear analysis
methods. In addition, the use of computer-aided tools such as Matlab/Simulink is also
necessary to complete many of the assignments in this course [2]. A summary of systems
dynamics concepts is available for those students who are in need of a review [3].
What this course is not
a) Modeling of IC engine thermo-fluid processes.
b) Modeling of mechanical dynamics of engines.
c) Design of transmission elements or subsystems.
Structure of the course
Module 1: Introduction to powertrain dynamics
Module 2: Intake and exhaust dynamics
ME 7236 Module 1 Autumn 2012
1.7
Module 3: Fuel injection and fueling dynamics
Module 4: Combustion, Ignition control and knock
Module 5: Crankshaft dynamics
Module 6: Overview of engine control systems
Module 7: Fluid couplings and torque converters
Module 8: Transmission mechanical system and vehicle longitudinal dynamics
Module 9: Transmission shift hydraulic system
Module 10: Open loop transmission control
Module 8: Shift schedules and continuously variable transmissions

Each of the above modules is presented according to the following sequence:
- Present physical phenomenon in intuitive terms
- Show physical components, where possible, and illustrate performance curves
- Define physical laws
- Derive equations of motion
- Interpret equations of motion
- Discuss computer simulation
- Discuss model identification experiments (where appropriate)

Required background
o Undergraduate level linear system theory
o System Dynamics (dynamics of electrical, electro-mechanical, fluid, thermal, and
mechanical systems) - review available in the technical presentation Signal and
System Dynamics Integration, by Profs. Rizzoni, Srinivasan and Yurkovich.
o Basic IC engine processes - review available in the technical presentation
Internal Combustion Engine Fundamentals, by Prof. Guezennec.
o Familiarity with the Matlab/Simulink environment - review available in the
GMTEP technical presentation Matlab/Simulink - Introduction.
ME 7236 Module 1 Autumn 2012
1.8
1.2. BACKGROUND AND MOTIVATION

Emissions, fuel efficiency and safety requirements have steadily become more
stringent over the last few decades. These regulations can be considered the driving force
behind most of the advancements in the technology of internal combustion engines,
which have been improving at increasing rate with multi-disciplinary effort at all stages
of development.

Fuel efficiency improvement
The recent surge in oil prices has dramatically increased the awareness on
improving the fuel efficiency of passenger vehicles. The problem becomes increasingly
alarming when the entire chain of oil supply, refinement, transportation and utilization is
considered (well-to-wheel analysis).

Figure 1: U.S. Energy Flow Trends in 2005 (Units in quadrillion BTUs)
ME 7236 Module 1 Autumn 2012
1.9
Figure 1 summarizes the energy utilization in the U.S. based on data collected in
2005. Focusing on the bottom part of the chart, it is possible to notice that the petroleum
supplies, which rely mostly on foreign sources (62%), are used almost exclusively for
transportation (out of which, 43% is used for fueling light-duty vehicles, such as
passenger cars and light duty trucks). Furthermore, the energy used for transportation
shows the worst overall efficiency, with values around 20%. This scenario suggests the
strong priority of the automotive industry to design fuel efficient vehicles, as well as the
efforts in improving the overall well-to-wheel efficiency and diversifying the fuel sources
for transportation (i.e., biofuels and natural gas).

Emission standards
Another important motivation for the development of modern and efficient
engines and powertrain systems is dictated by the emission legislations, which have
become increasingly stringent in the past years in the major world automotive markets
(U.S. and Europe), as well as in the emerging economies, such as Asia and Latin
America. These regulations can be considered the driving force behind most of the
advancements in the technology of internal combustion engines.
In the United States, emission standards are managed by the Environmental
Protection Agency (EPA) (www.epa.gov). Few state governments, however, implement
own regulations. This is the case of California, where the California Air Resource Board
(CARB) has applied some of the strictest standards in the world.
Within the Clean Air Act Amendments (CAAA) of 1990, the EPA defined two
sets of federal standards for light-duty vehicles (similar rules have been defined for
heavy-duty vehicles and trucks):
- Tier 1 Standards, which were published on 1991 and were effective from
1994 until 2003;
- Tier 2 Standards, which were initially adopted in 1999 with an
implementation schedule from 2004 to 2009.
ME 7236 Module 1 Autumn 2012
1.10
Currently, vehicles sold in the United States must meet Tier 2 standards, which
are characterized by lower emission limits and a number of additional changes that made
the standards more stringent for larger vehicles. Although the former Tier 1 standards
were different between automobiles and light trucks, the current Tier 2 regulation apply
the same emission standards to all vehicle weight categories (i.e., cars, minivans, light-
duty trucks, SUV), regardless of the fuel they use. With these regulations, large engines
(such as those used in light trucks or SUV) are forced to use more advanced emission
control technologies than smaller engines in order to meet the standards.
Within the Tier 2 standards, there is a ranking of 8 different emission levels,
named Certification Bins, ranging from BIN 1 (corresponding to zero emissions) to BIN
8. Vehicle manufacturers are allowed to certify their vehicles to any of the 8 categories.
At the same time, the average NOx emissions of the entire vehicle fleet sold by each
manufacturer has to meet the average NOx standard of 0.07 g/mi. Temporary emission
standard, less restrictive, have been set as a transitional step until the full implementation
or Tier 2 Standards in 2007.
The EPA Bins cover California LEV 2 emission categories, to uniform vehicles
certification to the Federal and California standards. The emission standards for all
pollutants, are shown in the following figure.

Figure 2.1: Tier 2 Federal Emission Standards (units in [g/mile]) (source: http://www.dieselnet.com)

ME 7236 Module 1 Autumn 2012
1.11
California emission standards have been traditionally more stringent than the EPA
requirements, but their evolution and structure is similar to that of the federal legislation.
The California Air Resource Board (CARB) adopted a former emission standard
program, applied until 2003, where six vehicle types are defined:
- Tier 1;
- Transitional Low Emission Vehicle (TLEV);
- Low Emission vehicle (LEV);
- Ultra Low Emission Vehicle (ULEV);
- Super Ultra Low Emission Vehicle (SULEV);
- Zero Emission Vehicle (ZEV).
Each category was characterized by more stringent emission restrictions. Tier 1
was the baseline used to determine the standards. Car manufacturers were required to
produce a percentage of vehicles certified to increasingly more stringent emission
categories.
In 1998 CARB defined the LEV 2 emission standards, adopted from 2004 to
2010. Under the LEV 2 standard, NOx and PM standards for all emission categories are
significantly tightened, and the same emission levels apply to both gasoline and Diesel
vehicles. Specific emission standards are defined for passenger cars (including light-duty
trucks and medium-duty vehicles below 8500 lbs gross weight) and for heavy-duty
vehicles. As a result, most pick-up trucks and sport utility vehicles are required to meet
the passenger car emission standards The TLEV emission category has been eliminated.
It is, therefore, believed that the LEV 2 emission standards can only be met by
vehicles fitted with advanced emission control technologies, such as particulate filters
and NOx catalysts, or by hybrid-electric vehicles. The following figure reports an excerpt
of the LEV 2 emission standards for passenger cars.
ME 7236 Module 1 Autumn 2012
1.12

Figure 1.2: California LEV 2 Emission Standards for passenger cars (units in [g/mile])
(source: http://www.dieselnet.com)

Emission controls for spark-ignition engines
To comply with the emission regulations, spark ignition (S.I.) engines require the
presence of a three-way catalytic converter to reduce emission levels of the three major
pollutants (HC, NO
X
, and CO). The catalytic converter operates most efficiently in a
narrow window around the stoichiometric value of the air-fuel ratio. Closed-loop
regulation of the air-fuel ratio is therefore required to achieve this goal; the feedback
signal used to close the loop in the air-fuel ratio controller is provided by an exhaust gas
oxygen, or lambda, sensor. The oxygen sensor contains a ceramic material that has an
electrical response to changes in the oxygen partial pressure in the exhaust stream
relative to ambient; this sensor becomes active when it reaches a temperature T > 250

C.
The concentration of the amount of oxygen relative to ambient contained in the exhaust is
related to the air/fuel ratio. The voltage output of the lambda sensor is processed by the
Engine Control Unit (ECU), a microcontroller, which outputs a signal to the fuel
injectors (pulse width) adjusting the amount of fuel injected according to the duration of
the input voltage pulse. In this way the air-fuel ratio is maintained as close as possible to
the stoichiometric value to take advantage of the higher conversion efficiency of the
catalytic converter.
To further reduce the formation of NO
X
, it is possible to recirculate a fraction of
the exhaust gas into the intake manifold through a controlled valve. This practice,
conventionally known as Exhaust Gas Recirculation (EGR), is nowadays in decline for
S.I. engines, but is largely adopted in Diesel engines, where it can reach up to 50 percent
of the total trapped charge. The use of EGR corresponds to adding an inert gas to the
ME 7236 Module 1 Autumn 2012
1.13
mixture and thereby reducing the combustion temperature, leading to the reduction of
NO
X
formation. In a typical automotive S.I. engine, 5 to 15 percent of the exhaust gas is
routed back to the intake as EGR.
The control of EGR is typically accomplished using an open-loop structure. The
engine ECU contains maps specifying the opening position of the valve across the engine
operating range. In modern EGR systems, the valve is electrically operated (through a
solenoid actuator), providing a feedback signal on the valve position. Although EGR
does measurably slow combustion, this can be compensated for by advancing the spark
timing.
Open-loop control using other available sensors (throttle position, manifold
pressure and temperature, mass air flow, etc.) is typically used at operating conditions
which require richer or leaner mixtures than in the case of closed-loop control; cold start
and heavy acceleration are two conditions among others which require the open-loop
control strategy to take over.
Additional actuation mechanisms may also be employed to aid in the control of
the engine exhaust emissions; these may include: heated catalysts; electronic throttle
control; auxiliary air handling during starting; idle air bypass actuator; and other systems.
Further, HC emissions which originate from evaporative sources (i.e. evaporation
of the fuel in the tank) are reduced by the use of an evaporative emission control system
which is designed to store and to dispose of the fuel tank vapors. This evaporative system
consists of an active-charcoal canister that stores the HC vapors; these vapors are then
purged into the intake manifold to be burned with the mixture at operating conditions that
require additional mixture enrichment. This system has a strong interaction with the fuel
control system because during the purge cycle air and fuel vapors are introduced into the
intake system in unknown quantities, disturbing the regulation of the lambda control.
The combination of the above described control strategies constitutes what is
generally known as an automotive engine emissions control system. The implementation
of such a control system with the performance required to satisfy emission regulations
requires the use of event-based control (i.e. the control inputs are not computed on a
ME 7236 Module 1 Autumn 2012
1.14
fixed time increment but are dependent on the position of the engine crankshaft), so as to
manage the individual fuel injection and spark ignition events, which occur at different
rates as the engine speed varies. Further, the complete control strategy requires the
measurement of all the signals necessary to obtain information of the engine operating
condition and the parallel implementation of several strategies, including: i) spark timing;
ii) fuel injection (quantity and timing); iii) exhaust gas recirculation (EGR); iv) canister
purge cycle; v) on-board diagnostics.

