Model Predictive Control of Modern High-Degree-of-Freedom Turboch
Model Predictive Control of Modern High-Degree-of-Freedom Turboch
TigerPrints
All Dissertations Dissertations
8-2018
Recommended Citation
Koli, Rohit Vishvanath, "Model Predictive Control of Modern High-Degree-of-Freedom Turbocharged Spark Ignited Engines with
External Cooled EGR" (2018). All Dissertations. 2211.
https://tigerprints.clemson.edu/all_dissertations/2211
This Dissertation is brought to you for free and open access by the Dissertations at TigerPrints. It has been accepted for inclusion in All Dissertations by
an authorized administrator of TigerPrints. For more information, please contact [email protected].
MODEL PREDICTIVE CONTROL OF MODERN HIGH-DEGREE-OF-FREEDOM
TURBOCHARGED SPARK IGNITED ENGINES WITH EXTERNAL COOLED EGR
A Dissertation
Presented to
the Graduate School
of Clemson University
In Partial Fulfillment
of the Requirements for the Degree
Doctor of Philosophy
Automotive Engineering
by
Rohit Vishvanath Koli
August 2018
Accepted by:
Dr. Robert Prucka, Committee Chair
Dr. Zoran Filipi
Dr. Qilun Zhu
Dr. Srikanth Pilla
ABSTRACT
Valve Timing actuation. Control of these sub-systems is challenging due to their inter-
dependence and the increased number of actuators associated with engine control. Much
research has been done on developing algorithms which improve the transient
newer technologies like external cooled EGR the control complexity has increased
exponentially.
physical engine model has been developed and analyzed for non-linearity. The
computational burden of implementing this control law has been addressed by utilizing a
semi-physical engine system model and basic analytical differentiation. The resulting
linearization process requires less than 10% of the time required for widely used
Program has been formulated and solved with preliminary validation applied to a 1D
system.
The MPC based control demonstrates the ability to co-ordinate different engine
violation. Avenues for further improvement have been identified and discussed.
ii
DEDICATION
iii
ACKNOWLEDGMENTS
guidance and support. He is among the few academic advisors who proactively prioritize
the students’ well-being. He has taught me that the easiest person to convince is the
person in the mirror, which has inspired me to cultivate and maintain the rigor of
fundamental research.
I would also like to thank my friend Dr. Qilun Zhu for his continual input and the
several hour long discussions we have had regarding engine control, science and life. His
work has been a great source of inspiration for my research. Thank you for having the
patience and the will to help me to the best of your ability I would also like to thank Dr.
Zoran Filipi for his valuable advice on turbocharged engine control on modeling. His
class lecturers on advanced IC engine controls was a memorable experience for me.
The research work performed was in collaboration with industrial partners Robert
Bosch LLC and Fiat Chrysler Automobiles. I would like to thank Jason Schwanke and
Dr. Shyam Jade from Robert Bosch LLC for their valuable industrial insight and quick
assistance in this research. I would also like to thank Michael Prucka and the folks at Fiat
Lastly, I would like to thank all my fellow CUICAR students and friends,
Konstantinos Siokos, Shuonan Xu, Harikesh Arunachalam, Anirudh Allam, Dan Egan,
Dennis Robertson, Tommy Powell, Ryan O’Donnell, and others for their help and
assistance.
iv
TABLE OF CONTENTS
Page
ABSTRACT ..................................................................................................................... ii
DEDICATION ................................................................................................................iii
ACKNOWLEDGMENTS .............................................................................................. iv
CHAPTER
I. INTRODUCTION ......................................................................................... 1
v
Table of Contents (Continued) Page
vi
Table of Contents (Continued) Page
vii
Table of Contents (Continued) Page
viii
LIST OF TABLES
Table Page
Table II-1. Engine specifications ...................................................................................... 14
Table III-1. Error metrics of the compressor mass flow rate models ............................... 25
Table III-2. Actuator combinations used for data collection for ANN model training .... 40
Table III-3 Inputs used for training ANNs ...................................................................... 42
Table III-4. Accuracy of trained ANNs ............................................................................ 48
Table III-5 Units used for normalization of state space model ......................................... 49
Table III-6 Error metrics of the 0D model for four different transient engine trajectories.
........................................................................................................................................... 51
Table IV-1 Minimum number of model linearizations required in horizon to maintain
IMEPn model accuracy for varying input frequency ........................................................ 57
Table IV-2 Minimum number of model linearizations required in horizon to maintain
IMEPn model accuracy for varying input frequency ........................................................ 59
Table A-1 Conventional PID control vs Smith Predictor Control performance comparison
......................................................................................................................................... 129
TABLE A-2. Performance summary of EGR % control using NMPC .......................... 150
ix
LIST OF FIGURES
Figure Page
x
Figure III-16. Modeled vs measured steady state exhaust temperature for all data +/- 10%
error bands represented by dashed red line. The accuracy of each of these neural
networks is summarized in................................................................................................ 47
Figure III-17 Open-loop system trajectory for 0D model vs experimentally collected data
........................................................................................................................................... 52
Figure IV-1 Deviation between the non-linear system model and the linearized system
model as a function of time step N from the linearization. The deviation is induced by
perturbing 𝑚𝐼𝑇𝑉 corresponding to commanded 𝐼𝑀𝐸𝑃𝑛 ................................................. 54
Figure IV-2 Model error as a function of commanded open-loop change in IMEPn
shows that the deviation between linearized and non-linear model varies as a function of
operating condition and the time –step after the point of linearization ............................ 57
Figure IV-3 Model error as a function of frequency of commanded open-loop model
input shows that the deviation between linearized and non-linear model generally peaks
between 1.5 Hz and 0.5 Hz depending on operating condition ..................................... 60
Figure V-1 Approximate relationship between horizon length considered for MPC, size
of the model and the computational expense of real-time execution................................ 62
Figure V-2 The selection of MPC sample time and prediction horizon time dictates the
choice of appropriate engine dynamics in the engine model ............................................ 63
Figure V-3. Graphical layout of Artificial Neural Network ............................................. 70
Figure V-4. Comparison between execution times of 𝐼𝑀𝐸𝑃𝑛 ANN model linearization
using numerical and analytical approach .......................................................................... 73
Figure V-5. Comparison between execution times of complete system model linearization
using numerical and analytical approach .......................................................................... 74
Figure V-6. Co-simulation setup between GT-Power and Simulink ................................ 79
Figure V-7 IMEPn response of MPC co-simulated with GT-Power ................................ 80
Figure V-8 Control actuator trajectories for MPC cosimulation with GT-Power, air-path
actuators are on the left hand side and cylinder actuators are on the right hand side. Black
dashed lines are the steady-state reference inputs and blue lines are the inputs generated
by MPC ............................................................................................................................. 81
Figure V-9 Compressor operation during co-simulation of GT-Power with MPC .......... 83
Figure VI-1. The gradient of the flow function becomes very steep as the pressure
difference across the valve diminishes ............................................................................. 85
Figure VI-2.Real-time estimated vs measured turbocharger speed using fixed point
iteration method ................................................................................................................ 88
Figure VII-1. Rapid control prototyping setup for experimental validation of MPC ....... 90
Figure VII-2. Tip-in IMEPn response for MPC and Open-loop control strategy............. 91
Figure VII-3. Intake throttle mass flow rate trajectories for MPC and Open-loop control
........................................................................................................................................... 92
Figure VII-4. Wastegate mass flow rate trajectories for MPC and Open-loop control .... 92
Figure VII-5. External cooled EGR mass flow rate trajectories for MPC and Open-loop
control shows that MPC deliberately stops EGR flow on the aggressive tip-in ............... 93
Figure VII-6. Spark timing trajectories for MPC and Open-loop control ........................ 94
Figure VII-7. Camshaft position timing trajectories for MPC and Open-loop control .... 94
xi
Figure VII-8 ANN Modeled volumetric efficiency using MPC inputs and open-loop
inputs for tip-in maneuver ................................................................................................ 95
Figure VII-9 Measured turbine inlet temperature for MPC and Open-loop control shows
that MPC commands late spark timing to increase specific exhaust gas enthalpy for
turbocharger response ....................................................................................................... 95
Figure VII-10 Turbocharger speed for MPC increases much more rapidly than open-loop
control ............................................................................................................................... 96
Figure VII-11 Measured 𝐾𝐼2 for MPC does not exceed the limit of 1 despite advanced
spark timing during tip-in ................................................................................................. 97
Figure VII-12. Tip-out IMEPn response for MPC and Open-loop control strategy......... 98
Figure VII-13 Air-path trajetories for aggressive tip-out maneuver ............................... 100
Figure VII-14 Cylinder actuator trajectories for aggressive tip-out maneuver............... 100
Figure VII-15 Compressor operation during the aggressive tip-out ............................... 101
Figure A-1 Control system setup .................................................................................... 110
Figure A-2 Dimensional Discharge coefficient map obtained from optimization routine
......................................................................................................................................... 113
Figure A-3 Measured vs modeled EGR valve mass flow rate ........................................ 114
Figure A-4 Modeled low pressure EGR path ................................................................. 115
Figure A-5. Closed loop EGR fraction controller with Smith predictor......................... 118
Figure A-6 Modeled vs measured transport delay .......................................................... 120
Figure A-7. Comparison between modeled and measured EGR fraction at EGR sensor
location; engine operated at 1700 RPM 3 Bar BMEP .................................................... 121
Figure A-8. Comparison between modeled and measured EGR fraction at EGR sensor
location; engine operated at 2100 RPM 5 Bar BMEP .................................................... 122
Figure A-9. Comparison between modeled and measured EGR fraction at EGR sensor
location; engine operated at 2100 RPM 2.5 Bar BMEP ................................................. 123
Figure A-10. Step response of EGR fraction controller with conventional PID controller
at 1500RPM 7 Bar BMEP .............................................................................................. 124
Figure A-11. Step response of EGR fraction controller with Smith Predictor at 1500RPM
7 Bar BMEP .................................................................................................................... 125
Figure A-12. Step response of EGR fraction controller with conventional PID controller
at 1300RPM 4 Bar BMEP .............................................................................................. 126
Figure A-13. Step response of EGR fraction controller with Smith Predictor ............... 127
Figure A-14. Step response of EGR fraction controller with conventional PID controller
at 2000RPM 9 Bar BMEP .............................................................................................. 128
Figure A-15. Step response of EGR fraction controller with Smith Predictor at 2000RPM
9 Bar BMEP .................................................................................................................... 128
Figure A-16. Schematic of Engine and Air path model.................................................. 133
Figure A-17. Step response of oxygen sensor based EGR measurement installed at
compressor outlet location .............................................................................................. 135
Figure A-18. Comparison of step response of transport delay models with true dead time
delay response ................................................................................................................. 136
Figure A-19 Closed-loop control structure of DL-EGR NMPC..................................... 139
xii
Figure A-20 Pressure states and intake manifold EGR dilution trajectories for HP-EGR
control only. .................................................................................................................... 141
Figure A-21 Input trajectories for HP-EGR control only. .............................................. 142
Figure A-22 Pressure states and intake manifold EGR dilution trajectories for LP-EGR
control only. .................................................................................................................... 144
Figure A-23 Input trajectories for LP-EGR control only................................................ 145
Figure A-24 Pressure states and intake manifold EGR dilution trajectories for Dual-loop
EGR control. ................................................................................................................... 147
Figure A-25 Input trajectories for Dual-loop EGR control. ........................................... 148
Figure A-26 Transient intake manifold EGR dilution for multiple EGR architectures .. 149
xiii
I. INTRODUCTION
requirement of a fleet average of 54.5mpg for the model year 2025 requires radical
improvements in fuel economy [4]. This can be facilitated partially by further downsizing
engine, there are undesired transient phenomenon which severely affect performance and
consequently driver perception during these situations. This phenomenon is called the
‘turbo-lag’ and is literally a time delay between the driver-demanded engine torque and
the response of the engine. Dedicated systems have been widely researched and utilized
Lighter turbocharger rotor assemblies have also been investigated by [6][7] which
Variable Geometry Turbines (VGT) has also shown significant improvement in transient
has not only shown improvement in transient response but also improved low speed
1
boosting ability [23]-[26]. However, exhaust gas temperature, cost, and reliability
[26][28] for improved turbocharger response and exhaust heat recovery. All of these
elevated cost and added system complexity which is unfavorable for mass production.
Modern engines are highly over-actuated with respect to torque control, wherein
multiple actuator combinations exist for a single engine torque level at a given engine
2
speed. Usually a unique combination exists which optimizes the steady-state fuel
economy of the engine. However, multiple optimum combinations with negligible fuel
economy difference may exist [11][12]. The actuator combinations themselves can be
very different from each other which further confounds the determination of the
The dynamics of the different subsystems on the engine are also vastly different
in timescales. The impact of each actuator on its corresponding subsystem and coupling
with other subsystems affects the overall engine’s dynamic response drastically. Because
of the fluid coupling control of the turbocharger actuators is critical for engine
performance and fuel economy. Early turbocharger control methods have been outlined
[14] with introduction to rule based electronic control of waste-gate and blow-off valve.
One of the most significant challenges associated with turbocharged engine control has
been the coordination of the waste-gate and the intake throttle valve in turbocharged SI
engines. The primary load control actuator for SI engines is the intake throttle valve.
However, in order to increase the intake manifold pressure beyond atmospheric pressure,
the turbocharger is utilized. Exhaust enthalpy drop across the turbine results in power
transfer to the compressor. The compressor performs work on the inlet air, thereby
increasing it’s temperature and pressure, also known as ‘boost’ pressure. The waste-gate
valve, which controls the turbine bypass flow has been typically used to limit the
However, with the advent of electronic waste-gate control this valve is kept wide open
under throttled operation to minimize the exhaust pressure and hence minimize pumping
3
work performed by the engine. However, under transient operation, the control of the
Figure I-2. Load control using intake throttle and wastegate on downsized turbocharged
engines
Much research has been conducted in the control and coordination of the
turbocharger and air-path actuators since then with an exhaustive list of papers and
patents [13][17]. Karnik, et. al have evaluated the performance of decentralized Single
Input Single Output (SISO) controllers vs full-state feedback and the reduced versions of
full state feedback control in [16] using linearized engine models. The results of this
laws for waste-gate and intake throttle coordination. Typically each of the subsystems
4
have been controlled with individual control loops targeting a particular set-point as
outlined in [13]. Non-linear control approaches have been introduced in [17] using
first order non-linear model. The feed-forward control law is derived based on inversion
of the reduced order model, which is then augmented with an error-integral based
feedback term. The potential of Variable Valve Timing (VVT) to improve turbocharger
response has been mentioned in [59][60] by improving air-flow into the cylinders and in
some conditions directly from the intake port to the exhaust port during the valve overlap
phase. Colin et al. [57][58] have proposed a decentralized control system which consists
of feed-forward intake throttle controller and NMPC based waste-gate controller for
engine torque control. The control of Variable Valve Timing (VVT) into the engine
torque controller for control of the trapped Residual Gas Fraction (RGF) has been
introduced using a static neural network model. Interaction between the turbocharger
response and VVT has been discussed but not exploited in the control strategy due to the
decentralized structure of the controller which targets RGF and Torque independently.