The importance of engine control systems
The directives for fuel efficiency and environmental protection are followed both
by the Industry and Research Centers focusing mostly on two key areas: design
optimization and control.
Figure 1.3 compares the main tasks related to design optimization and control for
mid class and luxury segment vehicles. In response to both environmental and
acceptability challenges, significant improvements are required to fulfill the targets
imposed in each vehicle segment. For mid class segment vehicles, a large effort is
dedicated to fuel economy improvement through engine downsizing, i.e. reducing the
displacement. This trend is also related to the upcoming regulation regarding CO
2
emissions. Moreover, in order to conquer a market share, the consumer acceptability
requirements (acoustic refinement and driving performance) must be improved.

Figure 1.3: Future engine development trends for mid class and luxury segment vehicles
ME 7236 Module 1 Autumn 2012
1.15
Slightly different effort must be dedicated to the development of luxury segment
vehicles. Here, the consumer acceptability implies a higher priority for driving
performance, combustion and structural refinements.
The tasks delineated must be met in very short time due to increased
competitiveness of the market. The most direct and effective way of improving engines
performance relies on design, either with the development of innovative technical
solutions or the refinement of existing products. The research effort in this field is
generally oriented in two directions. First of all, enhancements are sought for all the
various phenomena that affect the energy conversion process, such as combustion, gas
exchange processes, thermal management, etc Then it is also required to accommodate
physical constraints that are a result of the improvements sought, for example higher
peak pressures or temperatures.
On the other hand, control design improvement seeks to optimize the various
processes in order to deliver power most efficiently while meeting constraints on
emissions, safety and reliability. Generally, new design and control concepts are tested
extensively prior to prototype applications, a task that constitutes a labor intensive and
expensive process.
In this sense, one of the priorities of industry is to shorten the development time
as much as possible, therefore striving to reduce the time-consuming and expensive tests
and calibration phases typically associated to the engine control development. This can
be accomplished by relying more on mathematical modeling tools to assist both the
engine design phase and the development of its control systems. Such modeling
capabilities are the key to reduce time and costs related to traditional experimental tests
and calibration efforts.
Even though not able to replace experimental investigations in full, simulation
models are capable of shortening consistently the development time from the definition
of the control system tasks to final tests and prototyping.
For internal combustion engines, a wide range of models is available to assist all
the design phases, from the simulation of flow and combustion processes, to the
ME 7236 Module 1 Autumn 2012
1.16
implementation and test of the control system. For example, Computational Fluid
Dynamics (CFD) and Finite Elements Method (FEM) are largely adopted in the
mechanical and fluid dynamic design of engines. Even though computationally intensive,
these tools are extremely powerful in assisting engineers and designers in the
development of individual components, such as manifolds, runners, ports and cylinder
parts. Their use can also replace part of the experimental analysis, which would normally
require complex testing environments and equipment.
However, when the final objective of the study is the design of a controller, it is
essential to have a global understanding of the behavior of the entire engine system,
rather than focusing on the details of a single component. In addition, computation time
becomes a significant constrain.
A typical engine system model, conceived for engine and powertrain control, will
be designed to predict only certain features, namely the dynamic response to
environmental and control inputs, avoiding an overly complex description of phenomena
that are not relevant for the analysis.

1.3. INTRODUCTION TO POWERTRAIN DYNAMICS
Modern engines and powertrain systems have become complex units, their
behavior influenced by the interaction of several components. Consequently, the number
of variables that must be considered for control and diagnostics has dramatically grown.
As stated above, this forces the control engineers to rely more on models from the
early development phases of the control systems. Even though models are not able to
replace experimental investigations, they contribute to shorten development time from
the definition of design specifications to final road tests.
The number and the complexity of design requirements are now forcing a shift
from map-based control systems towards model-based algorithms able to support control
and diagnostics functions.
ME 7236 Module 1 Autumn 2012
1.17
The approach commonly adopted for the analysis of steady-state and transient
behavior of engine and powertrain systems has its foundations in the system dynamics
theory, a common practice in the study of mechanical, electrical and hydraulic systems.
The general idea is to divide the system studied in elementary components which are
interconnected. The connections, namely the inputs and outputs of each element, become
then the focus of the analysis.
The following block diagrams are intended to illustrate in increasing order of
complexity the relationships among components and subsystems in a typical automotive
powertrain.
Figure 1.4 depicts the simplest overview, in which the engine is interpreted as a
subsystem that converts air and fuel flows to torque and exhaust gases; the developed
torque is what provides the tractive force for the vehicle through a transmission and
driveline.

Figure 1.4: Overview of basic engine function

A more detailed view of basic engine functions, illustrating the difference
between ideal and actual external inputs, is shown in Figure 1.5. In the figure, it is
shown that the flows of air and fuel are actually regulated by valves or flow restriction
devices (throttle body and fuel injectors), and are affected by internal dynamics. In
particular, two specific subsystems are source of dynamic behavior, namely the intake
manifold for air and the evaporation dynamics for the injected fuel.
ME 7236 Module 1 Autumn 2012
1.18
Throttle
Body
Intake
Manifold
Combustion
Fuel Injector
Crankshaft
EGO
Sensor
E
Fuel Dynamics
Tind
Tload
A/Fexh
Vo2
Spark
Controller
ou
air in
fuel in
air in cyl.
fuel in cyl.
engine speed
Transmission
Vehicle

Figure 1.5: More detailed view of powertrain dynamics
Intake
Manifold
Fuel Dynamics
(Well Wetting)
Combustion
Exhaust
Manifold
Fuel Injection
EGR
Canister
HEGO
Catalytic
Converter
E
t
fuel
o
can
o
EGR
Throttle Body
E
Inertia
Damping/Friction
Drive Train
Vehicle
Load
ou
sp
Exhaust
engine
speed
engine
torque
Air
Fuel
Electronic
Throttle
Control
from ECM
from ECM
Canister
purge
from ECM
from ECM
from ECM
Open-loop
fuel control
Closed-loop
fuel control
EGR
Control
Ignition
Control

Figure 1.5: Complete block diagram of automotive powertrain.
ME 7236 Module 1 Autumn 2012
1.19
Figure 1.5 shows a third block diagram which depicts a more complete system
model, including closed-loop air-fuel ratio control, and including disturbances such as
EGR and canister purge flows. This representation allows one to better understand the
interactions between components, defining the inputs, outputs of each subsystem and the
related controlled variables.
The characterization of the engine and powertrain dynamics involves studying the
dynamic response of each subsystem, as well as the interactions between the different
components. Further complexity is added by the characteristic frequency associated to
the dynamic response of each component, as summarized in Table 1.1.

Table 1.1 Powertrain dynamics time constants.
Subsystem Bandwidth or Time Constant
1. Intake Manifold Fast Dynamics: 200 -400 Hz
Slow Dynamics: 1 - 2 Hz
2. Fuel Injector Time constant: 0.5 - 3 ms
3. Fuel Dynamics Evaporation: 0.5 s
Mixing: 1 - 10 ms
4. Combustion Delay: 1/2 engine cycle
5. Crankshaft 0 - 2000 Hz
6. Transmission/Vehicle 3 - 5 Hz
7. O
2
Sensor Time constant: of the order of a few ms
(speeds up with aging)

The engine system, as an assembly of mechanical, hydraulic, thermal and
electrical components, is characterized by a number of different phenomena that typically
occur at different time scales. For example, the typical response of a fuel injector
(electro-hydraulic actuator) is rather fast, in the order of 0.1 3 ms. Conversely, the fuel
evaporation dynamics, being mostly dependent on the slowly-varying intake manifold
temperature and the fuel vapor partial pressure in the port, is typically characterized by a
larger time constant.
ME 7236 Module 1 Autumn 2012
1.20
The separation of time constants is a typical problem that occurs when modeling
the dynamics of engines and powertrain systems, which may resolve in numerical issues
(i.e., stiffness) and complications when the model is used for control design.
It is important then to realize that no system can be modeled exactly in all its
detail, but that a good control-oriented model will capture with reasonable accuracy the
behavior of the system in a bandwidth sufficiently limited to allow for a simple but
reliable control design.
Powertrain models aimed at the development of control strategies have been
developed since 1980, following the advent of digital simulations of internal combustion
engine processes. To this extent, it is worth mentioning the early works of Dobner [1.7],
Powell and Cook [1.8], Moskwa and Hedrick [1.9], Hendricks and Sorenson [1.10], Turin
and Geering [1.11]. In recent years, various control approaches have been proposed to
use such models for control purposes, and model based control methods are gaining
wider acceptance in the research community. Relevant works in the field have been
presented by Amstutz et al. [1.12], Ault et al. [1.13], Azzoni et al. [1.14], Chang et al.
[1.15], Cho and Hedrick [1.16], Grizzle and Cook [1.17-1.19], Hendricks et al. [1.20],
Powell et al. [1.21], Kao and Moskwa [1.22, 1.23], Turin and Geering [1.11].

1.4. MODELING FOR POWERTRAIN CONTROL
The application of the system dynamics principles to the study of powertrain
systems allows one to operate a deconstruction of the plant into a series of interconnected
components.
Once each component is determined in its relevant inputs/states/outputs, it is
necessary to develop a mathematical model that allows to predict the states and outputs in
relation with the inputs and (if any) the control variables. In this process it is important to
maintain a system viewpoint, hence not focusing on the details of the individual
component, but considering it as part of a more complex entity. This approach, even with
the loss of details for each individual component, is necessary to develop a description of
ME 7236 Module 1 Autumn 2012
1.21
the system capable of dynamically accounting for the most relevant phenomena, while
maintaining a reasonable computational ease.
In general, a preliminary stage for the analysis of a system or component consists
of operating a classification, namely making assumptions on the categories of models that
would best address the problem. Although there are neither standard nor accepted
methods of classifying system models, it is possible to define several categories in
relation with the characteristics of the system and the scopes of the analysis.
The operation of an internal combustion engine can be seen as the result of
interactions between several phenomena that occur at different temporal and spatial
scales. When modeling a specific engine component, a fundamental choice has to be
made to determine the bandwidth of the model, i.e., the maximum temporal and spatial
resolution. This choice leads to implicitly operate a spatial and temporal average
(~lowpass filter), removing spatial and temporal scales from the model, which will be
unresolved. Therefore, the capabilities of a model to capture specific engine phenomena
will be affected.
Following the principles above, two possible classifications of engine models can
be made:
Classification by Space Scale (Characteristic Length):
Micro-length (multi-D) the boundaries of the systems are very small,
allowing for detailed characterization of scalar and vector fields;
Small length (1D) the boundaries are set to characterize the field in typically
one direction; properties are assumed constant on any plane orthogonal to the
chosen direction;
Large length (0D lumped, high-order) the boundaries are assumed equal to a
component (e.g., valve, receiver,), wherein the properties are considered
uniformly distributed;
Very large length (0D lumped, low-order) several components are included
within the boundaries.

ME 7236 Module 1 Autumn 2012
1.22
Classification by Time Scale:
Very short time scales - captures evolution of internal system states at micro-
scale (e.g.: sub-millisecond).
Short time scales - captures dynamics much faster than the excitation time
scale (e.g.: tens of milliseconds);
Medium time scales - captures input-output dynamic behavior of comparable
time scale relative to excitation (e.g.: fractions of one second);
Long time scales (quasi steady) - system reaches equilibrium very quickly
relative to the time scale of excitation (e.g.: seconds, minutes, hours).
This classification concept can be better explained through an example. A very
important engine subsystem is the intake manifold. Inside this component, several
thermodynamic, fluid dynamic and heat transfer processes occur, and the breathing
performance of the engine (i.e., the ability to draw fresh air/fuel charge at each cycle)
results from complex interactions.

Figure 1.6: Example of model classification applied to intake manifold modeling
ME 7236 Module 1 Autumn 2012
1.23
Several models can be used to study the flow through the intake manifold. Figure
1.6 illustrates the classification based on the spatial and temporal scales reporting the
most relevant model that are generally adopted for studying the system at different levels
of accuracy and computation time:
- 3D Computational Fluid Dynamics (CFD) Models: the system is studied
in four dimensions (space and time), adopting a very high resolution that
allows to characterize in great detail the velocity field. Such models are
typically used for design optimization;
- 1D Wave-Action Models (WAM): these models remove two spatial
dimension from the problem, assuming that all the properties (pressure,
temperature, etc) vary only with respect to time and one spatial
coordinate (length). The time resolution and discretization length are
typically small, allowing for the characterization of the high-frequency
pressure fluctuations (wave propagation dynamics) that are very relevant
for engine tuning. Wave-action models are also used for optimizing the
manifold design, with respect to tuning and volumetric efficiency. From a
numerical standpoint, WAM are typically based on nonlinear partial
differential equations, which are solved using numerical approximations;
- 0D Filling-and-Emptying (F&E) Models: these models approximate all
the spatial resolution of the system, assuming that the thermodynamic
properties are uniformly distributed. The time resolution is typically set to
provide an adequate characterization of the system properties during one
engine cycle (typically, one degree of crank angle). For these reasons,
such models are often named thermodynamic, crank-angle based
models. Models pertaining to this category are considered a fair
compromise between accuracy (most of the high-frequency variations in
pressures and flow rates due to the alternative motion of the piston are
captured) and computation time. These models typically result in a high-
order system of nonlinear ordinary differential equations;
- 0D , Low-Order Models: these models capture only the low-frequency
dynamics of the system, considering all the system properties as cycle-
ME 7236 Module 1 Autumn 2012
1.24
averaged. This approach is often known as Mean-Value Modeling
(MVM). The result of such approximations is a low-order system of
nonlinear ordinary differential equations, as typically occurs in system
dynamics theory. This facilitates the applications of these models in
control system design.
Figure 1.6 illustrates also that the choice of a specific model category implies that
all the phenomena occurring at higher spatial and temporal scales will not be considered.
However, in order to respect the conservation principles and the agreement with the
behavior of the physical component, the effects of unresolved scales must be
approximated on the resolved scales. This can be done by introducing corrective
coefficients (in engine modeling, these are typically known as, discharge coefficients,
friction coefficients, heat transfer coefficients). Such parameters, requiring
experimental calibration, approximate the unresolved physics that would be otherwise
removed from the model, generating errors.
As introduced above, mean-value models capture only the low frequency
spectrum of a system input/output behavior. For control oriented modeling approach, the
bandwidth of interest is typically associated to the transients that result from variations of
the engine load torque, which can be translated into correspondent variations of the
throttle position, intake manifold pressure and spark timing. With this approach, the
models developed have a time resolution which is adequate to capture the desired details
of the engine throttle-to-torque dynamics, but not the high-frequency behavior (which
can be associated to the fuel system dynamics, the engine cyclical behavior, or to the
tuning effects in the engine manifolds and runners). Such high frequency modes are
typically time-averaged.
The flexibility of these models and their capability of representing the
input/output behavior of the system with reasonable precision but low computational
complexity, have made them a very powerful tool in the analysis, simulation and design
of internal combustion engines.
The MVM technique is applied to each engine component to obtain a
mathematical representation of the input/output behavior. The characterization can be
ME 7236 Module 1 Autumn 2012
1.25
made with different modeling criteria as well as different levels of detail. This process
requires knowledge of all the system components behavior, hence knowledge of different
engineering fields. Focusing on the throttle-to-torque dynamics, all the necessary
principles and information concerning thermodynamics, fluid mechanics, heat transfer,
chemistry and combustion have to be used together. Often, notions of applied mechanics,
electrical and electromechanical systems, and automatic control are also required.
Once all components have been individually modeled, they are coupled together
in relation with the system block diagram representation. The result is a set of differential
and algebraic equations (DAE) describing the dynamic of the entire system. Models
created following this methodology can serve a number of applications which, depending
on the model complexity, range from feasibility and design studies, to optimization,
control design, calibration and validation.

Fundamental equations for modeling fluid systems dynamics
An important step for formulating a MVM of an engine (or any fluid system), is
stating the fundamental equations that represent the time evolution of the system
variables. The equations are based on conservation principles that stem from the
fundamental notions of thermodynamic and fluid mechanics.
The formulation of the fundamental equations for a fluid system assumes the
definition of a suitable model of the flow. Focusing on a macroscopic viewpoint, based
on the continuum scheme commonly adopted in fluid mechanics, two methods of
analysis are available. The first step in both cases consists of drawing an arbitrary closed
volume within a region of the flow field studied. This defines a control volume V, whose
boundary is defined by a control surface A.

ME 7236 Module 1 Autumn 2012
1.26

a) Finite control volume fixed in space

b) Finite control volume moving with the fluid
Figure 1.7: Overview of flow modeling approaches

The Eulerian approach states that the control volume is fixed in space and the
fluid moves through it, as depicted in Figure 1.7(a). Alternatively, as Figure 1.7(b)
shows, in the Lagrangian approach the control volume is moving with the fluid such that
the same fluid particles are always contained in the volume. The conservation laws are
applied to the fluid inside the control volume and, if the Eulerian approach is chosen, to
the mass and energy flows across the control surface.
Most of the practical cases in fluid systems dynamics deal with the study of the
flows into and out of components, an Eulerian approach is usually chosen, defining the
control volume as the physical volume of the component. Hence, the conservation laws
are applied to a control volume fixed in space, where mass, energy and momentum can
flow across its boundary. In this context, a general conservation equation for an extensive
system property can be written in the following form:
{net change in time} = {flow in through boundary}-{flow out through boundary} + {net
generation}-{net consumption}
The conservation laws, whose validity is independent of the nature of the
particular fluid or problem, are a summary of theoretical analyses and experimental
observations and rely on the assumption of nuclear and relativity effects being absent.
The conservation principles are usually expressed in the form of equations:
ME 7236 Module 1 Autumn 2012
1.27
mass conservation equation;
energy conservation equation;
momentum equation.
The principle of mass conservation, referred to an open thermodynamic system of
finite extension and fixed boundary, states that the overall mass of the system is constant,
hence the rate of accumulation of mass within the control volume is equal to the excess
of the incoming mass flow rate over the outgoing mass flow rate:

=
j
out
i
in
j i
m m
dt
dm

(1)
In steady-state (hence, with the left-hand side equal to zero), the equation is also
known as the continuity equation
The energy conservation equation is the expression of the first law of
thermodynamics for open systems, i.e. capable of exchanging mass and/or energy with
the boundary. Considering a flow model of a finite control volume fixed in space such as
in Figure 1.7(a), the principle states that rate of change of the total energy of the system
is equal to the difference between the rate of energy flowing into the system and the rate
of energy flowing out of the system:
( )

+ =
j
j out
i
i in
t
W Q h m h m
dt
me d
j i


(2)
where e
t
is the total energy per unit of mass comprising the thermodynamic internal
energy (u), as well as the kinetic (
2
2
1
c ) and potential energy ( ). The latter includes all
the forms of potential energy associated to conservative fields, such as gravitational or
electrical.
+ + =
2
2
1
c u e
t

(3)
ME 7236 Module 1 Autumn 2012
1.28
Moreover, h represents the enthalpy of the flow, Q

is the total heat exchanged through the


surface of the system and W

is the mechanical power generated by the system.