Spark timing also has a significant impact on the combustion process, exhaust
temperature and the cycle to cycle variability of combustion [54][71]. For best fuel-
economy, the spark-timing is calibrated to position fifty percent of the in cylinder fuel–air
mixture burnt at eight crank angle degrees after top dead center. However, on
turbocharged SI engines the spark timing at higher loads is often limited by the onset of
knocking in the cylinder due to end-gas auto-ignition [65] [68]. Such abnormal
5
combustion causes premature damage to the engine. The Spark timing is retarded to a
safe level such that knocking is minimized to acceptable levels. This is called the knock-
specific enthalpy of the exhaust gas have been shown to reduce turbocharger lag in
[72]and [9]. There is significant potential to improve transient response by utilizing this
actuator. However, due to the knock and combustion stability constraints mentioned
above, simultaneous control of spark timing along with other engine actuators is
challenging.
Single Output (SISO) type controllers with lookup-tables and maps. Although this
approach is simple to adopt for OEMs, it requires significant calibration effort and tuning
to ensure stability and robustness. This entails significant real-time testing on the
dynamometer and in some cases the vehicle as well. The recent findings of implementing
Exhaust Gas Recirculation (EGR) in SI turbocharged engines has also shown potential to
improve fuel economy drastically [11][12][29]-[36]. However, this adds yet another
control actuator which needs to be considered in the overall engine control system.
Moreover, low-pressure cooled EGR also has known issues associated with the inability
of feed-forward flow modeling through the valve and transport delay of the EGR Air
mixture through the intake air-path [37]-[41]. This delay is more significant for SI
engines compared to CI engines because the combustion process of SI engines is far more
sensitive to external dilution, since external EGR has a significant effect on the
6
Because the number of ‘control knobs’ on engines is ever increasing OEMs are
beginning to realize the potential of using model based control methodologies which
could significantly reduce the required calibration and tuning efforts. Multi-Input-Multi-
Output (MIMO) control systems have shown significant promise in being capable of
coordinating multiple engine actuators based on multiple targeted outputs. One particular
MIMO control approach which has gained attention in recent years is Model Predictive
Control (MPC).
actions for the system are generated as a result of an optimization problem over a fixed
time period into the future from the current time. This period is called the prediction
horizon. As the closed loop control system executes and proceeds in time, the horizon
recedes further equally. MPC is hence also known as Receding Horizon Control (RHC)
current condition and a dynamic control oriented model of the system. The general MPC
1. Generate a predicted future reference trajectory of the system for the prediction
horizon based on a known model, current state of the system and reference inputs.
function related to the behavior of the system for the prediction horizon.
3. Apply the inputs that correspond to the current time to the system
7
4. Re-evaluate the state of the system at periodic intervals and repeat the process
It is clearly visible that this strategy automatically considers the slow and the fast
behavior for a specific time period in the future. The main advantage of applying MPC to
engines is the significant reduction of calibration and tuning efforts required. MPC tuning
can be achieved easily by simple manipulation of the constant terms within the objective
function. This is greatly simplified compared to the conventional tuning and calibration
approach used for conventional SISO and LUT based methods. Another major advantage
of using MPC is the ability to handle constraints. These constraints can be simple
saturation limit constraints for the control actions or more complex operational
constraints for turbocharger protection, combustion stability and knocking which are
extremely detrimental to the mechanical integrity, fuel economy and exhaust emissions
powerful microcontroller with ample memory is required. In order to justify the usage of
MPC for engine control in production, the MPC has to be capable of operating at any
engine operating condition and any kind of engine transient, with consideration of fuel
8
economy, emissions, hardware protection, drivability and most importantly torque
delivery. The dynamic engine model complexity and accuracy begins to become a
order models to capture high levels of accuracy and mimic relevant dynamic behavior.
phase dynamics which complicates the model further. The turbocharger’s inertial
dynamics have a time constant which is in the order of seconds. This is significantly
slower compared to the cycle by cycle dynamics of the engine cylinders. Incorporating all
of these dynamics into one prediction model and consideration of a time period long
enough to account for the turbocharger’s inertial dynamics results in MPC based
A linear prediction model is the favorable choice for MPC formulation due to it’s
the computational cost of obtaining a linearized model from a more accurate non-linear
system model in real-time. The prediction model utilized in many of the past research
articles is a local linear approximation of the engine dynamics [73][75]. Many of these
research articles focus mainly on the control of the air-path of the diesel engines using
micro-controller. Some of these articles utilize black-box models identified directly from
9
experimental or high-fidelity simulation data of the engine. The number of models used
to approximate the engine dynamics is also an open question since utilizing lesser models
results in larger inaccuracy in the prediction of system behavior during transients and
complicates the transitioning between linear models. These approaches have been
demonstrated to work well mainly because of the linear-like behavior of the air-path
dynamics (intake manifold and compressor outlet pressure). However, this approach has
not been demonstrated to track highly non-linear outputs associated with engine behavior
like engine torque while simultaneously adhering to combustion constraints which are
also complicated non-linear functions of the system states, inputs and parameters.
et al in [77] for evacuation of EGR trapped in the intake manifold using a Linear Time
Varying MPC. This approach demonstrates the ability of the MPC to schedule the VVT
actuation in order to short-circuit the flow from intake ports directly to the exhaust port
using available preview of the future torque reference. Bemporad et al have explored the
ability of utilizing this short-circuit flow in order to improve turbocharger boost response
in [82] using a linear MPC with pre-stored linear black-box models to control the air-path
actuators. Similar to the diesel MPC research, the air-path dynamics of the SI
turbocharged engine are also linear-like, and hence the local-linear formulation has been
shown to work.
actuator and combustion constraints has been recently investigated in [79]-[81],[87] and
[54]. Combustion variability and knock constraint have been considered in the
10
formulation. Sequential Quadratic Programming based approach is used to solve the
MPC iteratively. The same approach has been further investigated in [83] for Air-path
physical and the cylinder model is purely data driven. The goal is to unify the control of
air-path and engine cylinder actuators into a single control system which automatically
respecting the engine’s operational constraints. In order to achieve this goal many sub-
1. Develop a control oriented model of the engine which captures the necessary and
2. Develop an algorithm that utilizes the control oriented model and formulates an
controller
11
5. Build a Simulation framework to tune the MPC controller by coupling it with a
12
II. EXPERIMENTAL METHODS AND RESEARCH ENGINE SETUP
actuators. Additionally a low-pressure, cooled EGR circuit was added to the stock engine
13
Table II-1. Engine specifications
Displacement 2.0L
Bore 86
Stroke 86
An aftermarket stainless-steel EGR cooler for a Ford F250 is utilized as the LP-
EGR system. The EGR valve from the Audi 1.8TDI is installed as the LP-EGR flow
control valve. Engine coolant is circulated through the LP-EGR cooler to emulate real-
world cooled EGR systems used in production passenger car diesel engines. Excessive
EGR cooling has also been known to cause condensation in the compressor inlet location
compressor throttle is also installed upstream of the EGR mixing location. The original
pneumatic waste-gate actuator has been replaced with an electronic linear actuator for
improved position control of the waste-gate valve. Figure II-3, Figure II-2 show the hot
side as well as the cold side of the engine. The actuators and systems most relevant to
14
Figure II-2. Exhaust side of the test engine
15
2.2 Data acquisition and dynamometer interface
High and low speed data acquisition systems have been utilized on the test engine. The
high speed data acquisition consists of the crank angle resolved pressure measurement of
the cylinders, intake and exhaust ports and across the low pressure EGR valve. The low
speed data acquisition systems mainly consists of time averaged pressure, temperature
and mass flow rate measurements at various locations on the engine. The test engine
FEV’s Test automation system is utilized to control the speed and torque setpoint of the
dynamometer.
16
2.3 Rapid control prototyping hardware
Since the base engine is a production engine, it has production intent sensors and
actuators associated with basic engine functionality. An open Bosch ECU MED17 has
been interfaced with the production engine sensors and actuators. This ECU is also
capable of real time cylinder pressure feedback based diagnostics. The embedded
application software deployed onto the ECU has CCP bypass functionality which allows
run-time read-access to all of the ECU variables and write-access to the variables
17
Additional subsystem like continuous blow off valve, electronic wastegate control
actuator and pre-compressor throttle are controlled using the ETAS ES930 and ES910
units. The ES910 is a rapid control prototyping ECU and calibration interface for the
Simulink/Stateflow and compiled using the Intecrio Realtime system target file. This
target file compiles the base level Simulink/Stateflow model into a Scoop IX file with the
extension .six. This file is then imported into a software called ETAS Intecrio which
facilitates interfacing with the MED17 ECU, ES930 and other systems via CAN. The
execution rate and several other diagnostics can also be configured in Intecrio. The
algorithms with large memory requirement and/or small execution time step. These
Estimation and complex plant-models for Hardware In the Loop simulation. For such
cases the Dspace DS1006 is utilized. The DS1006 system used has four processor boards
which can execute different applications independently. This system has high speed CAN
controllers which can be interfaced with other hardware. One of these controllers is
utilized to establish communication with the ES910, while the other is utilized to
communicate with the dynamometer control system for vehicle level simulations with
18
Figure II-6. Dspace rapid prototyping system
19
III. CONTROL ORIENTED ENGINE MODELING
A modular approach has been adopted for modeling, wherein the individual
components of the air path have been modeled separately and ultimately combined
Selection of the turbocharger model is critical for model accuracy and execution
speed of the engine model. The turbocharger models are split into the compressor models
To model the pressure dynamics in the boost manifold the compressor mass flow
rate is modeled as a function of turbocharger speed and the pressure ratio across the
compressor. There are several models available in literature with varying degrees of
accuracy and targeted applications. Artificial Neural Network (ANN) based models have
been proposed in [47]. Although the prediction of these models within the compressor-
map is good, low speed extrapolation is an issue with this approach. Furthermore, the
structure of the ANN required to ensure in-map and low-speed extrapolation accuracy
physics based classical approach was chosen to model the compressor mass flow rate.
Two models were evaluated for the compressor used in this research. The first one is the
ellipse model [44][46] The ellipse model is based on the ellipsoidal parametrization of
the iso-speed lines obtained from steady state gas-stand tests of the compressor. These
20
results of these tests are called the compressor maps and are provided by the
manufacturer for normal operating region of the compressor. The ellipse model supports
compressor mass flow rate prediction in the mild and deep surge region as well as the
choke region. However, the disadvantage of this approach is that the model changes
mathematical structure when transitioning from the normal region to the surge region of
the compressor map. Incorporating a variable structure model into MPC based control
algorithms is not favorable. Hence only the normal region of the compressor map was
modeled using the Ellipse model. The compressor map is modeled in the normal region
as follows.
1
Π 𝐶2 𝐶1
𝑚̇𝐶𝑜𝑚𝑝 = (𝑚̇𝑀𝑎𝑥 − 𝑚̇𝑍𝑆𝐿 ) [1 − (Π ) ] + 𝑚̇𝑍𝑆𝐿 1
ZSL
𝐶
𝐶2 = 𝑓𝐶 2 (𝜔 𝑇 ) = 𝐶2,0 + 𝐶2,1 𝜔𝑇2,2 3
𝐶
ΠZSL = 𝑓𝑍𝑆𝐿 (𝜔 𝑇 ) = 1 + 𝐶5,1 𝜔𝑇5,2 6
Where, 𝑚̇ is the mass flow rate and 𝐶𝑖,𝑗 have real values. Π is the pressure ratio
21
squares optimization to minimize the difference between the manufacturer’s and modeled
iso-speed lines. The suffixes 𝑍𝑆𝐿 and 𝑀𝑎𝑥 correspond to two distinct locations on the
iso-speed line. The former corresponds to the point on the surge line and the latter
evolution of the Jensen and Kristensen empirical mean value model [51]. The latter has
been widely used in publications relating to modeling, control and estimation of air path
nonlinear relationship between the compressor mass flow rate and the pressure ratio using
𝑚̇𝐶𝑜𝑚𝑝 𝑅𝑇𝐶𝑖𝑛
𝜙= (𝑃𝐶𝑖𝑛 𝑑𝑐3 𝜔𝑇 )
8
Where, 𝐶𝑃𝐶𝑖𝑛 and 𝑇𝐶𝑖𝑛 are the constant pressure specific heat capacity and
temperature of the gas at the inlet of the compressor. 𝑅 and 𝐾 are the specific gas
constant and the ratio of specific heat capacity respectively. 𝑈𝐶 is the compressor blade
tip speed which is calculated as a function of the turbocharger speed 𝜔 𝑇 and the
22
The flow parameter 𝜙 is not directly measureable and in fact it is the parameter
which needs to be predicted in order to derive the compressor mass flow rate. Several
functions have been explored to derive 𝜙 from 𝜓 depending on the shape of the iso-speed
lines of the compressor [52][53]. The most widely used relationship is the inverse
proportionality function which uses the Mach number across the ring orifice of the
derived from the turbo speed dependent Mach number. Preserving the structure of the
[49][50].
𝐶(𝜔𝑇 )−𝐴(𝜔𝑇 )𝜓
𝜙= 10
𝐵(𝜔𝑇 )+𝜓
turbocharger speed. The coefficients of these polynomial functions are obtained similar to
the method used for obtaining 𝐶1 and 𝐶2 for the ellipse model.