The momentum equation, unlike the previous two, is not the result of a
conservation principle but rather the application of Newtons second law of motion to an
open thermodynamic system. For this reason, the equation is actually a vector equation
(depending on three directions and time). This law states that the rate of change of the
system momentum in any direction is equal to difference between the incoming and the
outgoing rate of momentum flow and the sum of the external forces acting on the system:

+ =
i
i out in
F M M
dt
M d



(4)
The fundamental equations allow one to form a coupled system of nonlinear
ordinary differential equations in terms of several unknown variables. Therefore, it is
necessary to introduce further assumptions in order to create a model that can be solved
analytically or numerically. To this extent, constitutive relations can be introduced to
complete the equations set.
Unlike the conservation equations, whose validity is completely general, the
constitutive relations are a specific characteristic of the problem. Usually they are related
to the fluid considered and involve its properties, in the following form:
( ) 0 , , = T v p f (5)
where p is the fluid pressure, v the specific volume (per unit of mass) and T the
temperature.
The expression of a constitutive relation generally leads to a set of algebraic
equations, even though ordinary differential equations are possible in some cases.
Constitutive relations often derive from assumptions made on the nature of the fluid. For
instance, if the fluid can be considered incompressible, the relation will simply become
. const v = Moreover, for most of the applications involving liquids, constant temperature
is an acceptable assumption.
ME 7236 Module 1 Autumn 2012
1.29
Since internal combustion engines operate with gaseous fluids, another important
constitutive relation that will often be used is the equation of state for a perfect gas:
mRT pV = (6)
where R is the specific gas constant. The perfect gas model is a reasonable approximation
of the behavior of all compressible fluids, provided their thermodynamic state is far
enough from the saturation conditions (for example, during evaporation or condensation).
The perfect gas model implies further assumptions that allow one to close the
entire equations set. For example, the thermodynamic internal energy and enthalpy (used
in the energy equation) are functions of the sole gas temperature:
( )
( ) T c T h h
T c T u u
p
v
= =
= =

(7)
where c
v
is the specific heat at constant volume and c
p
the specific heat at constant
pressure. Equation (7), sometimes known as the caloric equation of state, will be further
applied in the following modules.

Model coupling techniques
One of the advantages of using the input/output representation is that this
approach emphasizes modularity. Hence, large and complex powertrain systems can be
decomposed into elementary components interconnected. If each component is modeled
to be easily interfaced at the input and output ports with other subsystems, it is possible
to use standard component to assemble complex systems in a simple and
straightforward procedure. At the same time the calibration effort is reduced, because it
can be made separately for the model of each component and then only minor
adjustments are required on the final assembly.
In order to achieve these benefits, a suitable coupling concept must be defined for
multiport interconnected systems. Several approaches have been proposed in literature,
the most important ones relying on transfer functions, transfer matrices and bond graphs.
However, in the analysis of internal combustion engines (or any fluid system), a simple
ME 7236 Module 1 Autumn 2012
1.30
and immediate representation can be obtained by dividing the system input and output
variables in two categories:
- Level Variables: in general they are differential variables provided by the
fundamental equations that indicate the amount of thermodynamic
properties stored inside a component;
- Flow Variables: they usually relate to the flow of a specific variable
through the control surfaces.
Examples of level variables may be mass (either total mass or mass of individual
components in a mixture), internal energy, or kinetic energy; flow variables may be mass
flow rate or enthalpy flow rate.
Likewise, when modeling any physical system there are two main classes of
objects that must be considered:
- Reservoirs: these components are characterized by one or more states that
represent the stored amount of level variables (state determined system);
- Flow Control Devices: they determine the amount of properties that flow
through the component itself, typically as a result of differences between reservoir levels
(purely algebraic system).
Reservoirs receive flow variables as inputs and their outputs are level variables.
Conversely, flow control devices receive level variables and determine the flows
associated. Figure 1.8 briefly summarizes the concept.

Figure 1.8: Classifications of systems components and signals

The representation adopted facilitates the connection between components,
solving typical causality problems that are associated with dynamical systems modeling.
As Figure 1.9 shows, alternating reservoirs (state determined systems) to flow control
devices allows one to respect the cause and effect priorities between the input and output
signals of each block, allowing to immediately identify the driving and driven variables.
ME 7236 Module 1 Autumn 2012
1.31

Figure 1.9: Example of connection between components

Modeling guidelines
The modeling approach described so far can be summarized with a sequence of
procedures:
1. apply the system dynamics approach to deconstruct the engine and/or
powertrain system into a series of fundamental components, therefore
determining the system boundaries, inputs and outputs);
2. identify the relevant sources of dynamics by choosing a number of
reservoirs and their corresponding level variables. The number of state
determined components will influence the order of the model (i.e., the
number of state equations);
3. using the fundamental equations and the fluid properties, formulate
differential equations for all the state determined components;
4. using the fundamental equations in quasi-static conditions (i.e., without
the time derivatives) and the fluid properties, formulate algebraic relations
for the flow resistance components, relating the flows between the
reservoirs as functions of the state variables;
5. as a result of the quasi-static approximations, the algebraic equations will
be characterized by a number of unknown parameters that need to be
identified from experimental data or other available information;
6. once the calibration is done, assemble the components into the overall
system model;
7. validate the complete model on a set of data points that have not been used
to identify the parameters.

ME 7236 Module 1 Autumn 2012
1.32
1.5. TWO IMPORTANT ENGINE CONTROL PROBLEMS
This section introduces representative control strategies used in production
vehicles for air-fuel ratio (AFR) and idle speed control. These control strategies use both
look-up table based control schemes and dynamic feedback or feedforward control
schemes. The engine control unit (ECU) has limited computational capability; however,
a large memory space is available through the use of Read Only Memory (ROM). The
use of sensors is limited due to cost factors. In keeping with the relatively slow
computational speed and large memory of the ECU, current look-up table based control
strategies work well. However the calibration processes associated with such tables
require a significant amount of time and effort. The development period of a new vehicle
therefore is significantly affected by the time duration of these processes. The use of
dynamic models may be useful in reducing this development time by reducing the overall
calibration effort.

The AFR Control Problem
The AFR control problem for an internal combustion engine is not limited
to the control of AFR about stoichiometry. There are multiple objectives that need to be
achieved via AFR control. These are essentially dependent on the various engine
operating conditions, including engine load engine speed, coolant temperature,
acceleration, deceleration, etc. However, we may generally classify the AFR control
problems into two groups: open-loop control and closed-loop control. Modern ECUs
have the capability to decide whether the control of AFR is to be closed-loop or open
loop. This is achieved by using information from various sensors installed on the engine
subsystems. These two types of strategies are discussed next.
Open-loop AFR Control
Open loop AFR control is essentially the control of the AFR at or about a desired
value through the use of (often elaborately) constructed look up tables that encompass
various engine-operating conditions. When the ignition key is turned on and the engine
cranks, a rich air-fuel mixture is required to guarantee the initiation and sustenance of
ME 7236 Module 1 Autumn 2012
1.33
stable combustion. The amount of fuel to be injected is primarily dependent upon the
temperature of the intake manifold with the assumption that the engine motoring RPM
remains constant, as does the intake pressure at each cranking instant. Since most of the
current mass-production engines do not have an intake manifold wall temperature sensor,
engine coolant temperature information is used. This is necessary in order to compensate
for the fuel dynamics due to the wall wetting phenomenon
1
. Air flow rate information,
which also plays an important part in these estimations, is supplied via a mass air flow
meter or other kinds of air metering sensors. Too much fuel, however, is not always
good for engine cranking. If too much fuel is injected during engine cranking, it may
cause flooding of the spark plug electrode gap that will in most cases results in no spark.
Therefore, when a repeated engine-cranking situation is detected by the ECU within a
relatively short time period, it is easily implied that the previously injected liquid fuel has
not had sufficient time to evaporate. The ECU should therefore have an algorithm to limit
the amount of fuel injected such that the overall liquid fuel present in the intake system is
compatible with optimum engine performance under a cranking scenario.
At the onset of engine cranking, all the injectors start to inject fuel simultaneously
(full group injection) upon the detection of crankshaft tooth wheel signal. Normally, full
group injection is performed only once, and the amount of fuel to be injected is
dependent on the coolant temperature, and not on the air flow rate. This is due to the
inherent delay of the air-metering sensor not allowing sufficient time to both sense and to
calculate the amount of air entering the engine. During the engine cranking, the fuel
amount from the injector solenoid valve is heavily dependent on the battery voltage,
because the starter motor draws large amounts of current from the battery, causing a
temporary drop in battery terminal voltage. The battery voltage may drop to around 7
Volts in the worst case (at extreme cold engine conditions), thus decreasing the amount
of current that passes through the injector solenoid coil. The less the current at the coil

1
I.e.: the tendency of injected fuel to condense in a puddle in the inlet port, and then to evaporate
at a rate dependent on local pressure and temperature conditions. A more detailed account is given in
Module 3.
ME 7236 Module 1 Autumn 2012
1.34
the lower the lift of injector plunger, thereby injecting less fuel. Therefore, compensation
of battery voltage drop via increasing injection pulse width at cranking phase is another
important consideration. The detection of the specific cylinder that is to be fired after the
first firing is decided upon by the combination of the crankshaft sensor signal, the
number 1 cylinder Top Dead Center (#1 TDC) sensor signal, and the engine firing order.
Since full group injection causes excessive fuel at the beginning of cranking, this results
in a high HC emission immediately after the engine cranking. To reduce the cold HC
emission, some car manufacturers are adopting a static #1 TDC sensor, typically mounted
on the camshaft, thereby avoiding full group injection. Half group injection, which
injects fuel on one bank of the cylinders, is thus a good alternate technique for firing
initiation at cranking.
In addition, the AFR is controlled to be rich at Wide-Open Throttle (WOT)
condition. Since the primary objective at WOT acceleration is to get enough engine
power for better driveability performance, AFR control remains in a rich state. When
there is an abrupt acceleration command from the driver, the engine experiences a lean
spike mainly due to the time lag of the air-metering sensor and ECU calculation time
delay. Thus it usually results in under fuelling during sudden accelerations. One typical
strategy to compensate the lean spike at the sudden acceleration situation is the non-
synchronous injection timing scheme. Under normal operating conditions, fuel is
injected based on predefined engine events, that is the firing order. During instances of
sudden acceleration the ECU drives additional injection through an interrupt signal. This
is non-synchronous injection. Non-synchronous injection is applied to the injector of the
next firing cylinder only on the detection of a hard acceleration scenario by the ECU.
This situation is detected through the use of the TP sensor signal. The ECU is able to
discriminate between normal acceleration rates and hard accelerations by looking at the
rates of the throttle position change. The amount of fuel to be injected in these
conditions is dependent on the engine speed, the pressure at the intake runner, and
temperature of the intake subsystem. In most of the mass-production engines, the so-
called transient fuel compensation on throttle variation is based on the above mentioned
engine speed, pressure and coolant temperature.
ME 7236 Module 1 Autumn 2012
1.35
The injection timing of each cylinder will be optimum if injection is terminated
before the start of the intake valve opening. This provides sufficient time for the liquid at
the intake port and valve to evaporate, thereby ensuring a homogeneous mixture of air
and fuel. The homogeneity of mixture plays an important part in the reduction of exhaust
emissions over the entire engine running range.
Not only does the lean spike worsen driveability and exhaust emissions but it also
tends to induce engine knock. A rich AFR has other advantages. Traditionally the
engine-knocking problem at the beginning of a sudden acceleration was rectified mainly
by spark timing control. In most mass-production vehicles today, enrichment of air-fuel
mixture is also introduced to reduce knocking tendencies in the engine at the initiation of
a sudden acceleration. This is because a rich AFR condition improves flame propagation
speed in the combustion chamber.
On the other hand, fuel cut-off (or reduction) is necessary for deceleration
condition. The purpose of fuel cut-off is to improve fuel economy. Among the factors
which affect the fuel cut-off are: coolant temperature; engine speed; and air conditioner
compressor engagement status. The status of the compressor engagement is an important
factor in deciding the fuel cut-off. A fuel cut-off with the compressor engaged, which
can be considered an external load, could result in an unstable engine operation or in
the extreme case engine stall. Hence it becomes necessary to consider the compressor
engagement status even when looking at speed ranges for fuel cut-off. Several other
considerations come into play. For example, when the engine is below the base engine
warm-up temperature, there will be relatively large engine torque loss due to increased
engine friction as a result of high lubricant viscosity. Therefore, if the fuel-cut is
executed at low engine temperature conditions while decelerating, such undesirable states
may be achieved as discussed above.
The basic injection look-up table, which is obtained through steady state engine
dynamometer test, has two independent variables: engine speed and engine load. By
breaking up the engine speed and load into several points, various steady state test
condition-operating points are determined. These operating conditions are referred to as
break points. By running the engine at these break point conditions, the injection pulse
ME 7236 Module 1 Autumn 2012
1.36
width for the desired AFR (stoichiometry) is thus calibrated. The fuel injection look-up
table uses units of milliseconds as the time scale. The basic fuel injection look-up table,
therefore, implicitly includes such static engine parameters as charge efficiency, residual
gas, fuel evaporation at the intake ports, injector dead time and air metering sensors
(pressure sensor, MAF, etc) static characteristics.
Since the basic fuel injection look-up table is constructed for standard conditions,
there are many other factors relating to ambient conditions and powertrain system that
must be taken into account as correction terms for open-loop control. These correction
groups consist of battery voltage, coolant temperature, ambient air temperature, altitude
compensation, transient compensation (throttle and engine load variations), anti-
knocking, automatic transmission compensation, etc. The corrections are added or
subtracted from the basic fuel injection look-up table, some of them are even multiplied
to calculate open-loop injection pulse width. Increasing the fuel pulse from the open-
loop value is very important for robust AFR control of an engine. The open-loop fuel
and closed-loop fuels are in a trade-off relationship with each other. That is to say that if
for a total fuel amount to be injected, if the open-loop related fuel quantity is increased,
then the closed-loop portion of the fuel has to be decreased. By reducing the portion of
fuel commanded in closed-loop, a faster and more stable AFR control can be achieved.
The ECU being used today still often consists of 16-bit microcomputers with an
increasing number of applications employing 32-bit architecture. Automotive
microcontrollers are characterized by fixed-point arithmetic, relatively slow processor
speeds, and substantial amounts of memory. For this reason, the use look-up tables for -
loop control has been historically a preferred choice for control applications. With the
introduction of 32-bit controllers, it is conceivable that in some instance dynamic control
strategies might replace some of todays look-up tables, and that the calibration load
might be subsequently reduced.

ME 7236 Module 1 Autumn 2012
1.37
Closed-Loop AFR Control
Closed-loop AFR control corrects the quantity of fuel injected as a function of the
signal sensed by the EGO sensor. In most gasoline engine applications, binary EGO
sensors have been used as the exhaust gas AFR sensing device. Due to the cost and
insufficient reliability of the Universal Exhaust Oxygen (UEGO) sensors, these sensors
are seldom used for the AFR control except for lean-burn engine applications.
The basic AFR control strategy that is commonly applied to mass production
vehicles is a Proportional and Integral (PI) control algorithm with gain scheduling
dependent on how long the EGO sensor signal stays at one level (lean or rich). The
closed-loop correction term oscillates as a function of the state of the EGO sensor. The
ECU employs fuel compensation by using a P-correction that intervenes when the EGO
sensor switches, and an I-correction that intervenes when the EGO sensor stays at the
same level. Therefore, the P-gain comes into play instantaneously with the EGO sensor
switching and the I-gain compensates progressively while the EGO sensor stays at one
level (lean or rich). The PI gains are not always the same; instead, these gains change to
adjust according to the engine operating conditions, and on the number of steps of I-
correction already applied. The duration of the EGO signal staying at on one level also
has an impact in gain scheduling. This variable gain concept affords a fast control of
AFR to stoichiometry in the presence of large AFR deviations.
For a well-tuned engine system with an appropriately calibrated ECU, the mean
value of PI-correction will oscillate about a zero value. However, there are various
factors that keep the mean PI-correction from being zero. These factors include the
following:
- Fuel system degradation and tolerance: degradation of fuel pressure regulator, fuel
pump deterioration, injector clogging, and the use of gasoline other than that used for
calibration may result in the PI control behaving as if for a lean system. This
misinterpretation by the control scheme will result in a shift of the mean correction
value from zero.
ME 7236 Module 1 Autumn 2012
1.38
- Leakage of air into the intake system leading to an unmeasured quantity of air being
added to the engine, thus leading to a lean excursion. The effect of leakage on MAF
sensor-based systems might be more critical than on speed-density systems.
- Aging and tolerance of air measurement sensors: degradation of not only the air
measurement sensors but also of such sensors as those that are used for fuel
calibration, air temperature and coolant temperature will affect the mean PI-
correction shift.
- Air density change: the altitude effects on air density will affect fuelling over the
entire engine operating range.
If no closed-loop AFR control compensation is introduced, the previously listed
factors may force the engine to have a steady-state fuelling error. Therefore, a self-
adaptive AFR control algorithm is necessary to effectively control for each of the engine
specific circumstances. When the listed factors occur, the closed-loop PI-correction
strategy comes into play and compensates for the AFR deviation, thus resulting in the
mean PI-correction value to shift from zero. If the PI-correction deviation remains for a
specified time duration, the ECU may add/subtract fuel corresponding to the amount of
PI deviation from the open-loop fuel calculation sum (additive compensation). Thereby,
keeping the AFR at stoichiometry with the mean PI-correction being forced back to a
zero value. The self-adaptive additive compensation is especially effective when an
engine is experiencing intake air leakage or changes of injector delay time. If the shift of
mean PI-correction value is caused due to an air density change, a multiplication of scale
factors is more effective to cover the whole engine operating range (multiplicative
compensation).
The engine and ambient conditions that introduce the deviation of mean PI -
orrection have different magnitude of influence depending on the engine speed and load.
Therefore, it is necessary to distinguish the engine operating areas into several sections to
find which factors are dominant in that particular area. Figure 1.10 specifies these areas.
Intake leaks need to be compensated at low-load conditions as an additive term to the
open-loop injection map, because leakage air might be magnified at low engine load
ME 7236 Module 1 Autumn 2012
1.39
(area 2: high vacuum condition). Injector delay, which affects each engine stroke, has to
be compensated as an additive term at high engine speed and low-load (area 3). Finally,
fuel system degradation and air density changes might effect the entire engine operation
range. So the open-loop fuel map has to be multiplied by scaling factors to compensate
across the whole engine operating range (area 1) except WOT condition where the AFR
is not controlled to stoichiometry.
Area
2
Area
1
Area
3
WOT
Max
speed
Min Load
Min
Speed
Engine
Load
Engine
Speed

Figure 1.10: Areas for additive and multiplicative correction.