23
Figure III-1. Polynomial approximation for Jensen compressor model coefficients
The modeled iso-speed lines of both of the models are shown in Figure III-3. The
accuracy metrics for predicted mass flow rate as a function of compressor pressure ratio
for both the models is shown in Table III-1. Error metrics of the compressor mass flow
rate models. Although the Ellipse model accuracy is superior to the Jensen model, the
without the prior knowledge of the mass flow rate across it. The Jensen model structure
24
can be utilized all the way till the zero compressor mass flow rate. However, the true
compressor behavior in the deep surge region is not an accurate representation of the
compressor behavior predicted by the Jensen model. Since the scope of this research is to
not model these abnormal modes of operation, the Jensen model is used since it
transitions continuously from the normal operation region to the surge region
Figure III-2. Modeled compressor iso-speed lines vs manufacturer provided map points
Table III-1. Error metrics of the compressor mass flow rate models
25
The compressor model is a quasi-steady state model [49]. During highly transient engine
operation Π can be calculated to have a value which is much higher than the maximum
possible pressure ratio that the compressor model can deliver at the corresponding
turbocharger speed and zero mass flow rate. This leads to an erroneous computation of 𝜓
and hence, to limit the operation of the compressor model to have realistic 𝜙 values even
𝐶(𝜔 )
𝜓𝑀𝑎𝑥 = 𝐴(𝜔𝑇 ) 11
𝑇
𝜓𝐹𝑖𝑛𝑎𝑙 = min(𝜓𝑀𝑎𝑥 , 𝜓) 12
The compressor efficiency can vary significantly depending upon the operating
significant impact with regards to turbocharger inertial dynamics since the turbocharger
speed dynamics are directly coupled with this parameter. The compressor isentropic
𝑃
𝜂𝐶𝑜𝑚𝑝 = 𝑃 𝐶𝐼𝑑𝑒𝑎𝑙 13
𝐶𝐴𝑐𝑡𝑢𝑎𝑙
Where, 𝑃𝐶𝐼𝑑𝑒𝑎𝑙 is the power required to compress the gas from inlet to the outlet
pressure, defined by the isentropic process. 𝑃𝐶𝐴𝑐𝑡𝑢𝑎𝑙 is the actual power required by the
compressor to pressurize the gas. 𝑃𝐶𝐴𝑐𝑡𝑢𝑎𝑙 is always lesser than 𝑃𝐶𝐼𝑑𝑒𝑎𝑙 due to the heat
26
transfer and other losses associated with the pumping of inlet gas. Since 𝑃𝐶𝐴𝑐𝑡𝑢𝑎𝑙 is not
measured on the engine, determination of 𝜂𝐶𝑜𝑚𝑝 cannot be made directly from this
equation. [53] has shown that it can be identified as a function of 𝜙. However, 𝜙 is the
𝜂𝐶𝑜𝑚𝑝 = 𝑏1 𝜓3 + 𝑏2 𝜓 2 + 𝑏3 𝜓 + 𝑏4 14
Where, 𝑏1 , 𝑏2 , 𝑏3 and 𝑏4 are identified to minimize the error between the manufacturer’s
data and the modeled 𝜂𝐶𝑜𝑚𝑝 . Figure III-3 shows the fit between the modeled and the
efficiency curve is within 5% of the manufacturer’s data with some points exceeding the
27
Figure III-3. Modeled compressor efficiency curve vs manufacturer provided compressor
map points
compressor efficiency and the pressure ratio across the compressor as follows.
𝐾−1
𝑃 𝐾
( 𝐵𝑠𝑡 ) −1
𝑃𝐶𝑖𝑛
𝑇𝐶𝑜𝑚𝑝𝑂𝑢𝑡 = 𝑇𝐶𝑖𝑛 ( + 1) 15
𝜂𝐶𝑜𝑚𝑝
Where, 𝑇𝐶𝑖𝑛 is the compressor inlet temperature calculated using the adiabatic mixing
equation as follows.
28
𝑚̇𝐸𝑔𝑟 𝑇𝐸𝑔𝑟 +𝑚̇𝐴𝑖𝑟 𝑇𝐴𝑖𝑟
𝑇𝐶𝑖𝑛 = 16
𝑚̇𝐸𝑔𝑟 +𝑚̇𝐴𝑖𝑟
Where, 𝑚̇𝐸𝑔𝑟 is assumed to be modeled and 𝑚̇𝐴𝑖𝑟 is directly measured from the
engine air mass flow rate sensor which is a few inches upstream of the EGR mixing
inlet and outlet temperature models is shown in Figure III-4 and Figure III-5
29
Figure III-5 0D Modeled vs GT-Power modeled compressor outlet temperature
Two turbine mass flow rate models have been developed. One model which
predicts the mass flow rate across the turbine as a function of the pressure ratio across it
and another model which predicts the pressure ratio as a function of known mass flow
rate across the turbine. The turbine mass flow rate is represented as a reduced Turbine
30
𝑚̇𝑇𝑢𝑟𝑏 √𝑇𝑇𝑖𝑛 105
𝑇𝐹𝑃 = 17
𝑃𝑇𝑖𝑛
conversion to mass flow rate is required in order to fit a model to the data. The model
used to predict turbine mass flow rate as a function of pressure ratio is a modified version
of the standard restriction model described by Eriksson in [10] has been utilized to model
the relationship between the pressure ratio and mass flow rate across the turbine. The
structure of this model is very similar to the simplified orifice flow model and is defined
as follows.
𝐴𝑇𝑢𝑟𝑏 .P𝑇𝑖𝑛 𝑃
𝑚̇ 𝑇𝑢𝑟𝑏 = . ψPR (√𝑃 𝑇𝑖𝑛 ) 18
√𝑅. 𝑇𝑇𝑖𝑛 𝑇𝑜𝑢𝑡
𝐴𝑇𝑢𝑟𝑏 = 𝑗1 𝜔𝑇 + 𝑗2 20
The parameters 𝑗1 and 𝑗2 are identified using the manufacturer’s data. The main
difference between this flow function and the one used in the orifice model is that ψPR is
applied to the square root of the pressure ratio across the turbine as mentioned in [53].
The flow function of the standard orifice model changes its structure when the pressure
31
ratio reduces below the critical pressure ratio. This causes the standard orifice model to
operate under the choked flow regime where the flow function ψPR loses coupling with
the turbine outlet pressure. Utilizing the square root of the pressure ratio instead of the
pressure ratio itself, significantly improves the accuracy of the restriction based model.
Figure III-6 shows the modeled and the manufacturer’s turbine TFP points.
Figure III-6. Modeled turbine mass flow rate vs manufacturer provided turbine map
points
Due to it’s non-linear nature and the square root terms present in the model, the
inversion of this restriction based model is very difficult. Due to this concern, a simple 3rd
order polynomial model is defined which predicts turbine pressure ratio as a function of
32
𝑃𝑇𝑖𝑛
𝑃𝑇𝑜𝑢𝑡
= 𝛼1 (𝑚̇ 𝑇𝑢𝑟𝑏 )2 + 𝛼2 (𝑚̇ 𝑇𝑢𝑟𝑏 ) + 𝛼3 21
Where, 𝛼3 is a constant with value 1. The rest of the 𝛼𝑖 terms are not constsant since
there is a dependency on turbine inlet temperature and the mass flow rate. Based on the
turbine map, the 𝛼𝑖 terms were identified to be linear and second order polynomial terms
𝛼1 = 𝛽1 𝑇𝑇𝑖𝑛 + 𝛽2 22
2
𝛼2 = 𝛽3 𝑇𝑇𝑖𝑛 + 𝛽4 𝑇𝑇𝑖𝑛 + 𝛽5 23
Where, the 𝛽𝑖 terms are constants identified by minimizing the error between the model
and the manufacturer data. The fit of the model vs the manufacturer’s data is shown for
33
Figure III-7. Turbine pressure ratio as a function of turbine mass flow rate with map
34
3.1.4 Turbine efficiency model
Where, 𝑐𝑝𝐸𝑥ℎ and 𝐾𝐸𝑥ℎ are the constant pressure specific heat capacity and the
ratio of specific heat capacities of the exhaust gas, modeled as polynomial functions of
the compressor and turbine models. The turbine efficiency is shown in Figure III-8 as a
function of BSR. The model is accurate within five percent for majority of the turbine
map points except for the six points at the lowest turbo speed.
35
Figure III-8. Modeled turbine efficiency curve vs map points
The compressor and turbine torque models are given by the following equations [42]
𝐾−1
𝑃𝐵𝑠𝑡 𝐾 1
𝜏𝐶𝑜𝑚𝑝 = 𝑚̇𝐶𝑜𝑚𝑝 . 𝑐𝑝𝐴 . 𝑇𝐶𝑖𝑛 [(𝑃 ) − 1] . [𝜂 ] 26
𝐶𝑖𝑛 𝐶𝑜𝑚𝑝 .𝜔𝑇
𝐾𝐸𝑥ℎ −1
𝑃𝑇𝑖𝑛 𝐾𝐸𝑥ℎ 𝜂𝑇𝑢𝑟𝑏
𝜏 𝑇𝑢𝑟𝑏 = 𝑚̇ 𝑇𝑢𝑟𝑏 . 𝑐𝑝𝐸𝑥ℎ . 𝑇𝐸𝑥ℎ [1 − (𝑃 ) ].[ ] 27
𝑇𝑜𝑢𝑡 𝜔𝑇
These equations are used in the turbocharger rotational dynamics equation as follows
36
𝑑𝜔𝑇 1
= 𝐽 (𝜏 𝑇𝑢𝑟𝑏 − 𝜏𝐶𝑜𝑚𝑝 − 𝜏𝐹𝑟𝑖𝑐𝑡 ) 28
𝑑𝑡 𝑇
Where, 𝜏𝐹𝑟𝑖𝑐𝑡 is the friction torque which is assumed to be zero, since this
parameter is associated with the mechanical efficiency term which is lumped into the
turbine efficiency as mentioned in [53]. 𝐽𝑇 is the moment of inertia of the rotor of the
Alternatively it can be modeled if the exact geometry and material composition of the
turbocharger rotor assembly is available. These parameters can be used on the part file in
any commercial computer aided design software to estimate the moment of inertia of the
entire rotor assembly. For this research, this parameter has been derived experimentally
by comparing the turbocharger speed behavior of the 1D Engine simulation model to the
experimental data for a transient engine response. The moment of inertia utilized in the
1D Model is adjusted till the 1D Engine model mimics the turbo speed behavior of the
real engine. The final value was determined to be 1.7 × 10−5 𝑘𝑔/𝑚2 which is within
the range of values for similar sized turbochargers. The turbocharger speed response for
the tuned 𝐽𝑇 can be seen in Figure III-9 and Figure III-10. The time constants of the 1D-
37
Figure III-9. Comparison of measured and GT-Power modeled transient turbocharger
38
3.2 Data driven parameter models for engine cylinders
active research area since the advent of internal combustion engines. Parametric models
based on physics, empirical relations and measured data have been widely developed and
investigated. Since, engine modeling is not the focus of this research and the means to
obtain measurement data is easily available, a data driven approach has been adopted to
model these processes. The data was obtained by operating the engine at multiple
actuator combinations at 2000 and 3000 RPM. Each combination was allowed to settle
for some time before transitioning to the next combination. This time period is critical
because lesser time results in some of the slower dynamic behavior like exhaust
temperature to be transient even at the point of transition to the next combination. More
time per operating point is desirable but would lead to increased time required to obtain
the entire data set. A time period of five seconds was chosen per operating combination
which resulted in a total of forty hours of dynamometer test time. The actuator
combinations are shown in Table III-2. The actuator combinations are programmed into
an automated test cycle and deployed onto the ES910’s real-time controller.
In the control oriented model of the engine, there are four models that correspond
to the gas-exchange and in-cylinder combustion process of the engine. The gas-exchange
model is the volumetric efficiency (𝜂𝑉𝐸 ) model The remaining cylinder process models
are used to predict the Net Indicated Mean Effective Pressure (𝐼𝑀𝐸𝑃𝑛 ), Coefficient of
39
Variation of 𝐼𝑀𝐸𝑃𝑛 (𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 ), Turbine inlet gas temperature (𝑇𝐸𝑥ℎ ), and the square of
Table III-2. Actuator combinations used for data collection for ANN model training
The data acquisition is done conveniently using ETAS INCA’s Open Hardware
is required for the calculation of 𝐼𝑀𝐸𝑃𝑛 and 𝐾𝐼 2 . AVL Indicom’s calc-graf tool is
per cycle. The INCA OHI interface is configured to communicate in real-time with AVL
Indicom to receive a configured list of parameters computed from the cylinder pressure.
This list of parameters is then available in the INCA workspace which also consists of the
control actuator signals sent to the engine. This test structure shown in Figure III-11
40
facilitates collection of data in one file per test using only one software (INCA) which
Figure III-11. Experimental setup to facilitate data acquisition for engine cylinder models
portion of the five seconds per combination of input actuators. This is done because the
measurement of 𝜂𝑉𝐸 requires measurement of mass flow rate across the intake valves
which is not done directly on the engine. Instead, the mass air flow rate sensor which is
far upstream of the cylinders is utilized, which during transient operation is not
a steady state. Typically, this calculation is done based on hundreds of engine cycles of
steady state data. For the un-boosted test conditions upto 600 consecutive steady-state
41
cycles could be identified from the data. However, for some of the high-load boosted
conditions, the 𝐼𝑀𝐸𝑃𝑛 would still be in a slow transient at the end of the five seconds due
to the slow dynamics (exhaust manifold temperature, intercooler outlet gas temperature)
of the engine. For such conditions, the last twenty cycles of the engine operation before
Input Unit
Intake temperature K
RPM rpm
42
3.2.1 𝜼𝑽𝑬 model
Neural network based methods have been investigated for engine air-flow
performed using Matlab’s neural network toolbox [86]. The accuracy for the model for
the collected data can be seen in Figure III-12. This network is modeled using a fifteen
neuron ANN with a single hidden layer. Majority of the test data points fall within the +/-
Figure III-12. Modeled vs measured volumetric efficiency [%] for all data with +/- 10%
43
3.2.2 𝑰𝑴𝑬𝑷𝒏 model
Similar to the volumetric efficiency model the IMEPn is modeled using a fifteen
neuron, single hidden layer ANN. The model seems to lose accuracy at the lower load
points due to the high combustion variability noticed at these conditions resulting in
Figure III-13. Modeled vs measured IMEPn for all data with +/- 10% error bands
44
3.2.3 𝑪𝑶𝑽𝑰𝑴𝑬𝑷𝒏 model
The 𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 is computed as follows for steady state ensembles of data.
𝑠𝑡𝑑𝑑𝑒𝑣(𝐼𝑀𝐸𝑃𝑛 )
𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 = III-29
𝑚𝑒𝑎𝑛(𝐼𝑀𝐸𝑃𝑛 )
the average 𝐼𝑀𝐸𝑃𝑛 . The model is able to capture trends but lacks accuracy.
Figure III-14. Modeled vs measured 𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 for all data with +/- 10% error bands
45
3.2.4 𝑲𝑰𝟐 model
The 𝐾𝐼 2 model has the worst accuracy of all the models. This is primarily due to
the lack of training data in the heavy knocking region. Since knocking is very detrimental
Figure III-15. Modeled vs measured 𝐾𝐼 2 for all data with +/- 10% error bands
46
3.2. 𝑻𝑬𝒙𝒉 model
Similar to the volumetric efficiency and 𝐼𝑀𝐸𝑃𝑛 models, the steady state exhaust
temperature model is able to predict most of the points within the +/- 10% region. The
model was fitted to steady state exhaust temperature because the time constant associated
with the transient exhaust temperature is much longer than the turbocharger time
constant. Including these dynamics in the control model would require lengthening of the
horizon and consequently increasing the sampling time of the MPC to maintain real-time
execution feasibility, which would then require removing some of the pressure dynamics
Figure III-16. Modeled vs measured steady state exhaust temperature for all data +/- 10%
error bands represented by dashed red line. The accuracy of each of these neural
47
Table III-4. Accuracy of trained ANNs
𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 1.4%
𝐾𝐼 2 0.297 Bar2
𝜂𝑉𝐸 6.7%
𝑇𝐸𝑥ℎ 14 Deg K
48
3.3 Continuous time state space model of the system
Based on the models described in the previous sections, a continuous time state-
space model of the system is formed. The inputs and outputs are also normalized in
magnitude to have single digit values during operation. The units which are used for
State/input/parameter Unit
Pressure Bar
IMEPn Bar
COV [Ratio]
Temperature Deg K
Volume m^3
Based on the physical unit based normalization scheme above the state dynamics and the
49
̇ = − 𝑉𝑑. 𝜂𝑉𝐸 .𝑅𝑃𝑀.𝑃𝐼𝑚 +
𝑃𝐼𝑚
𝑅.𝑇𝐼𝑚
6
𝑚̇𝐼𝑇𝑉 30
𝑉 .120
𝐼𝑚 𝑉 𝐼𝑚 .6×10
𝑅(𝑇𝐶𝑜𝑚𝑝𝑜𝑢𝑡 +𝑇𝐼𝐶𝑂𝑢𝑡 )
̇ =(
𝑃𝐵𝑠𝑡 ) (𝑚̇𝐶𝑜𝑚𝑝 − 𝑚̇𝐼𝑇𝑉 − 𝛽𝐵𝑜𝑣 (𝑃𝐵𝑠𝑡 − 𝑃𝐶𝑖𝑛 )) 31
2𝑉𝐵𝑠𝑡
𝐽−1
𝑇
𝜔̇ 𝑇 = 1000 (𝜏 𝑇𝑢𝑟𝑏 − 𝜏𝐶𝑜𝑚𝑝 ) 33
𝑇
𝑥 = [𝑃𝐼𝑚 𝑃𝐵𝑠𝑡 𝑃𝐸𝑔𝑟 𝜔 𝑇 ] , 𝑥 ∈ ℝ4×1 34
Where, suffixes 𝐼𝑚, 𝐵𝑠𝑡and 𝐸𝑔𝑟 mean Intake manifold, post-compressor boost
manifold, and the EGR in the intake manifold. The input vector to the model is defined as
follows.
𝑇
𝑢 = [𝑚̇𝐼𝑇𝑉 𝛽𝐵𝑜𝑣 𝑚̇𝑊𝐺 𝑚̇𝐸𝑔𝑟 𝑆𝑃𝐾𝐼𝐶𝐿𝐸𝐶𝐿] , 𝑢 ∈ ℝ7×1 35
Where, suffixes 𝐼𝑇𝑉, 𝐵𝑜𝑣and 𝑊𝐺 denote Intake throttle valve, Blow-off valve, and
Turbine waste-gate respectively. The output that is to be tracked and the constraints are
𝑦 = 𝐼𝑀𝐸𝑃𝑛 , 𝑦 ∈ ℝ 36
𝑇
𝑧 = [(𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 )(𝐾𝐼 2 )(𝑃𝐼𝑚 − 𝑃𝐵𝑠𝑡 )(𝑚̇𝑆𝐿 − 𝑚̇𝐶𝑜𝑚𝑝 )(𝑚̇𝑊𝐺 − 𝑚̇𝐸𝑛𝑔 )] , 𝑧 ∈ ℝ5×1
37
50
Where, the constraint 𝑃𝐼𝑚 − 𝑃𝐵𝑠𝑡 < 0 and 𝑚̇𝑊𝐺 − 𝑚̇𝐸𝑛𝑔 < 0ensures that the flow across
the intake throttle valve and the waste-gate is physically meaningful. The constraint
𝑚̇𝑆𝐿 − 𝑚̇𝐶𝑜𝑚𝑝 < 0 ensures that the compressor mass flow rate is always on the right hand
side of the surge line modeled as a linear function of 𝑃𝐵𝑠𝑡 as follows similar to [88].
Where, 𝑆𝐺1 and 𝑆𝐺2 are the coefficients of the surge-line of the compressor.
Open-loop validation of the state space model against experimentally measured engine
data has been shown in Figure III-17 for 3000RPM. The complete error metrics for the
0D engine model validated against four different engine transients at 2000 and 3000
RPM are shown in
Table III-6. The model is within +/- 10% accuracy for the states and the IMEPn output.
Table III-6 Error metrics of the 0D model for four different transient engine trajectories.
Wt 2.39 9.15%
51
Figure III-17 Open-loop system trajectory for 0D model vs experimentally collected data
52
IV. NON-LINEARITY ANALYSIS OF CONTROL ORIENTED MODEL
approximation . Previous research has used approximated linear state-space models for
utilization in MPC. The underlying assumption of these approaches is that the behavior of
the non-linear engine system is linear-like for the entirety of the prediction horizon of the
MPC. In order to understand the effect of this assumption on the performance of MPC,
The deviation of the locally linearized model from the non-linear model is quantified
The LAE has been evaluated for a sinusoidal as well as a step change in inputs
The analysis has been done for six different steady state operating points which are
combinations of 2000 and 3000 RPM and IMEPn load levels of 5, 10 and 15 Bar.
of the non-linear model 30 to 37 of the system is often made while designing MPC
algorithms. The IMEPn accuracy of the linearized model is of specific importance for
MPC designed for torque reference tracking. The LAE for the IMEPn output has been
quantified by first linearizing the model at a steady state operating point. Then the open
53
loop inputs which correspond to a deviation from the steady state IMEPn are applied to
both the models. The LAE for IMEPn is then examined as a function of time-step N from
Figure IV-1 Deviation between the non-linear system model and the linearized system
model as a function of time step N from the linearization. The deviation is induced by
The IMEPn LAE is quantified as the absolute error between the IMEPn from the non-
54
The 𝛿𝐼𝑀𝐸𝑃𝑛 is computed for different open loop step input 𝑚̇𝐼𝑇𝑉 for which the final
values correspond to the steady-state deviation from the linearization point IMEPn. This
is denoted in general as ∆𝐼𝑀𝐸𝑃𝑛. Figure IV-2 shows the 𝛿𝐼𝑀𝐸𝑃𝑛 as a function of the
time-step after linearization N and ∆𝐼𝑀𝐸𝑃𝑛 for all six operating conditions. As expected
for ∆𝐼𝑀𝐸𝑃𝑛 = 0, all of the conditions show 𝛿𝐼𝑀𝐸𝑃𝑛 = 0 since there is no change in
𝛿𝐼𝑀𝐸𝑃𝑛 begins to grow larger. The highest deviation is seen at the 3000 RPM 15 Bar
case for ∆𝐼𝑀𝐸𝑃𝑛 = −4. This implies that the non-linear 𝐼𝑀𝐸𝑃𝑛 output model varies
more with time at this operating condition compared to the others. Similarly, the least
deviation is seen at 3000 RPM 5 Bar case. Implementing MPC using such approximated
models would result in possibly reduced optimality and constraint violation. Based on the
analysis done on the six operating conditions shown in the errors induced by linear
linearization within the horizon causes 𝛿𝐼𝑀𝐸𝑃𝑛 to naturally reduce due to improved
local accuracy of the linearized model at the additional linearization points. However,
identify the number of times the non-linear model needs to be linearized in order to
maintain a prescribed model accuracy. Table IV-1. shows the minimum number of
𝐼𝑀𝐸𝑃𝑛.
55
56
Figure IV-2 Model error as a function of commanded open-loop change in IMEPn
shows that the deviation between linearized and non-linear model varies as a function of
operating condition and the time –step after the point of linearization
[%]
15 Bar 1 2 5
10 Bar 1 1 5
5 Bar 10 2 5
15 Bar 1 5 3
10 Bar 2 2 3
5 Bar >10 5 3
15 Bar 5 5 1
10 Bar 5 5 1
5 Bar >10 10 1
57
4.3 Sinusoidal input transient testing
Hz was applied to the model to understand the LAE as a function of input frequency. The
significance of this test is to analyze the dependence between the LAE and the rate at
which the engine transient occurs. Similar to the step input test, Table IV-2 shows the
𝛿𝐼𝑀𝐸𝑃𝑛 as a function of the time-step after linearization N and frequency of the input
that was applied to the models. At every operating condition a certain frequency results in
the maximum LAE. This frequency ranges between 0.5 and 1.5 Hz depending on the
operating condition. However, similar to the step-input test the minimum 𝛿𝐼𝑀𝐸𝑃𝑛 is
seen at the 3000 RPM 5 Bar case. The maximum 𝛿𝐼𝑀𝐸𝑃𝑛 is seen at the 3000 RPM 10
Bar case for 0.9 Hz input frequency. Similar to the step-input case the minimum number
of linearizations required for different prescribed accuracy levels have been quantified. In
In both of the tests we can see that some of the operating conditions require the model to
be linearized more than 10 times within the horizon to meet the desired accuracy level.
This is due to the inaccuracy associated with discretization of the continuous model.
Since the sample time chosen here is 50ms, for a prediction horizon of 500ms, the
58
Table IV-2 Minimum number of model linearizations required in horizon to maintain
15 Bar 1 2 5
10 Bar 2 5 5
5 Bar >10 5 5
15 Bar 2 5 3
10 Bar 5 5 3
5 Bar >10 5 3
15 Bar 10 5 1
10 Bar 10 10 1
59
Figure IV-3 Model error as a function of frequency of commanded open-loop model
input shows that the deviation between linearized and non-linear model generally peaks
60
The choice of prediction model utilized for MPC is critical for the stability,
optimality and robustness of the closed loop controller performance. This research
presents a method of evaluating the impact of local linear approximation error on the
modeled IMEPn by manipulation of the throttle mass flow rate input. The same method
could be applied to every other input of the model to quantify the effect on the rest of the
The transient tests conducted in this study show clear deviation between the non-
linear model and its linearized version for different operating conditions. Based on this
analysis it is clear that this deviation varies significantly with the type of transient
induced in the system. The step transients are generally more forgiving with respect to
the number of model linearizations required, compared to the sinusoidal transients. The
sinusoidal transients clearly show a narrow frequency range where the deviation in the
models is highest.
condition and the commanded transient to the MPC. However, for some operating
conditions, linearization even at every stage within the prediction horizon is not sufficient
61
V. NON-LINEAR MODEL PREDICTIVE CONTROL DESIGN
control. Complex cycle-cycle behavior and fast dynamics can be handled effectively by
sample time 𝛿𝑡 is dependent on the engine dynamics considered in the MPC control
model, the prediction/control horizon considered and the computational ability of the
accommodate the slowest dynamics in the system [55][56]. The turbocharger speed has
the slowest dynamics in the system with time constant in the order of seconds (ignoring
the exhaust manifold wall temperature and the intercooler outlet temperature dynamics).
Figure V-1 Approximate relationship between horizon length considered for MPC,
62
Figure V-2 The selection of MPC sample time and prediction horizon time dictates the
horizon length of 10 steps which is 0.5 s. This allows consideration of the turbocharger
dynamics, the filling and emptying dynamics of the boost manifold and the intake
manifold. The exhaust manifold and the other control volumes on the engine are too
small and hence have almost instantaneous dynamics which can then be handled by lower
level controllers.
Based on the study shown in the previous chapter, it is evident that utilizing one
linearized model to predict the behavior of the engine over a period of time similar to the
prediction horizon of the MPC can result in significant errors in the modeled outputs.
Ideally, linearization at every stage within the prediction horizon would be desirable to
conserve accuracy. However, the process of obtaining the linearized model requires
63
computational resources. The following section will explore the methods used to
The system model described in (30) to (37) can be linearized as follows using the
Taylor’s series expansion (ignoring higher terms) in continuous time domain as follows.