The Idle Speed Control Problem
The objective of idle speed control is to maintain a smooth and comfortable
driving condition while minimizing fuel consumption rates. To get good fuel efficiency,
the engine idle speed should be set to a very low RPM. If we set the idle RPM too low,
however, then the idle speed control tends to become very unstable as the engine may be
producing insufficient torque at this engine RPM. Therefore in deciding the idle RPM
one has to consider an operating speed where torque production is robust enough to reject
disturbances from various sources. Moreover for drive smoothness and comfort it is
necessary to control the idle speed within a very narrow range. This would imply
maintaining the engine speed at almost a constant level. This would be a trivial task if
there were no unexpected disturbances. However, in a real life scenario there are various
kinds of disturbances present during an engine idle condition; these are:
ME 7236 Module 1 Autumn 2012
1.40
- Engagement of automatic transmission: shifting the transmission from neutral to the
drive range adds a torque disturbance to the engine through the torque converter.
The reverse procedure (drive to neutral shift) also introduces the same torque
amount as a subtractive disturbance to the engine. Moreover, the amount of this
disturbance is not always same; instead it is dependent on the temperature of the
torque converter fluid because torque transmissibility is related to the viscosity of
Automatic Transmission Fluid (ATF).
- Electrical disturbance: due to the operation (on/off) of such electrical devices as
defroster, head lamp and direction lamps, the engine speed is affected. The engine
speed drop is dependent in the electrical capacity of such devices.
- Direct engine torque disturbances: the power steering pump and air conditioning
compressor is good examples of direct torque intervention.
- Canister purge valve on and off at engine idle: as discussed in the Chapter 1, the
opening of CPV causes unmeasured air and fuel introduction to the engine system,
thereby prompting an increase engine speed. In contrast, closing of the CPV will
induce an engine speed drop.
For engine applications with small displacement volumes, the above mentioned
disturbance effects are more dominant as compared to engines with a larger
displacement volume. This necessitates the use of more sophisticated idle speed control
strategies for small displacement volume engines. Driver comfort is a strong motivation
for controlling the idle speed to the nominal value. It is common knowledge that a
change in the nominal idle speed affects driver comfort, as discussed above. However,
it needs to be mentioned that a drop in the nominal idle speed can affect the comfort
level more drastically than a similar increase in the idle speed. Researchers in the
automotive industry have therefore been concentrating on investigating methods of
avoiding engine speed decrease during idle conditions. Ignition timing and idle air
control strategies are two methods popularly adopted for controlling idle speed.
ME 7236 Module 1 Autumn 2012
1.41
Idle Air Control
Idle air control is often accomplished through the use of either a stepper motor or
a rotary solenoid valve actuating the air by-pass valve. This affords control of the section
area of the air passage thereby allowing a control over the amount of air entering the
engine. The airflow through the Idle Air Control Valve (IACV) can be managed through
a control over the valve flow area; this can be accomplished by varying the actuator duty
cycle. The factors affecting the calculation of duty cycle vary with engine running
conditions.
During engine cranking and warm-up phase, the IACV is opened as a function of
coolant temperature corrected for altitude. The nominal engine speed is then decided
upon by relying on the engine coolant temperature; for cold engine conditions, the
nominal engine speed is maintained high so as to make up for the high friction torque
loss. The idle RPM is then gradually reduced to the idle speed normal for a warmed up
engine. This procedure can be achieved using a two-dimensional look-up table (coolant
temperature vs. idle valve duty cycle). Altitude compensation is implemented by means
of a multiplying factor for the look-up table, thereby ensuring sufficient airflow into the
engine at high altitudes. Thus, the altitude compensation factor should be larger than
unity. Detection of altitude other than the level of calibration can be easily done using
the Manifold Absolute Pressure (MAP) sensor. When the ignition key is turned on
without the engine running, the MAP sensor senses the ambient pressure. With the
engine running, ambient pressure can be continuously updated by sensing the maximum
intake pressure for low RPM at high engine loads. This is so because the pressure
existing in the intake manifold during such a condition closely approximates ambient
pressure. At the engine warmed-up condition, the idle speed duty cycle calculation is
dependent on engine events and conditions. The calculation of idle duty cycle can be
subdivided into four main parts; these are open-loop duty cycle, vehicle event based
corrections, closed-loop, and self-adaptive compensations.
ME 7236 Module 1 Autumn 2012
1.42
Open-loop Base Idle
Duty Cycle
Battery Correction
Engine Load
Correction
Dash Pot Correction
Closed-loop PI
Correction
Self-adaptive
Correction
Others
Bat Voltage
Duty
Load
Duty

Final Idle
Duty
Command
Vehicle
Condition
Based
Corrections

Figure 1.11: Final idle air bypass valve duty command calculation.
Figure 1.11 shows how these correction terms are added to calculate the final idle duty
cycle. The detection of vehicle on or off events such as air/con, cooling fan, electric load
and transmission neutral to drive shift can be done via various sensory systems of a
vehicle. Each of the above mentioned four-correction terms are calculated respectively
and added to finalize the total idle duty cycle to be output. Following is a brief discussion
of how each of the corrections are calculated and calibrated:
ME 7236 Module 1 Autumn 2012
1.43
Leakage Air
Altitude Correction
Nominal Engine
Speed Correction
Cooling Fan
Correction
Transmission Neutral
Drive Shift
A/C
Others
Coolant Temp.
Duty
Air Temp.
Duty
Ambient Press.
Factor
Coolant Temp.
Duty
ATF Temp.
Or
Coolant Temp.
Duty

Switch
Switch
Switch
Switch
Open-Loop
Base Idle
Duty Cycle
Base
Terms
Vehicle
Event
Related
Terms

Figure 1.12: Schematic of Open-loop Base Duty Cycle.

Open loop duty cycle calculation: the open-loop duty cycle is calibrated by
referring to the coolant and air temperatures, and the ambient pressure along with the
corrections associated with vehicle events that inject torque disturbances to the engine.
During the engine idle condition, the intake manifold pressure level is low (high vacuum)
hence the airflow through idle valve can be modeled as a one-dimensional choked flow.
In addition, most of the idle speed valves are designed to allow a linear relationship
between the duty cycle and airflow rate. The airflow rate is primarily dependent on
ambient pressure. Therefore, air temperature and ambient pressure are base factors that
ME 7236 Module 1 Autumn 2012
1.44
decide the amount of air that passes through the idle valve. As is shown in Figure 1.12,
all the open-loop duty can be sub-grouped either as base terms or as vehicle events
related terms. The base terms consist of leakage air, nominal engine speed correction and
corrections for air and coolant temperatures. Air leakage defines the air that leaks past
throttle valve seating circumferentially while the throttle is closed. The leakage occurs
through an annular aperture because the valve can not complete seal the throttle body.
Selection of such valves is therefore necessarily based on the valve satisfying the
specifications on leakage limits at standard operating conditions. Air leaking past the
throttle valve will therefore add to the idle airflow, hence throttle valves with small
leakage are preferred for idle air control purposes. This allows the idle air control valve
to have a larger control span. The corrections based on air and coolant temperature
contribute to the idle duty calculation. Since the coolant temperature level determines
nominal idle RPM, this correction factor is added to compensate the nominal idle speed
differences. The vehicle event related terms are added to the base terms as a correction
factor. Turning on the A/C compressor upon the drivers command will result in a steady
state engine speed error if the system lacks compensation. Thus increment of idle duty
cycle is required to maintain the nominal engine idle speed. The neutral to drive shift of
the transmission also introduces a similar kind of steady state error and thus needs to be
compensated. Each of the vehicle events related terms have to be decided upon by
empirical means. This is done by subjecting the engine to each of these conditions. The
calibration procedure thus requires a lot of time and effort. However, this open-loop
calibration procedure reduces the closed-loop contribution thus leading to a fast and
robust idle speed control. In most of the engine applications, the nominal idle speed
between A/C on and off are set differently. The nominal engine speed level for A/C on
condition is higher than that of A/C off condition to provide more engine torque to
compensate the disturbance from the A/C compressor.
Anticipating Disturbance Torque: One of the most effective ways of rejecting a
torque disturbance is anticipating it before it occurs. Due to the large delay between the
opening of the idle valve to the production of engine torque, an engine will experience
abrupt engine speed fluctuation whenever accessory loads disturbance are applied or
ME 7236 Module 1 Autumn 2012
1.45
removed. Many accessory load disturbances can be anticipated to allow for air
compensation via the idle valve. For instance, if an ECU receives a command to turn on
A/C from driver, then it begins to compensate air via the idle air valve before the
compressor is actually turned on. After the torque production time delay the engine will
have developed the torque lost to the compressor related disturbance. The ECU will then
send a command to turn on the compressor without affecting the engine RPM.
Therefore, anticipating the disturbance torque is a problem that determines how much air
needs to be compensated. The time period between the start of idle valve compensation
and the instant of compressor turning on also should be calibrated. The neutral to drive
shift of the transmission can also be anticipated through the detection of transmission
lever movements. However, the amount of idle duty compensation required is based on
the temperature of the ATF. The viscosity differences, due to the various possible
temperatures of the ATF, result in different levels of torque disturbances acting on the
engine. If the ECU does not have access to the ATF temperature information, engine
coolant temperature may be used instead. Using the procedures explained above, almost
all of the torque disturbances can be effectively anticipated. As long as the vehicles
electrical components are under the control hierarchy of the ECU, the electrical
disturbances to the idle speed control can be anticipated too. These electrical loads
include radiator cooling fans, headlamps, defroster, etc. The anticipating control scheme
becomes more essential for small displacement volume engines.
Vehicle Events Based Corrections: There are other vehicle-running conditions,
however, that must also be considered. These consist of corrections that compensate for
battery voltage, engine load and dashpot function, etc. The idle valve is essentially a
solenoid valve or stepper motor hence the opening section area is effected by the battery
voltage. If a vehicles battery voltage deviates from the standard calibration condition,
then the voltage needs to be compensated. The engine load, if it is different from the
value at the time of calculation, needs to be compensated too. Different engine loads
other than those during calibration may effect the amount of air that passes through idle
valve. In addition, the load difference also effects the charge efficiency, which
eventually changes the air that enters the engine. The dashpot correction plays an
ME 7236 Module 1 Autumn 2012
1.46
important role at the moment of throttle valve closing. When the throttle valve abruptly
closes during deceleration, the intake manifold will be in a high vacuum, presenting a
possible risk of damaging the rubber hoses and connections attached to the intake
manifold. Moreover, the Positive Crankcase Ventilation (PCV) system can leak
lubricating oil into the intake manifold. Another contribution of the dashpot function is
to reduce the rich spike of the AFR during deceleration. At this condition, the wall-
wetted fuel from the previous cycle continues to enter the cylinder although deceleration
fuel cut-off occurs. Thus leading to a rich AFR spike. However, by opening the idle
valve at a pre-defined mode, above listed high vacuum and AFR rich excursion problems
can be solved.
Closed-loop Idle Air Control: The closed-loop idle control strategies rely on the
conventional PI-control scheme with various gain-scheduling concepts. Each gain is
tuned according to the difference of the specific target RPM and current engine RPM.
The Crankshaft Position Sensor (CPS) usually performs the measurement of engine
RPM. As in the same case of the injection pulse width control, the P-gains intervene as
a correction for instantaneous engine RPM fluctuation while I-gains compensate stead-
state engine RPM error. For a new, well-tuned engine system, the mean value of PI-
correction is expected to be centered at zero. However, as the engine intake system ages,
deposits of dust and oil mixtures can be found on the throttle and idle valves. This
process leads to among other effects a decrease in the idle airflow rate owing to a
constriction of the air passages. Variations among engines attributable to manufacturing
processes may also force the closed-loop controller to come into play. The amount of
shift can be separately compensated through self-adaptive correction, thus returning the
mean PI correction value zero. The self-adaptive correction, therefore, provides more
room for closed-loop idle air control.
ME 7236 Module 1 Autumn 2012
1.47
Leakage
Self-adaptation
Open-loop
Closed-loop
Idle Air
Controllable
Air