𝜕𝑓 𝜕𝑓
𝑥̇ = 𝑓(𝑥0 , 𝑢0 ) + 𝛿𝑥̇ = 𝑥̇ 0 + 𝜕𝑥 | 𝛿𝑥 + 𝜕𝑢| 𝛿𝑢 40
𝑥0 ,𝑢0 𝑥0 ,𝑢0
𝜕𝑔 𝜕𝑔
| |
𝑦 𝑦0 𝜕𝑥 𝑥0 ,𝑢0 𝜕𝑢 𝑥0 ,𝑢0
[ ] = [ 𝑧 ] + [𝜕ℎ ] 𝛿𝑥 + [𝜕ℎ ] 𝛿𝑢 41
𝑧 0 | |
𝜕𝑥 𝑥0 ,𝑢0 𝜕𝑢 𝑥0 ,𝑢0
Where, 𝑥0 and 𝑢0 are the nominal state vectors and input vectors respectively. 𝑦0
and 𝑧0 are the corresponding nominal tracking output and constraint output vectors. 𝑥̇ 0 is
the vector of derivatives of the states evaluated at the nominal point. 𝛿𝑥 is the deviation
𝜕𝑓 𝜕𝑓
of the states from the nominal point. | , | are the Jacobian matrices evaluated
𝜕𝑥 𝑥0 ,𝑢0 𝜕𝑢 𝑥0 ,𝑢0
at 𝑥0 and 𝑢0 by partial differentiation of the state dynamics 3-30 to 3-33 with respect to 𝑥
𝜕𝑔 𝜕𝑔 𝜕ℎ 𝜕ℎ
and 𝑢. Similarly, 𝜕𝑥 | , 𝜕𝑢| , 𝜕𝑥 | and 𝜕𝑢| are the Jacobian matrices
𝑥0 ,𝑢0 𝑥0 ,𝑢0 𝑥0 ,𝑢0 𝑥0 ,𝑢0
evaluated at 𝑥0 and 𝑢0 by partial differentiation of the output functions 3-36 to 3-37 with
respect to 𝑥 and 𝑢.
64
5.2.1 Drawbacks of numerical linearization
𝜕𝑐(𝑥) (𝑐(𝑥+Δ𝑥)−𝑐(𝑥))
≈ 42
𝜕𝑥 Δ𝑥
to approximate the slope of the arbitrary non-linear function 𝑐. Utilizing this method is
However, since the engine system model has four states and seven inputs, utilizing this
method to obtain the linearized system model at one nominal operating condition will
require thirty eight evaluations of the system dynamics 3-30 to 3-33. Similarly, in order
to obtain the linearized output models, the output functions 3-36 to 3-37 would have to be
evaluated fifty six times. This would lead to ninety four evaluations of the system model
in order to obtain the linearized system dynamics. This process would have to be repeated
at every stage in the prediction horizon in order to minimize the linear approximation
error, resulting in nine hundred and forty evaluations of the system model at every time-
step of execution. This process takes a large amount of memory and processor throughput
and leaves no resources left for the optimization algorithm. In the worst case, it will cause
a task overrun wherein the computation fails to execute within the given time-step.
65
5.2.2 Hybrid linearization of the system model
approach described in the previous section, the analytical solution is exact and it requires
significantly less number of executions of the system model. A series of steps taken in
order to maximize analytical differentiation within the system model is described in the
following sections.
The system model is first analyzed for potential to apply analytical partial
differentiation. In this the first process is to identify the fixed zeros in the system model.
𝜕𝑓𝑃𝐼𝑚 𝜕𝑓𝑃𝐼𝑚
𝟎 𝟎
𝜕𝑃𝐼𝑚 𝜕𝑃𝐸𝑔𝑟
𝜕𝑓𝑃𝐵𝑠𝑡 𝜕𝑓𝑃𝐵𝑠𝑡
𝟎 𝟎
𝜕𝑓 𝜕𝑃𝐵𝑠𝑡 𝜕𝜔𝑇
𝜕𝑥
= 𝜕𝑓𝑃𝐸𝑔𝑟 𝜕𝑓𝑃𝐸𝑔𝑟 43
𝟎 𝟎
𝜕𝑃𝐼𝑚 𝜕𝑃𝐸𝑔𝑟
𝜕𝑓𝜔𝑇 𝜕𝑓𝜔𝑇 𝜕𝑓𝜔𝑇 𝜕𝑓𝜔𝑇
[ 𝜕𝑃𝐼𝑚 𝜕𝑃𝐵𝑠𝑡 𝜕𝑃𝐸𝑔𝑟 𝜕𝜔𝑇 ]
66
𝜕𝑓𝑃𝐼𝑚 𝜕𝑓𝑃𝐼𝑚 𝜕𝑓𝑃𝐼𝑚
𝜕𝑚̇𝐼𝑇𝑉
𝟎 𝟎 𝟎 𝟎 𝜕𝐼𝐶𝐿 𝜕𝐸𝐶𝐿
𝜕𝑓𝑃𝐵𝑠𝑡 𝜕𝑓𝑃𝐵𝑠𝑡
𝜕𝑚̇𝐼𝑇𝑉 𝜕𝛽𝐵𝑂𝑉
𝟎 𝟎 𝟎 𝟎 𝟎
𝜕𝑓
= 𝜕𝑓𝑃𝐸𝑔𝑟 𝜕𝑓𝑃𝐸𝑔𝑟 𝜕𝑓𝑃𝐸𝑔𝑟 44
𝜕𝑢
𝟎 𝟎 𝟎 𝜕𝑚̇𝐸𝑔𝑟
𝟎 𝜕𝐼𝐶𝐿 𝜕𝐸𝐶𝐿
𝜕𝑓𝜔𝑇 𝜕𝑓𝜔𝑇 𝜕𝑓𝜔𝑇 𝜕𝑓𝜔𝑇
[ 𝟎 𝟎 𝜕𝑚̇𝑊𝐺
𝟎 𝜕𝑆𝑃𝐾 𝜕𝐼𝐶𝐿 𝜕𝐸𝐶𝐿 ]
𝜕𝑔 𝜕(𝐼𝑀𝐸𝑃𝑛 ) 𝜕(𝐼𝑀𝐸𝑃𝑛 )
=[ 𝟎 𝟎] 45
𝜕𝑥 𝜕𝑃𝐼𝑚 𝜕𝑃𝐸𝑔𝑟
𝜕(𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 ) 𝜕(𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 )
𝜕𝑃𝐼𝑚
𝟎 𝟎
𝜕𝑃𝐸𝑔𝑟
𝜕(𝐾𝐼 2 ) 𝜕(𝐾𝐼 2 )
𝟎 𝟎
𝜕𝑃𝐼𝑚 𝜕𝑃𝐸𝑔𝑟
𝜕ℎ
= 1 −1 𝟎 𝟎 46
𝜕𝑢
𝜕(𝜉𝑆𝑢𝑟𝑔𝑒 ) 𝜕(𝜉𝑆𝑢𝑟𝑔𝑒 )
𝟎 𝟎
𝜕𝑃𝐵𝑠𝑡 𝜕𝜔𝑇
𝜕(𝜉𝑊𝐺 ) 𝜕(𝜉𝑊𝐺 )
𝟎 𝟎
[ 𝜕𝑃𝐼𝑚 𝜕𝑃𝐸𝑔𝑟 ]
The fixed zeros are identified by simply inspecting the differential equations in
state dynamics and the output equations. A fixed zero would simply denote the absence
of the corresponding state or output within that equation/model. It is evident that there is
a significant number of fixed zeros in the model which substantially alleviates the
computational burden of model linearization. Since the forward throttle flow constraint
𝜕ℎ
𝜉𝐼𝑇𝑉 is a linear function of the states the corresponding elements of 𝜕𝑢 are 1 and -1
respectively.
67
5.2.2.2 Analytical and hybrid differentiation of system model
The analytical differentiation of the system dynamics for 𝑃𝐼𝑚 and 𝑃𝐸𝐺𝑅 is simple
𝜕𝑓𝑃𝐼𝑚 𝑉 𝑅𝑃𝑀 𝜕𝜂
𝜕𝑃𝐼𝑚
𝑑
= −120𝑉 (𝑃𝐼𝑚 𝜕𝑃𝑉𝐸 + 𝜂𝑉𝐸 ) 47
𝐼𝑚 𝐼𝑚
𝜕𝑓𝑃𝐼𝑚 𝑉 𝑅𝑃𝑀 𝜕𝜂
𝑑
= −120𝑉 (𝑃𝑖𝑚 𝜕𝑃 𝑉𝐸 ) 48
𝜕𝑃𝐸𝑔𝑟 𝐼𝑚 𝐸𝑔𝑟
𝜕𝑓𝑃𝐼𝑚 𝑇𝐼𝑚 𝑅
=𝑉 6
49
𝜕𝑚̇𝐼𝑇𝑉 𝐼𝑚 6×10
𝜕𝑓𝑃𝐸𝑔𝑟 𝑉 𝑅𝑃𝑀 𝜕𝜂
𝑑
= −120𝑉 (𝑃𝐸𝑔𝑟 𝜕𝑃𝑉𝐸 ) 52
𝜕𝑃𝐼𝑚 𝐼𝑚 𝐼𝑚
𝜕𝑓𝑃𝐸𝑔𝑟 𝑉 𝑅𝑃𝑀 𝜕𝜂
𝑑
= −120𝑉 (𝑃𝐸𝑔𝑟 𝜕𝑃 𝑉𝐸 + 𝜂𝑉𝐸 ) 53
𝜕𝑃𝐸𝑔𝑟 𝐼𝑚 𝐸𝑔𝑟
𝜕𝑓𝑃𝐸𝑔𝑟 𝜕𝑓𝑃
= 𝜕𝑚̇ 𝐼𝑚 54
𝜕𝑚̇𝐸𝑔𝑟 𝐼𝑇𝑉
68
It is visible in this analytical framework that there are common computations
𝜕𝜂𝑉𝐸
shared between the models. E.g. 𝜕𝐼𝐶𝐿
is common in (50) and (55) and is computed only
once, further reducing the computational load. For 𝑃𝐵𝑠𝑡 a hybrid approach is utilized. The
𝜕𝑓𝑃𝐵𝑠𝑡 𝜕𝑓𝑃𝐵𝑠𝑡
numerical differentiation as shown in (42) is used for and because of the
𝜕𝑃𝐵𝑠𝑡 𝜕𝜔𝑇
However, the linearization w.r.t. the inputs is still done analytically as follows.
𝜕𝑓𝑃𝐵𝑠𝑡 𝑇𝐵𝑠𝑡 𝑅
= −𝑉 6
57
𝜕𝑚̇𝐼𝑇𝑉 𝐵𝑠𝑡 6×10
𝜕𝑓𝑃𝐵𝑠𝑡 𝑇𝐵𝑠𝑡 𝑅
= (𝑃𝐵𝑠𝑡 − 𝑃𝐶𝑖𝑛 ) (−𝑉 6
) 58
𝜕𝛽𝐵𝑂𝑉 𝐵𝑠𝑡 6×10
solutions which are marginally faster than the numerical solutions obtained from. Hence,
the numerical approach was utilized in obtaining these partial derivatives but after some
simplification as follows.
𝜕𝑓𝜔𝑇 1 𝜕𝜏
= 1000𝐽 ( 𝜕𝑃𝑇𝑢𝑟𝑏) 59
𝜕𝑃𝐼𝑚 𝑇 𝐼𝑚
𝜕𝑓𝜔𝑇 −1 𝜕𝜏𝐶𝑜𝑚𝑝
= 1000𝐽 ( ) 60
𝜕𝑃𝐵𝑠𝑡 𝑇 𝜕𝑃𝐵𝑠𝑡
69
𝜕𝜏𝐶𝑜𝑚𝑝 𝜕𝜏𝑇𝑢𝑟𝑏
Where, the numerical computation associated with and is eliminated
𝜕𝑃𝐼𝑚 𝜕𝑃𝐵𝑠𝑡
due to the absence of the corresponding states in these equations, similar to the fixed
zeros.
Majority of the functions used to generate the tracking and constraint output are
the ANN models described in the previous chapter. Hence, the linearization of the
outputs requires linearization of the ANNs w.r.t. their inputs. A typical ANN model has
70
This model can be mathematically represented as follows.
Where, 𝑊 𝑖 and 𝑏 𝑖 are the weight and bias vectors for the ith layer of the ANN
model respectively. 𝑥 is the input vector and n is the total number of layers in the
function. s is the sigmoid activation function of the following form. Hence as shown in
𝛼1 = 𝑊 1 𝑥 + 𝑏1 V-62
𝛼 2 = 𝑊 2 𝑠(𝛼 1 ) + 𝑏2 V-63
1
𝑠(𝛼) = 65
1+𝑒 −𝛼
Similar to the ‘Back Propagation’ algorithm used to train the ANN models, the
71
It is noteworthy that 𝑠(𝛼) is in fact computed during the forward execution of the ANN
model and is already available. Hence the partial derivative of the model with respect to
𝜕𝑦
= 𝑠 ′ (𝑊 𝑛 𝑠(𝑊 𝑛−1 𝑠(… 𝑠(𝑊 1 𝑥 + 𝑏1 ) + 𝑏 𝑛−1 ) + 𝑏 𝑛 )(𝑊 𝑛 )𝑇 𝑠 ′ (𝑊 𝑛−1 𝑠(… 𝑠(𝑊 1 𝑥 +
𝜕𝑥
Where, (66) is utilized to compute 𝑠′ using results of the computation of 𝑠 from the
forward execution of the model. Figure V-4 shows the difference in execution time
between the analytical differentiation method shown above and the numerical
differentiation for different combinations of inputs. The execution of both methods was
performed on a desktop computer. The analytical method is over ten times faster than the
numerical approach.
72
Figure V-4. Comparison between execution times of 𝐼𝑀𝐸𝑃𝑛 ANN model
Figure V-4 shows the difference in execution time between the numerical linearization
and the hybrid model linearization for the entire system model using the methods
described in this section. Similar to the difference in execution time for the ANN models
shown in Figure V-5. The execution time for the system model linearization using the
hybrid approach is also over ten times faster than the numerical approach.
73
Figure V-5. Comparison between execution times of complete system model
model in time. The non-linear model is open-loop stable in continuous time domain,
hence it is necessary to ensure that the discrete-time, linear model used for MPC is open-
loop stable as well to preserve accuracy. The manifold filling and emptying dynamics
have time constants which vary with engine load and speed. At high engine flow rates,
these time constants becomes significantly shorter than 𝛿𝑡. Such fast dynamics could be
74
eliminated from the model at these conditions, but that would require a variable-order
model which is not desirable for MPC formulation. Euler’s finite difference
time to discrete time domain. However using this method for the linear system leads to
numerical instability at some high load/speed conditions. . Utilizing the State Transition
Matrix (STM) method for time domain discretization generally preserves numerical
stability to a higher degree compared to the finite difference approach [83]. The discrete
𝛿𝑡
𝛿𝑥𝑘+1 = 𝑒 𝐴𝛿𝑡 𝛿𝑥𝑘 + (∫0 𝑒 𝐴(𝛿𝑡−𝜏) 𝑑𝜏𝐵) 𝛿𝑢𝑘 68
Where, 𝑒 𝐴𝛿𝑡 and 𝑒 𝐴(𝛿𝑡−𝜏) ) are matrix exponentials computed using the Pade
approximation. Hence, the discrete time model of the following form is derived from the
STM method.
𝑦𝑘 𝑦0
[ 𝑧 ] = [ 𝑧 ] + 𝐶𝑑 𝛿𝑥𝑘 + 𝐷𝑑 𝛿𝑢𝑘 70
𝑘 0
controller, the commands generated by the MPC algorithm are available only at the
beginning of the next time step. This results in a unit time-step delay in the control
actions. However, the prediction model () has a non-zero D matrix which does not
75
account for this delay and also isn’t favorable for MPC-QP formulation. In order to
𝛿𝑥𝑘+1 𝐴 𝐵𝑑 𝛿𝑥𝑘 0
[ ]=[ 𝑑 ][ ] + [ ] 𝛿𝑢𝑘 71
𝛿𝑥̃𝑘+1 0 0 𝛿𝑥̃𝑘 𝐼
𝑦𝑘 𝑦0 𝛿𝑥𝑘
[ 𝑧 ] = [ 𝑧 ] + [𝐶𝑑 𝐷𝑑 ] [ ] 72
𝑘 0 𝛿𝑥̃𝑘
Augmentation of the original state vector with 𝑥̃𝑘 automatically induces a unit
step delay in the model and eliminates the 𝐷 matrix. The prediction model (71) to (72) is
time-varying. Hence, it is derived at every stage within the prediction horizon using the
method shown above and the set of these derived models is utilized to formulate the
MPC-QP.
The formulation of the MPC-QP largely depends on the chosen objective function
as follows.
𝑘+𝑁−1
𝑖=𝑘
76
Where,
𝑄 and 𝑄𝑁 are the tuning matrices for the stage and terminal penalties on 𝐼𝑀𝐸𝑃𝑛
tracking error. 𝑄𝑁 is chosen large enough for closed loop stability [55][56]. 𝑅 is the
tuning matrix to penalize tracking error of the inputs and is always positive definite.
The reference input 𝑢𝑅𝑒𝑓 is the input vector which minimizes fuel consumption
of the engine at steady state conditions. It is also utilized to generate the reference
trajectory of the prediction model around which the MPC finds the optimal solution.
𝑚̇𝐼𝑇𝑉𝑟𝑒𝑓 , 𝑚̇𝑊𝐺𝑟𝑒𝑓 and 𝑚̇𝐸𝐺𝑅𝑟𝑒𝑓 which are part of 𝑢𝑅𝑒𝑓 are particularly important to derive
the reference trajectory of the model in the prediction horizon. The steady state values of
𝑚̇𝐼𝑇𝑉𝑟𝑒𝑓 under boosted conditions can only be achieved by increasing the turbine power
by sufficiently reducing 𝑚̇𝑊𝐺 . Hence using the steady state 𝑚̇𝐼𝑇𝑉𝑟𝑒𝑓 can lead to a
reference trajectory which is physically impossible and undesirable. E.g. 𝑃𝐵𝑠𝑡 may drop
well below 𝑃𝐼𝑚 in the reference trajectory, even low enough to drop the compressor
pressure ratio below 1, resulting in choked operation. At this condition, the numerically
𝜕(𝑚̇𝐶𝑜𝑚𝑝 )
derived becomes 0, decoupling 𝑃𝐵𝑠𝑡 from the compressor mass flow rate model.
𝜕𝑃𝐵𝑠𝑡
This results in only 𝜔 𝑇 influence on the compressor mass flow rate. Since 𝜔 𝑇 has the
slowest dynamics, the compressor mass flow rate practically is ‘locked’ at the choked
position corresponding to 𝜔 𝑇 at that instant in the reference trajectory. For large tip-in
77
maneuvers, this problem results in an infeasible MPC-QP problem due to the imposed
𝜉𝐼𝑇𝑉 constraint. Similarly, utilizing the steady state 𝑚̇𝑊𝐺𝑟𝑒𝑓 and 𝑚̇𝐸𝐺𝑅𝑟𝑒𝑓 to obtain the
reference trajectory results in the 𝜉𝑊𝐺 and 𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 constraints being violated in the
reference trajectory. To address these problems in deriving 𝑢𝑅𝑒𝑓 , the following steps have
been taken. 𝑚̇𝐼𝑇𝑉𝑟𝑒𝑓 is derived using the orifice flow model and the corresponding steady
min
𝑎𝑟𝑔 𝐽(𝑥(𝑘), 𝑈(𝑘))
𝑈(𝑘)
Subject to:
𝑧(𝑖) ≤ [0.151000]𝑇
Based on the cost function Error! Reference source not found. and the prediction
min 1 𝑇
𝑈 𝐻𝑈 + 𝐿𝑇 𝑈
𝑈 2
𝑀𝑈 ≤ 𝐹
Where,
𝐻 ∈ ℝ70×70
>0 , 𝐿 ∈ ℝ70 , 𝑀 ∈ ℝ70×70
>0 , 𝐹 ∈ ℝ70 , 𝑈 ∈ ℝ70
78
Solving the MPC-QP described above results in the optimal solution 𝑈 ∗ which is a
sequence of sub-optimal inputs in the vicinity of the reference inputs 𝑈𝑟𝑒𝑓 out of which
the input vector which corresponds to the current time step 𝑢∗ (𝑘) is chosen and applied
A 1D-dynamical engine model designed in GT-Suite was coupled with the control
79
Figure V-7 IMEPn response of MPC co-simulated with GT-Power
immediately visible that the MPC tracks the commanded IMEPn with some steady state
offset. This occurs due to the 𝑢𝑟𝑒𝑓 calibration being derived from the experimental engine
data. This reference input corresponds to a steady state IMEPn which is higher than the
desired value. Due to the tuning penalty𝑅 on the reference inputs, the MPC generates
80
inputs which are a ‘trade-off’ between the steady-state reference inputs𝑢𝑟𝑒𝑓 and the
Figure V-8 Control actuator trajectories for MPC cosimulation with GT-Power, air-path
actuators are on the left hand side and cylinder actuators are on the right hand side. Black
dashed lines are the steady-state reference inputs and blue lines are the inputs generated
by MPC
81
5.7.2 Air-path actuator trajectories
The control actuator trajectories for the IMEPn transient are shown in Figure V-8
The MPC shows differences in air-path actuation for the two tip-in maneuvers shown.
For the first tip-in at 2 seconds, the MPC commands minimal change in 𝑚̇𝐸𝑔𝑟 whereas,
for the second tip-in at 12 seconds the MPC commands a complete stoppage of 𝑚̇𝐸𝑔𝑟 for
a short period of time. 𝑚̇𝑊𝐺 is also stopped temporarily at the time of tip-in followed by a
return to it’s reference value. 𝛽𝐵𝑂𝑉 actuation on the second tip-in mitigates the drop in
𝑃𝐵𝑠𝑡 caused by the rapid increase of 𝑚̇𝐼𝑇𝑉 . For the tip-out at 7 seconds MPC mainly
The 𝑆𝑃𝐾 (Spark timing) actuator is significantly retarded from the reference
value during both the tip-in maneuvers. MPC also advances the intake cam position 𝐼𝑉𝑂
during the tip-in maneuvers for volumetric efficiency and air-flow improvement. During
the tip-out the MPC advances 𝑆𝑃𝐾 to handle the 𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 constraint.
The last three constraint outputs in the constraint vector are purely air-path
constraints. Out of these constraints the constraints 𝑃𝐼𝑚 − 𝑃𝐵𝑠𝑡 < 0 and
𝑚̇𝑊𝐺 − 𝑚̇𝐸𝑛𝑔 < 0 are always satisfied on the GT-Power model independent of the MPC
commands. These constraints are in place merely to ensure physically feasible behavior
in the control oriented model (30) to (37). The surge constraint handling is demonstrated
82
in Figure V-9 where, the compressor operation momentarily crosses the surge-line but
Since the knock model is not tuned and the 𝐶𝑂𝑉𝐼𝑀𝐸𝑃𝑛 cannot be evaluated using the GT-
83
VI. LOWER-LEVEL CONTROL AND ESTIMATION
The commands generated by MPC require conversion into actuator set-points for
real-time application. This lower-level actuation is categorized into two sections, air-path
control and engine-cylinders control. Real time estimation of the values of the states is
The air-path control consists of resolving the valve mass flow rate commands 𝑚̇𝐼𝑇𝑉 , 𝛽,
𝑚̇𝑊𝐺 , and 𝑚̇𝐸𝐺𝑅 into valve position set-points using the orifice flow equation as follows.
𝑚̇√𝑅𝑇
𝐴𝑒 = 𝑃𝐼𝑛 76
𝜓( )𝑃
𝑃𝑂𝑢𝑡 𝐼𝑛
𝜃 = 𝑙𝑢𝑡(𝐴𝑒) 77
Where, 𝑙𝑢𝑡 is the lookup table function to convert commanded valve area 𝐴𝑒 to a valve
position 𝜃. 𝜓 is the flow function which has an infinite gradient when the pressure
difference ∆𝑃 across the valve approaches as shown in Figure VI-1. This results in
84
Figure VI-1. The gradient of the flow function becomes very steep as the pressure
In order to address this issue a simple filter with the following structure is applied.
𝑤(𝑘) = |𝑔∆𝑃(𝑘)| 79
∆𝑃(𝑘) ≈ 0 while also having sufficiently fast transient response. The engine-cylinders
85
6.2 Low-pressure EGR ∆𝑷control
Recent findings show the difficulty in control of mass flow rate across the LPE
system [37][38]. The standard orifice model accuracy deteriorates rapidly in a highly
pulsating ∆𝑃𝐿𝑃𝐸 ≈ 0. Increasing ∆𝑃𝐿𝑃𝐸 to atleast 5 kPa was observed to maintain the
orifice flow model’s accuracy within 5%. Hence, the 𝑃𝐶𝑇 valve is actuated using𝐴𝑒=
𝑚̇√𝑅𝑇
𝑃𝐼𝑛 (76). The mass air-flow sensor unit which is shortly upstream of the 𝑃𝐶𝑇
𝜓( )𝑃
𝑃𝑂𝑢𝑡 𝐼𝑛
valve is used for the 𝑚̇ and 𝑃𝐼𝑛 input. 𝑃𝑂𝑢𝑡 is simply set to 𝑃𝐼𝑛 − 5𝑘𝑃𝑎. Since, the
turbine outlet pressure is always higher than 𝑃𝐼𝑛 , ∆𝑃𝐿𝑃𝐸 ≥ 5𝑘𝑃𝑎 is guaranteed. However,
this method leads to much higher ∆𝑃𝐿𝑃𝐸 at higher power conditions where the ‘natural’
∆𝑃𝐿𝑃𝐸 without PCT throttling is close to 5 kPa. This method results in unnecessary
compressor inlet throttling, potentially limiting peak torque of the engine. Although this
Since the prediction model is physics based, the states have physical meaning as
well. 𝑃𝐼𝑚 and 𝑃𝐵𝑠𝑡 measurements are part of the production engine sensor set. An intake
oxygen sensor is installed at the inlet of the intake manifold for real time measurement of
𝑃𝐸𝑔𝑟 . These measurements are also low pass filtered to minimize noise and oscillations
induced from the discrete-event gas exchange processes. The cutoff frequency high
86
enough to maintain transient accuracy within 50ms. 𝜔 𝑇 is estimated using the Fixed Point
Iteration method described in [89]. The compressor model is utilized for estimation.
̂𝐶𝑜𝑚𝑝 )
𝛼(Π𝑐𝑜𝑚𝑝 −Π ̂̇ 𝐶𝑜𝑚𝑝 )
(1−𝛼)(𝑚̇𝐶𝑜𝑚𝑝 −𝑚
̂ 𝑇 (𝑘) = 𝜔
𝜔 ̂ 𝑇 (𝑘 − 1) + 𝐾 ( ̂
𝜕Π
+ ̂̇
𝜕𝑚
) 80
𝐶𝑜𝑚𝑝 𝐶𝑜𝑚𝑝
̂𝑇
𝜕𝜔 ̂𝑇
𝜕𝜔
𝑃𝐵𝑠𝑡
Π𝐶𝑜𝑚𝑝 = 81
𝑃𝐶𝑖𝑛
1
𝛼= ̂ 𝐶𝑜𝑚𝑝
𝜕Π
82
1− ̂
𝜕𝑚̇𝐶𝑜𝑚𝑝
̂ 𝐶𝑜𝑚𝑝 and 𝑚
Where, Π ̂̇ 𝐶𝑜𝑚𝑝 is the predicted pressure ratio and compressor mass flow rate
compressor model. 𝛼 dictates how much emphasis is placed on utilizing Π𝐶𝑜𝑚𝑝 over
turbocharger speed for a fully transient engine test. Some of the scatter in the Figure VI-2
is caused by the noise in measurement of turbocharger speed. The accuracy of the turbo
speed estimation using this method was observed to be within the 15% error band even
for low speed extrapolated region of the compressor map. The RMS error and max error
87
Figure VI-2.Real-time estimated vs measured turbocharger speed using fixed point
iteration method
88
VII. EXPERIMENTAL VALIDATION OF MPC
engine and it’s ECU is shown in Figure VII-1. The MPC derived in the previous section
maintain real-time execution ability using this formulation, a Dspace DS1006 has been
utilized to execute the MPC program similar to [90][91]. The DS1006 executes in two
timer-task mode. The MPC program operates at the sample time 𝛿𝑡. At the end of each
time of 2ms. This block consists of the CAN transmit subsystems which send 𝑢∗ to the
ES910 over a 1Mbit/s J1939 CAN bus. The same channel is used to receive states 𝑥 and
parameters 𝑘 from the ES910. These are utilized to execute the next step of the MPC
program. During testing, the message transmission and receive time was observed to be
less than 2ms which is negligible compared to the unit-step delay induced from discrete
time execution.
simulation controller. Because of the additional actuators and sensors associated with LP-
EGR and WG control an ES930 I/O extension unit has also been added to the ES910
controller. The 𝑃𝐸𝑔𝑟 measurement is also performed using an Oxygen sensor controller
box which outputs analog voltage as a function of Oxygen concentration. The lower
level control layer also has the position controllers for the LP-EGR, PCT and BOV
89
valves. These are conventional PID based controllers. Because of the faster dynamics of
the lower-level actuators, the sample time of the low- level controller and estimator is set
to 1ms.