Figure 1.13: Idle Air Control.
Strategies for idle valve control have been explored. The contributions of
leakage, open loop, closed-loop and self-adaptive air correction have been reviewed.
Figure 1.13 illustrates how these corrections are linked together. Elaborate calibration
work makes it possible to reject almost all disturbances that are expected for various
engine conditions. However, for unanticipated disturbances this scheme suffers in its
effectiveness due to the inherent delay of the idle speed air control strategy.

Ignition Timing Control
Ignition timing control schemes are used to effectively compensate for the
inherent delay of the idle air control method. The torque production delay associated
with the airflow dynamic is substantially larger than that for ignition. It is also known
that changes in ignition timing will affect the engine torque production instantaneously.
This characteristic can therefore be used as an idle speed control method under dynamic
engine conditions.
ME 7236 Module 1 Autumn 2012
1.48
Ignition Timing (BTDC)
[Degree]
MBT
Normalized
Torque
1
0.9
35 20 10 0

Figure 1.14: Schematic of Engine Torque vs. Ignition.
Engine torque production is dependent on ignition timing. Figure 1.14 shows the
effect of variations in ignition timing on brake torque for typical spark-ignition engine
[6]. Ignition timings can typically be varied to be advanced or retarded. An ignition
advance implies increasing the angular position of ignition initiation with respect to Top
Dead Center (TDC). Retarding the ignition timing would mean moving the initiation of
ignition closer to TDC. The engine torque increases as the ignition timing is advanced
until it reaches to the Maximum Brake Torque (MBT) point. Further advancing the
ignition timing may result in the engine knocking. Substantially retarding the ignition
timing on the other hand will cause the engine to misfire. Hence the importance placed
by automotive manufactures on efficient calibrations for optimal ignition tuning.
Calibration of MBT or optimum ignition timing at various engine loads and
speeds forms a basis for ignition control. To find the MBT point, an engine is set up on a
dynamometer. Running the engine at various predefined load and RPM breakpoints, the
ignition timing is adjusted to for MBT. Thus the MBT ignition timing provides the best
torque at each of the engine running conditions tested. Knocking points exist for all
engines; it is preferred to have knock points after the MBT point. However, hardware
design may force the ideal MBT point to lie after the knock point. Continuous knock,
however, needs to be avoided since it can seriously damage the combustion chamber
ME 7236 Module 1 Autumn 2012
1.49
including valves and piston. Once the knocking ignition timing is found, the optimum
ignition timing should be decided by retarding a few degrees away from the knocking
point. This provides a margin of safety to allow for engine to engine production
variations. If an engine has a knock detection sensor installed, the ignition margin of
safety can be reduced, thus advancing the ignition timing by a few degrees. This will
result in the production of more engine torque. Ignition timing relies heavily on engine
RPM when the engine load is constant, the ignition timing needs to be advanced as the
engine RPM increases. This is done to allow the combustion process similar time
duration at all engine speeds. The basic ignition map for part load and full load
conditions is found by considering the previously discussed factors. This leads to the
steady-state ignition map for part load and full load.
Retarding the ignition timing from MBT or optimum torque ignition timing
should set the basic ignition map for the idle condition. Idle ignition map is necessary if
the engine were to be able to operate within a torque range. For instance, selecting the
basic ignition timing at, 10 degrees before TDC, the ECU can have room for controlling
the torque through ignition timing variation. The ECU may therefore retard or advance
the ignition timing through control algorithms to reject unknown disturbances.
Generally, the ignition timing can vary by about 10 degrees centered at the calibrated
basic idle ignition timing. In most cases, the map of idle ignition timing relies on the
measured coolant temperature. The basic ignition timing is usually set to be advanced,
with the coolant temperature low, to provide more torque. Then as the coolant
temperature increases, the basic idle ignition timing is retarded gradually. To
accommodate different disturbances like A/C or transmission shift, several maps will be
useful for compensation.
Ignition control at idle is done mostly by proportional correction with various
gains. Since the prime aim for the use of ignition timing is to compensate dynamic
engine rpm fluctuation, integral control is seldom used. Instead the steady-state error is
corrected by idle air control. Knowing that the engine torque is linearly proportional to
the ignition advance in the vicinity of idle ignition control range, the amount of ignition
timing compensation can be calculated simply by multiplying a gain to the difference of
ME 7236 Module 1 Autumn 2012
1.50
the filtered and the current engine RPM. In addition, since the shape of the engine torque
curve is different when the coolant temperature changes, gain scheduling is needed to
take into account the coolant temperature variation with the engine running. Therefore,
the proportional gains are determined based on the different coolant temperatures.
Although the ignition timing can reject unknown and high frequency disturbances
effectively, it is not appropriate to use the ignition timing to compensate steady-state
engine RPM deviation. Consequently, the idle ignition map provides control capability
by sacrificing engine torque.

Summary
Typical production control strategies for both AFR and idle speed control have
been explored. Most of these control strategies use event-based methods rather than
using model-based control method. Thus lots of experimental works are involved in
calibration procedure. Also, since there are too many calibration factors involved each
other like a web, changing one calibration factor may influence to the other control
performance that is difficult to predict. Although all the current production control
strategies may not be replaced by the model-based control schemes, engine subsystem
models may replace part of them; thus, providing a more simple and systematic approach
to the control problems.
ME 7236 Module 1 Autumn 2012
1.51
1.6. REFERENCES
[1.1] J. Lowen Shearer, Bohdan T. Kulakowski, and John F. Gardner Dynamic Modeling and Control of
Engineering Systems", 2nd Ed., Prentice-Hall, 1997.
[1.2] " Matlab/Simulink Introduction", GMTEP Technical Presentation.
[1.3] G. Rizzoni, K. Srinivasan, S. Yurkovich, "Signal and System Dynamics Integration", Notes for GMTEP
Technical Presentation GM 1989.
[1.4] M. Ross, R. Goodwin, R. Watkins, M. Wang, and T. Wenzel, Real-world emissions from Model Year 1993,
2000 and 2010 passenger cars, distributed by the American Council for an Energy-efficient Economy, USA, November
1995.
[1.5] PNGV
[1.6] H Heywood, J. R., Internal Combustion Engine Fundamentals, McGraw Hill Publishing Company, 1988.
[1.7] D. J. Dobner, An engine model for dynamic engine control development, ASME Winter annual meeting,
Paper No. WA4-11:15. 1986.
[1.8] B. K. Powell and J. A. Cook, Nonlinear low frequency phenomenological engine modeling and analysis,
Proceedings of the American Control Conference, pp.332-340, Minneapolis, MN, June, 1987.
[1.9] J. J. Moskwa and J. K Hedrick, Modeling and validation of automotive engines for control algorithm
development, Advanced Automotive Technologies-1989, pp.237-247, ASME DSC-vol.13.
[1.10] E. Hendricks and S. C. Sorenson, SI engine controls and mean value engine modeling, SAE Technical
Paper No. 910258, 1991.
[1.11] R. C. Turin and H. P. Geering, Model based adaptive fuel control in an SI engine, SAE Technical Paper
No. 940374. 1994.
[1.12] A. Amstutz, N. P. Fekete and J. D. Powell, Model-based air-fuel ratio control is SI engines with a switching
type EGO sensor, SAE Technical Paper NO. 940972, 1994.
[1.13] B. A. Ault, V. K. Jones, J. D. Powell and G. F. Franklin, Adaptive air-fuel ratio control of a spark-ignition
engine, SAE Technical Paper No. 940373, 1994.
[1.14] P. Azzoni, D. Moro, F. Ponti and G. Rizzoni, Engine and load torque estimation with application to
electronic throttle control, SAE Technical Paper No. 980795, 1998.
[1.15] C. F. Chang, N. P. Fekete, A. Amstutz, and J. D. Powell, Air-fuel ratio control in spark ignition engine
using estimation theory, IEEE Transactions on Control System technology, Vol. 3, pp. 22-31. March, 1995.
[1.15] D. Cho, D. and J. K. Hedrick, A nonlinear controller design Method for fuel-injected automotive engines,
ASME Journal of Engineering for Gas Turbines and Power, Vol. 110, pp. 313-320, July, 1988.
[1.16] J. W. Grizzle, J. A. Cook and W. P. Milam, Improved cylinder air charge estimation for transient air fuel
ratio control, Proceedings of the American Control Conference, Baltimore, MD, June, 1994.
[1.17] J. W. Grizzle, K. L. Dobbins and J. A. Cook, Individual cylinder air-fuel ratio control with a single EGO
sensor, IEEE Transactions on Vehicular Technology, Vol. 40, No.1, pp. 280-286, February, 1991.
[1.18] A. G. Stefanopoulou, J. A. Cook and J. W. Grizzle, Modeling and control of a spark ignition engine with
variable cam timing, Proceedings of the American Control Conference, pp. 2576-2581, Seattle, WA, June, 1995.
[1.19] E. Hendricks, T. Vesterholm and S. C. Sorenson, Nonlinear, closed loop, SI engine control observers, SAE
Technical Paper No. 920237, 1992.
[1.20] N. P. Fekete, U. Nester, I. Gruden and J. D. Powell, Model-based air-fuel ratio control of a lean multi-
cylinder engine, SAE Technical Paper No. 950846, 1995.
[1.21] M. Kao and J. J. Moskwa, Turbocharged diesel engine modeling for nonlinear engine control and state
estimation, Journal of Dynamic Systems, Measurements, and Control, Vol. 117, pp. 20-30. 1995.
[1.22] M. Kao and J. J. Moskwa, Nonlinear diesel engine control and cylinder pressure, Journal of Dynamic
Systems, Measurements, and Control, Vol. 117, pp. 183-192. 1995.
ME 7236 Module 1 Autumn 2012
1.52
1.7. EXAMPLES
Example 1.1: Water tank filling dynamics model
This example shows how to derive a model of a cylindrical water tank as shown
in Figure 1.15. The scope of the problem is to determine a mathematical model for the
water level in the tank, assuming the entering mass flow rate of water as an independent
variable (control variable). An exit valve is connected to the tank.
in
m
out
m