Figure VII-1. Rapid control prototyping setup for experimental validation of MPC
90
7.2.1 Experimental test results 2900 RPM: Aggressive tip-in and tip-out maneuvers
open-loop fuel-economical control input 𝑢𝑅𝑒𝑓 . The analysis is done for step commands of
𝐼𝑀𝐸𝑃𝑛 at 3000 RPM to show the tip-in and tip-out response of both the control
approaches. Figure VII-2 shows the tip-in response of the controllers. It is evident that the
tip-in response of the MPC is much quicker than the open-loop control. The MPC
response also has a slight overshoot followed by stable steady state tracking. The air-path
actuator trajectories have been shown in Figure VII-3, Figure VII-4 and Figure VII-5.
Figure VII-2. Tip-in IMEPn response for MPC and Open-loop control strategy
At the instant of the tip-in the commanded reference 𝑚̇𝐼𝑇𝑉 is momentarily increased and
dropped based on the reference steady state intake throttle position which corresponded
to a pressure difference of approximately 15kPa across the intake throttle. Because of this
91
the MPC commanded 𝑚̇𝐼𝑇𝑉 impulse is slightly higher as the forward throttle flow
constraint is inactive in the reference trajectory. MPC completely stops 𝑚̇𝑊𝐺 and 𝑚̇𝐸𝐺𝑅
to accelerate turbocharger spool-up. 𝑚̇𝐸𝐺𝑅 converges to the reference value in the latter
part of the transient. The cylinder inputs for both the control strategies is shown in Figure
VII-6 and Figure VII-7. MPC retards the spark timing and advance the intake cam timing
Figure VII-3. Intake throttle mass flow rate trajectories for MPC and Open-loop control
Figure VII-4. Wastegate mass flow rate trajectories for MPC and Open-loop control
92
Figure VII-5. External cooled EGR mass flow rate trajectories for MPC and Open-loop
control shows that MPC deliberately stops EGR flow on the aggressive tip-in
Similar to the air-path actuators, the deviation of the MPC inputs from the
reference input is mainly to improve the turbocharger spool-up. Figure VII-8, Figure
VII-9 and Figure VII-10 shows the modeled volumetric efficiency, measured exhaust
temperature and measured turbocharger speed for both of the control strategies. The
advanced intake cam timing is clearly to improve the transient volumetric efficiency of
the cylinders whereas the retarded spark timing is to increase the exhaust temperature.
93
Figure VII-6. Spark timing trajectories for MPC and Open-loop control
Figure VII-7. Camshaft position timing trajectories for MPC and Open-loop control
94
Figure VII-8 ANN Modeled volumetric efficiency using MPC inputs and open-loop
Figure VII-9 Measured turbine inlet temperature for MPC and Open-loop control shows
that MPC commands late spark timing to increase specific exhaust gas enthalpy for
turbocharger response
95
Figure VII-10 Turbocharger speed for MPC increases much more rapidly than open-loop
control
significantly faster than the open-loop control. It is noteworthy, that once the IMEPn
tracking has reached steady state, the deviation between the MPC and the open-loop
inputs is minimal.
The most relevant constraint for the tip-in maneuver is the 𝐾𝐼 2 constraint. The
MPC advances the spark timing well beyond the reference value between 3 and 4
seconds. However, MPC also commands 𝑚̇𝐸𝐺𝑅 beyond the reference value as well during
this time period. The measured 𝐾𝐼 2 for cylinder 1 is shown in Figure VII-11. Due to the
spark timing advancement, the 𝐾𝐼 2 value increases well beyond the open-loop control
trajectory. However, due to the addition of extra 𝑚̇𝐸𝐺𝑅 into the engine, the constraint is
not violated.
96
Figure.
Figure VII-11 Measured 𝐾𝐼 2 for MPC does not exceed the limit of 1 despite advanced
Because the open-loop control strategy does not have active compressor surge handling
ability, only the MPC trajectories are shown. Figure VII-12 shows that there is a slight
97
Figure VII-12. Tip-out IMEPn response for MPC and Open-loop control strategy
The air path trajectories for the tip-out maneuver are shown in Figure VII-13. The MPC
commands the intake throttle mass flow rate to zero to reduce the intake manifold
pressure. MPC commands stoppage of EGR mass flow rate for combustion stability
concerns. Simultaneously the MPC commands an increase in waste-gate mass flow rate
and blow off valve mass flow rate to address the tip-out surge problem. Figure VII-15
shows that even though the MPC commands the blow off valve to open instantaneously,
the compressor operation actually crosses over into the surge region. This is mainly
because, the boosted engine operation before the tip-out is very close to the surge line. At
the instant of the tip-out the blow off valve command issued by the MPC cannot be
instantly fulfilled by the lower level controller due to bandwidth limitations. The true
compressor mass flow rate dynamics during the surge is also governed by very fast limit-
cycle like behavior which can be modeled using the classical Moore-Greitzer model
98
mentioned in [53]. However, addition of this model introduces dynamics which are
significantly faster than the MPC sample time to and should effectively be used in the
Apart from compressor surge avoidance, tip-out dilution from trapped external
EGR in the air-path is a concern as well. Since this EGR cannot be evacuated by any
other means other than by passage through the engine itself, the MPC can only control
the engine actuators to handle this problem. It is visible from Figure VII-14 that MPC
combustion stability. Although, internal residual gas fraction is not included in the
control model, the COV model indirectly captures the effect of excessive valve overlap
on COV which is the most likely reason for valve timing manipulation at the instant of
tip-out.
99
Figure VII-13 Air-path trajetories for aggressive tip-out maneuver
100
Figure VII-15 Compressor operation during the aggressive tip-out
101
VIII. CONCLUSIONS AND RESEARCH CONTRIBUTIONS
was performed in order to identify the lapses in research and potential for improvements
open research topic. The recent research work published for MPC based subsystem
control of turbocharged engines was examined in detail and it was quickly identified that
majority of research articles utilize linear system models which are either physics based
or of black-box nature and derived offline. Although this approach works well for air-
path control, is questionable when engine cylinder actuators are also included in the MPC
framework. Hence, in order to assess the error associated with the linearity approximation
a control oriented physics based model of the turbocharged engine air-path was
developed with ANN based sub-models for outputs and constraints. A study was
performed to evaluate the deviation of the non-linear model from the linearized model
over the prediction horizon which revealed that the linear approximation error is not
trivial for the highly non-linear and the most important outputs of the engine, IMEPn.
Extending this study further, a simple analysis was performed to evaluate the number of
re-linearizations required within the prediction horizon for a given required accuracy
level and operating point. The results of this analysis show that the model has varying
levels of linear approximation error depending on the operating condition at which the
analysis is done and the frequency of the reference inputs applied to the model.
102
Under the knowledge that the engine model is non-linear and MPC strategies
require linearization of the model, avenues to speed up the linearization process have
been explored. Analytical differentiation has been widely utilized to reduce the number
of computations associated with model linearization. The resulting approach adopted has
This is a significant advantage with cascaded benefits especially when the linearization is
done more frequently within the prediction horizon. Based on this approach, an NMPC
where linearization is done at every stage within the prediction horizon is formulated. A
preliminary co-simulation with 1D-Engine simulation has been performed and the results
have demonstrated the ability of the control system to coordinate air-path and engine
cylinder actuators simultaneously while tracking the desired IMEPn within 1 bar.
and turbocharger speed estimation layer was developed. Simple inversion of the orifice
flow model was performed with an additional adaptive filter to mitigate chattering of the
intake throttle control at very low pressure differential values. A fixed-point iteration
based turbocharger speed estimation algorithm was utilized to estimate the turbocharger
rotational speed using the production sensor set. A two micro-controller based rapid-
was established between the CAN controllers. Using this layout, real-time experimental
validation of the NMPC algorithm on a 2.0 L Turbocharged gasoline engine with cooled
103
external EGR was performed. In the absence of a production type engine control
algorithm for this engine configuration, the NMPC performance was evaluated against
the reference open loop control which is optimized for fuel economy. The NMPC
performs control actuator maneuvers similar to those observed using the GT-Power co-
simulation. The NMPC demonstrates the ability to manipulate Spark and cam timing to
improve engine breathing and exhaust gas enthalpy to maximize turbocharger response
while simultaneously controlling the air-path actuators. However, the surge control has
pitfalls due to the proximity of the compressor operation to the surge line and the
bandwidth limitations of the lower level control which were unaccounted for in the GT-
Power co-simulation.
The algorithm outlined in this work is derived mainly from a physics based air-
path model coupled with ANN based engine model derived from steady state data. These
models were derived using 40 hours of steady state engine dynamometer test data. This
time period is negligible compared to the time required to acquire transient data to tune
black-box models of the engine. Furthermore, the physics based nature of the air-path
implies that changes in air-path geometry and turbocharger hardware can be quickly
implemented in the model without the need to re-identify parameters associated with air
path dynamics. These attributes of this approach imply that the algorithm can be ported to
104
Due to demonstrated ability of the algorithm to co-ordinate multiple actuators
simultaneously, calibration efforts required are minimal with only 15 constants which
require tuning, which is negligible compared to the large number of lookup tables and
formulated based on the current state of the system, user defined objective function and
the reference inputs applied to the system. Although an MPC-QP is convex due to
linearized models used in the formulation, the original Non-linear optimization problem
can be non-convex. As a result, applying the sub-optimal control inputs derived from the
MPC-QP optimization may violate one or more constraints. This approach was used in
this research because 2 out of the 5 constraints on the mainly to impose physical
feasibility in the control model and can never be violated in reality. The remaining 3
operational constraints could possibly be violated with this approach. In order to find sub-
computational resources.
The tuning of the NMPC is also an open question, wherein the duality between Q
and R implies that IMEPn tracking affects the cost associated with tracking of the
reference inputs and vice versa. Drive cycle or extended transient simulation could be
105
performed with different tuning values to study impact on fuel-economy. Improved
transient turbocharger response may not necessarily imply reduced fuel economy, as
106
A. APPENDIX
A Control Algorithm for Low Pressure – EGR Systems using a Smith Predictor with
gasoline engines has been shown to improve efficiency at many engine operating
conditions [1][2][3]. Previous research work in the domain of LP-cEGR control focuses
mainly on utilization of mean value models of air and EGR paths for determination of
EGR and air fraction [12]. An open-loop EGR valve mass flow model is utilized in these
methods [4][5]. The primary challenges associated with control of LP-cEGR systems are;
(1) mass flow modeling across EGR valve due to low-pressure differential across the
system [4][5], and (2) EGR control is difficult during transients due to long length of
EGR path [4][5][12]. Additionally, the methods described in [4] and [12] also utilize an
intake throttle valve upstream of the EGR mixing location, which increases pressure
differential across the EGR system. Additional throttle valves for EGR control are not
Universal Exhaust Gas Oxygen (UEGO) sensors are widely used in engine exhaust
systems for control and diagnostics. These sensors measure the concentration of oxygen
in the sampled gas. Previous research has demonstrated the feasibility of using UEGO
107
sensors to accurately measure the EGR fraction in the intake system [7][8]. The potential
of these sensors to provide real-time EGR fraction feedback control is explored in this
research. Water condensation from exhaust [1] can severely damage the high temperature
sensing element in these sensors. Also, in order to ensure proper EGR and Air mixing at
the sensor location, the sensor location has to be adequately downstream of the mixing
location. However, moving the sensor farther downstream of the EGR valve increases the
EGR valve to sensor time delay. This severely restricts the calibrated ‘aggressiveness’ of
the feedback control system. In systems with dead time delays, Smith Predictor based
feedback control is a widely used method for process control [13][6]. Smith Predictor
based control is a predictive method of process control where the measured output of the
process is delayed by a known time duration. Fluid flow through a pipe and engine
The paper begins with description of experimental setup including the engine and control
system hardware. As required by the Smith Predictor method, models are defined for the
EGR and air path. These consist of open-loop modeling of EGR valve mass flow and the
transport delay across the EGR and air path. These models are validated in isolation to
evaluate their accuracy. Finally, the complete control structure is tested at steady state
controls engine speed and/or loading for all experiments. A set of production-intent
sensors is installed on the engine for basic engine functionality (air, fueling, ignition,
108
valve timing and turbo control). A low-pressure and cooled EGR (LP-cEGR)
configuration is retrofitted on the engine. Exhaust gases are extracted downstream of the
turbine and upstream of the catalyst. EGR passes through a cooler and is delivered to the
intake air-path system upstream of the compressor. The EGR cooler is a tube-core type
chosen for low pressure loss characteristics. Twin liquid-to-air intercoolers have been
used to allow high boost/load capability in the dynamometer cell, where air/air heat
SI
Displaced volume 1998 cc
Bore x Stroke 86 x 86 mm
109
The overall EGR control system architecture is outlined in Figure 2, and a description of
include software hooks on specific engine control parameters. An ETAS ES910 Rapid
Prototype Controller (RPC) is used in conjunction with the base engine controller using
data interface between the base engine controller and the rapid prototype controller. A
variant of an oxygen sensor is also installed post compressor (Figure 1.). This sensor is
similar in operation to UEGO sensors, however it has been developed and optimized to
be operated specifically in the intake environment, where temperature range and gas
composition is different from the exhaust. EGR fraction is calculated based on change in
intake oxygen concentration caused by EGR dilution [7][8]. Hereafter, the intake oxygen
Base
engine
controller
……
Engine
Additional Sensors
CCP bypass
and actuators
RPC EGR
Controller
110
𝑚̇𝐴𝑖𝑟 Fresh air mass flow measured before EGR mixing
𝑃𝐶𝑜𝑚𝑝−𝐼𝑛 Compressor inlet pressure measured at same
𝑃𝐶𝑜𝑚𝑝−𝑂𝑢𝑡 Boost pressure measured post intercooler
𝑇𝐴𝑖𝑟 location as 𝑚̇𝐴𝑖𝑟
Inlet air temperature measured at 𝑚̇𝐴𝑖𝑟 location
𝑁𝐸𝑛𝑔 Engine speed [rpm]
𝑇𝐸𝐺𝑅 EGR cooler outlet exhaust temperature
𝑥𝑂2𝑆𝑒𝑛𝑠 Mole fraction of Oxygen measured by EGR sensor
Mass flow rate across the EGR valve is modeled using an isothermal orifice flow model
in equation (76). This model is specifically used for exhaust gases at lower temperatures
83.