( ) t h
F
) (t y

Figure 1.15: Cylindrical water tank system, ( ) t m mass of water in the tank, and
corresponding height ( ) t h , F tank-floor area, A exit orifice area
The model of the system can be built following the general modeling guidelines
summarized in this module. In particular, the only relevant "reservoir" is the mass
( ) t m

of water in the tank (state variable), which is proportional to the water height
( ) t h

(output variable). The input for the system is
in
m
, i.e. the inflowing water mass flow,
while the out flowing water
out
m
depends on the exit valve opening area A (control
variable).
Assuming the reservoir effects in the measuring device can be neglected (as
typically very fast), and the water temperature (and therefore its density) changes very
slowly such that it may be assumed constant, it is possible to apply the mass conservation
equation (1) to the mass of water contained in the tank:
( ) ( ) ( ) t m t m t m
dt
d
out in
=

(e-1)
ME 7236 Module 1 Autumn 2012
1.53
where the water mass in the tank can be written as ( ) ( ) t Fh t m = , being F the tank area.
Assuming the incoming flow as the independent variable (control variable), the outgoing
mass flow can be determined considering the exit valve as a flow control device. For this
component, the steady-state continuity equation for an incompressible fluid (Bernoulli's
law) is applied:
( ) ( ) ( )
( )
( ) ( ) t gh t p
t p
t v t v A t m
out

= A
A
= = ,
2
, (e-2)
which results in the following expression:
( ) ( ) t gh A t m
out
2 = (e-3)
Combining (e-1) with (e-3), the model equation for the system in Figure 1.15 is
obtained:
( ) ( ) ( ) | | t gh A t m
F
t h
dt
d
in
2
1

=
(e-4)
which represents a first order nonlinear model.



ME 7236 Module 1 Autumn 2012
1.54
Example 1.2: Compressible flow through an isentropic nozzle
This example is similar to the one discussed in the previous problem and shows
how to utilize the equation describing the isentropic flow of a compressible fluid through
a nozzle to determine the pressure in a vessel filled with gas. The initial pressure
( ) 0 = t p
and temperature
( ) 0 = t T
in
are known, as well as the geometric data of the
vessel and the nozzle. The outlet pressure
out
p
is assumed to be constant and equal to
ambient conditions.

Figure 1.16: Cylindrical, pressurized vessel with exit orifice
Following the general modeling guidelines summarized above, it is possible to
observe that there is no input mass flow (
0 =
in
m
) and the state variable of the system is
the mass
( ) t m
of fluid in the vessel. Applying the continuity equation to the system, the
resulting differential equation is:
( ) ( ) t m t m
dt
d
out
=

(e-5)
The mass flow rate of gas exiting the system can be calculated using the equation
of the flow through an isentropic nozzle (see [1.6]).
ME 7236 Module 1 Autumn 2012
1.55

|
|
.
|

\
|
+
<
|
|
.
|

\
|
|
|
.
|

\
|
+
=
|
|
.
|

\
|
+
>
|
|
.
|

\
|
(
(
(

|
|
.
|

\
|

|
|
.
|

\
|
=

+

1 1
1
1
1 1
1
2
1
2
1
2
1
1
2



in
out
in
in d
out
in
out
in
out
in
out
in
in d
out
p
p
if
RT
Ap C
m
p
p
if
p
p
p
p
RT
Ap C
m



(e-6)
where p is the pressure inside the tank, R is the universal gas constant, T is the
temperature of the fluid in the tank, A is the area of the restriction and is the ratio of the
specific heat coefficients. In (e-6), C
d
is known as discharge coefficient: this model
parameter is typically introduced to account for the real gas flow effects and requires
calibration with experimental data.
Assuming the fluid as a perfect gas and that the temperature of the fluid in the
tank does not change significantly, the ideal gas law can be used in conjunction with (e-
5) to determine the pressure in the vessel:
( )
( )
( ) t m
RT
V t p
dt
d
t m
dt
d
out
=
(

=
(e-7)
Which, combined with (e-6), allows one to estimate the instantaneous pressure in the
vessel.
The expression (e-6) is nonlinear and particularly complex to model. Its
formulation can, however, be considerably simplified if the pressure ratio between the
tank and the external ambient is high (above the critical value) and the tank is sufficiently
large compared to the area of the nozzle. Under these approximations, it is possible to
assume that the nozzle operates in choked conditions. Therefore, by combining equations
(e-6) and (e-7), the differential equation that governs the dynamics of the pressure inside
the vessel can be derived as follows:
p
RT
A C
V
RT
dt
dp
d

(
(
(

|
|
.
|

\
|
+
=

+
1
1
1
2


(e-9)
which is a first order linear differential equation with constant coefficients.
ME 7236 Module 1 Autumn 2012
1.56
Example 1.3: Torsional system dynamics
A simple model of an engine drivetrain is shown in Figure 1.17, where the input
to the system is the torque generated by a multicylinder engine and the desired output is
the engine angular velocity,
du
E
dt
= e
E
. Please note that the damping term B
E is

proportional to the flywheel speed, while the clutch damping is proportional to the
difference between vehicle speed,
du
V
dt
= e
V
, and flywheel speed.
Using the method of determinants, find the transfer function between torque and
engine speed in symbolic form. Expand and group numerator and denominator to obtain
the transfer function in the form of a ratio of polynomials.
crankshaft and
flywheel inertia
vehicle
equivalent inertia
stiffness and damping of
clutch coupling
J
E
J
V
K
C
B
C
T
E
u
E
u
V
B
E
damping due to
bearing friction
at flywheel
T
L

Figure 1.17: Simplified model of engine cranktrain dynamics
The equations of motion can be obtained by applying Newtons law to the system
represented in Figure 1.17:
) ( ) (
V E C E E V E C E E E
B B k T J u u u u u u

=

) ( ) (
V E C V E C L V V
B k T J u u u u u

+ + =

(e-10)
(e-11)
Equations (e-10 e-11) become
( ) ( ) ( )
( ) ( )
L V C C V E C C
E V C C E C E C E
T s K s B s J s K s B
T s K s B s K s B B s J
= + + + +
= + + + +
) ( ) (
) ( ) (
2
2
u u
u u

(e-12)
ME 7236 Module 1 Autumn 2012
1.57
or, in matrix form,
( ) ( ) ( )
( ) ( )
(

=
(

+ + +
+ + + +
L
E
V
E
C C V C C
C C C E C E
T
T
s
s
K s B s J K s B
K s B K s B B s J
) (
) (
2
2
u
u

(e-13)

u
E
(s) =
det
T
E
B
C
s + K
C
( )
T
L
J
V
s
2
+ B
C
s + K
C
( )



(

(
det
J
E
s
2
+ B
C
+ B
E
( )s + K
C ( )
B
C
s + K
C
( )
B
C
s + K
C
( )
J
V
s
2
+ B
C
s + K
C
( )




(

(
(
=
J
V
s
2
+ B
C
s + K
C
( )
T
E
B
C
s + K
C
( )T
L
J
E
s
2
+ B
C
+ B
E
( )s + K
C ( )
J
V
s
2
+ B
C
s + K
C ( )
B
C
s + K
C
( )
2

u
E
(s)
T
E
( s)
T
L
(s)=0
=
J
V
s
2
+ B
C
s + K
C ( )
J
E
s
2
+ B
C
+ B
E
( )
s + K
C ( )
J
V
s
2
+ B
C
s + K
C ( )
B
C
s + K
C
( )
2

Since we desire the transfer function from T
E
to e
E
, we multiply both sides by s
to obtain

su
E
(s)
T
E
(s)
T
L
(s)=0
=
e
E
(s)
T
E
(s)
=
s J
V
s
2
+ B
C
s + K
C
( )
J
E
s
2
+ B
C
+ B
E
( )s + K
C
( )
J
V
s
2
+ B
C
s + K
C
( )
B
C
s + K
C
( )
2
=
J
V
s
3
+ B
C
s
2
+ K
C
s
J
E
J
V
s
4
+ B
C
+B
E
( )
J
V
+J
E
B
C
( )
s
3
+ J
E
K
C
+ B
C
+B
E
( )
B
C
+K
C
J
V
( )
s
2
+K
C
2
B
C
2
s
2
2B
C
K
C
sK
C
2
=
J
V
s
2
+ B
C
s + K
C
J
E
J
V
s
3
+ B
C
+ B
E
( )J
V
+ J
E
B
C
( )s
2
+ J
E
K
C
+ B
E
B
C
+ K
C
J
V
( )s 2B
C
K
C

Now we can substitute numerical values.

Das könnte Ihnen auch gefallen