𝐶𝐷 is the discharge coefficient and 𝐴𝑉𝑎𝑙𝑣𝑒 is the valve open area. 𝑃𝐼𝑛 is the valve inlet
pressure. To determine pressure at the inlet of the EGR valve, the pressure at the inlet of
the EGR cooler and the pressure drop across the EGR cooler needs to be known.
Calculation of pressure drop across the EGR cooler requires prior knowledge of mass
flow across the EGR system. For the sake of simplicity, this research considers EGR
valve inlet pressure to be the same as EGR cooler inlet pressure. A physics based model
is utilized to predict this pressure as a function of engine operating conditions [68]. 𝑃𝑂𝑢𝑡
111
is the valve outlet pressure which is considered to be the same as 𝑃𝐶𝑜𝑚𝑝−𝐼𝑛 . 𝑅 is the
specific gas constant for stoichiometric exhaust gas. LP-cEGR systems are required to
pressure ratio correction is very sensitive to minor inaccuracies in the pressure ratio at
such low pressure differences. In addition, the pressure at the inlet of the EGR valve
pulsates[4][5], which influences the flow through the EGR system. The model shown in
(83) holds strictly true only for steady flow conditions. Pressure pulsations reduce the
accuracy of the model as the flow is unsteady at the valve and averaged pressures are
In this research the discharge coefficient used in the model has been characterized as a
function of engine speed and EGR valve angle. This is done to partially capture the effect
of the primary frequency of pressure pulsations on the discharge coefficient since the
speed [11]. The discharge coefficients were obtained by running an optimization routine
which minimizes the error between the modeled mass flow and the measured mass flow.
The measured mass flow is derived from experimental sweeps of EGR valve position
between fully closed and fully open at different engine operating conditions. The engine
load was swept from 3 to 11 Bar BMEP and theengine speed was swept from 1200 to
2700 RPM. The final result of this optimization routine is the two-dimensional discharge
coefficient map shown in Figure A-2. Typical maximum values of discharge coefficient
for butterfly type valves are close to 0.6 [11]. The values obtained for this research were
signficantly lower (0.15) mainly due to consideration of EGR cooler inlet pressure as
112
valve inlet pressure. A minor inaccuracy in the mapping between EGR valve angle and
Modeled and measured mass flow rates along with the +/-10% error lines are
shown in Figure A-3. The majority of operating points lie within the +/-10% range.
Engine speed has been used to capture the effect of exhaust pressure pulsations. The
primary frequency of exhaust pressure pulsations is a linear function of engine speed and
number of cylinders/exhaust blowdown events per cycle. This has been done to partially
characterize the effect of exhaust pressure pulsation frequency on the EGR mass flow.
Effects of peak amplitudes and pulsation shape may also have an impact, but are outside
0.15
Discharge coefficient [-]
0.1
0.1
0.05
0.05
80
70
0
60
2700
50 2500
40 2300
2000
30 1800
EGR valve angle [deg] 20 1500 Engine speed [rpm]
1200
Figure A-2 Dimensional Discharge coefficient map obtained from optimization routine
113
Figure A-3 Measured vs modeled EGR valve mass flow rate
The dead-time delay between EGR valve and the EGR sensor consists of two
components. The first is transport delay from the EGR mixing location to EGR sensor.
The second component is sensor response time, consisting of transport from exterior of
sensor to sensing element and electrical signal delay. The sensor response time is
approximated as a linear function of flow velocity in the control volume in which the
sensor is mounted. A simplified equation based on the assumption of plug flow is derived
using the steady state continuity equation for mass flow rate through a control volume.
114
This method is utilized to estimate the transport delay from the entry to the exit of the
control volume. The equation used to estimate the transport delay 𝜏𝑖 across a generic
In this equation 𝐿𝑖 and 𝐴𝑖 are the length and flow area respectively. 𝑚̇𝑖 , 𝑃𝑖 and 𝑇𝑖 are the
mass flow rate, pressure and temperature in the control volume respectively. Although
the equation is derived for strictly steady state conditions, it is still used for the quasi
steady state operating conditions tested in this research where only EGR valve transients
were performed.
EGR fraction model with no transport delay and the EGR fraction model with modeled
transport delay. These models are described as per the schematic in Figure A-4.
115
The EGR fraction model with no transport delay is used to predict the EGR
fraction at mixing point A. Under the assumption that the mixing is instantaneous and
𝑚̇𝐸𝐺𝑅
𝐸𝐺𝑅𝑓𝐴 = 𝑚̇ 85
𝐴𝑖𝑟 +𝑚̇𝐸𝐺𝑅
The EGR fraction at location B and time, 𝑡, 𝐸𝐺𝑅𝑓𝐵 (𝑡)which is also the EGR
sensor location is the transport delayed value of the EGR fraction at A. The modeled
Where 𝜏(𝑡)is the modeled transport delay at time 𝑡 from location A to location B,
Where 𝜏𝐴𝑡𝑜𝐶 (𝑡) and 𝜏𝐶𝑡𝑜𝐵 (𝑡) are the modeled transport delays from A to the
compressor and compressor to B respectively. 𝜏𝐶 (𝑡) is the transport delay within the
compressor which has been assumed to be zero. These delays are calculated separately
because the pressure and temperature within the control volumes from A to the
116
EGR fraction measurement using Oxygen sensing
The EGR mole fraction, 𝐸𝐺𝑅𝑓𝐵 𝑆𝑒𝑛𝑠 , is the measurement performed by the EGR
sensor. It is calculated using the measured mole fraction of oxygen 𝑥𝑂2 𝑆𝑒𝑛𝑠 with equation
(88).
𝑥𝑂2 −𝑥𝑂2
𝐴𝑚𝑏 𝑆𝑒𝑛𝑠
𝐸𝐺𝑅𝑓𝐵 𝑆𝑒𝑛𝑠 = 88
𝑥𝑂2 −𝑥𝑂2
𝐴𝑚𝑏 𝐸𝑥ℎ
𝑥𝑂2 𝐴𝑚𝑏 and 𝑥𝑂2 𝐸𝑥ℎ are the mole fractions of oxygen in fresh air and the exhaust gases
respectively. In order to derive the EGR mass fraction, the following assumptions are
made:
The difference between EGR mass fraction and EGR mole fraction is considered
A classical Smith Predictor was implemented as a dead time delay compensator. Previous
research shows that Smith Predictor based classical feedback control for EGR systems in
Large Diesel Engines improves control stability [61]. The application of the Smith
predictor is visible in the EGR fraction controller shown in Figure A-5. Linear,
117
Figure A-5. Closed loop EGR fraction controller with Smith predictor
𝑃𝑃𝑟𝑒𝑑𝐴 (𝑠) = 𝑃𝑀𝑜𝑑𝑒𝑙𝐴 (𝑠) + (𝑃𝑃𝑙𝑎𝑛𝑡 𝐴 . 𝑒 −𝜏𝑃𝑙𝑎𝑛𝑡 (𝑡)𝑠 − 𝑃𝑀𝑜𝑑𝑒𝑙𝐴 (𝑠). 𝑒 −𝜏(𝑡)𝑠 ) 90
The primary objective of this architecture is to eliminate the delay dynamics from the
output. The EGR Path + Sensor subsystem system has the output 𝐸𝐺𝑅𝑓𝐵 𝑆𝑒𝑛𝑠 which is
inherently delayed by the physical transport delay 𝜏𝑃𝑙𝑎𝑛𝑡 (𝑡) as shown in equation (89)
𝑃𝑃𝑙𝑎𝑛𝑡𝐴 (𝑠) is the EGR Path + Sensor subsystem without the transport delay. The output
of augmented system approaches that of the Mixing point EGR fraction model as the
term in the parenthesis in equation (90) tends to 0 i.e. as the output of 𝑃𝑀𝑜𝑑𝑒𝑙 𝐴 (𝑠) tends
to 𝑃𝑃𝑙𝑎𝑛𝑡 𝐴 (𝑠) and 𝜏(𝑡) to𝜏𝑃𝑙𝑎𝑛𝑡 (𝑡). This emphasizes the importance of accuracy of the
EGR flow model and transport delay model. Effectively, the output 𝐸𝐺𝑅𝑓𝐴 𝑃𝑟𝑒𝑑 of the
118
augmented system𝑃𝑃𝑟𝑒𝑑 𝐴 (𝑠) has minimal delay dynamics in it. This reconstructed output
along with the desired EGR fraction command 𝐸𝐺𝑅𝑓𝐵 𝐶𝑜𝑚𝑚 is used to generate the error
upon which the PID controller operates. Therefore, instability and oscillation induced by
The benefits of LP-cEGR are seen at different engine operating conditions due to
different mechanisms [36]. This research mainly focuses on LP-C-EGR control at low to
mid load. This is done partially due to the generally longer transport delay 𝜏(𝑡)associated
with these load conditions. Longer transport delays can exist for high load and very low
engine speeds but for this research these conditions were not considered.
The transport delay model was tested experimentally by performing step changes in EGR
valve position at quasi steady engine operating conditions. These conditions were
intake throttle valve, valve timing, waste-gate and blow-off valve position, etc. The
number of engine cycles elapsed between the EGR valve step and the effect of the step at
the EGR sensor was compared with the delay calculated by the transport delay model.
This is done at various engine operating conditions to account for varying engine cycle
119
time. The majority of test points lie within a +/- 1 engine cycle band, as shown in Error!
Recall that 𝐸𝐺𝑅𝑓𝐵 is the transport delayed version of 𝐸𝐺𝑅𝑓𝐴 using the transport delay
model. In order to evaluate accuracy of the EGR fraction models EGR valve steps were
performed, and modeled and measured EGR fractions at the EGR sensor location are
compared Figure A-7 to Figure A-9. show this comparison at various engine operating
conditions for different EGR valve position steps. At all of the operating conditions, the
fast change in 𝐸𝐺𝑅𝑓𝐵 and 𝐸𝐺𝑅𝑓𝐵 𝑆𝑒𝑛𝑠 caused by the EGR valve position step are within
+/- 1 engine cycle. The steady state difference between 𝐸𝐺𝑅𝑓𝐵 and 𝐸𝐺𝑅𝑓𝐵 𝑆𝑒𝑛𝑠 is highest
120
at 2100 RPM and 2.5 Bar BMEP. This is attributed to the inaccuracy of the EGR mass
Figure A-7. Comparison between modeled and measured EGR fraction at EGR sensor
121
Figure A-8. Comparison between modeled and measured EGR fraction at EGR sensor
122
Figure A-9. Comparison between modeled and measured EGR fraction at EGR sensor
In order to test the feedback control architecture outlined in Figure A-5, step inputs are
commanded for desired EGR fraction. These tests are performed using conventional PID
control in Figure 11, Figure 13 and Figure 15. Step response of EGR fraction controller
with conventional PID controller at 2000RPM 9 Bar BMEP as well as Smith Predictor
based control as seen in Figure 12, Figure 14 and Figure 16. Step response of EGR
fraction controller with Smith Predictor at 2000RPM 9 Bar BMEP The controller gains
Kp, Ki and Kd were determined heuristically as 130, 6 and 0.01 respectively by step
123
response analysis. They were kept constant for all of the tests. In the case of the
The EGR valve position overshoots the position required to reach the desired
EGR fraction
The EGR valve position appears to be 90 degrees out of phase with the sensed
Oscillations can be mitigated with appropriate controller tuning. However, the controller
gains would also have to be scheduled as per engine operating condition as the flow
within the EGR and air path is mainly dependent on engine operating condition.
Figure A-10. Step response of EGR fraction controller with conventional PID controller
124
Figure A-11. Step response of EGR fraction controller with Smith Predictor at 1500RPM
7 Bar BMEP
In case of the Smith Predictor based controller the following observations are made:
Oscillations related to phase lag between EGR valve and sensor are greatly
Rise time and fall time (time duration from 10% to 90% of the setpoint value) are
125
There are small overshoots in the EGR valve position, but they are small enough in
magnitude and time to not affect EGR fraction significantly, as evident in Figure A-13 at
3 and 27 seconds. There are some steady state oscillations in EGR fraction as well. These
were determined to be a result of a ‘sticky’ EGR valve and controller combination. The
increase in rise and fall time are mainly due to the difference between modeled and
measured EGR fractions. The comparison between the performance criteria of these two
Figure A-12. Step response of EGR fraction controller with conventional PID controller
126
Figure A-13. Step response of EGR fraction controller with Smith Predictor
127
Figure A-14. Step response of EGR fraction controller with conventional PID controller
Figure A-15. Step response of EGR fraction controller with Smith Predictor at 2000RPM
9 Bar BMEP
128
Table A-1 Conventional PID control vs Smith Predictor Control performance comparison
Summary/Conclusions
A closed loop LP-cEGR feedback controller with Smith predictor based dead time delay
compensator was implemented and tested. Open loop models for EGR fraction at the
mixing and EGR sensor locations along with a transport delay model were developed and
129
validated. EGR valve step changes were performed to evaluate step response of a
The Smith Predictor control showed significant reduction in average overshoot as well as
improved steady state stability. However, due to mismatch between EGR sensor feedback
and modeled EGR fraction, rise time and fall time increased as compared to PID control.
Accurate open loop modeling of EGR mass flow in LP-cEGR systems is still an area
where development is required. The conventional orifice flow model begins to lose
accuracy at low pressure differentials and with pressure pulsations. These conditions
130
Nonlinear Model Predictive Control of Dual Loop - Exhaust Gas Recirculation in a
Compression Ignition (CI) engines for its ability to reduce NOx emission [69]. This
technology was introduced to Spark Ignition (SI) engines recently due to its ability to
mitigate knock and reduce pumping loss[11]. However, SI engines require much more
precise EGR concentration control than the CI engines. The employment of High Pressure
(HP) and Low Pressure (LP) dual loop (DL) EGR provides more degrees-of-freedom for
EGR concentration control. The LP-EGR loop has a characteristic transport delay much
longer than the HP-EGR loop, which introduces additional difficulty in air fraction control
in the intake manifold. Conversely, exhaust enthalpy exchange through the turbine is
reduced in the case of HP-EGR thereby reducing turbine power and hence adversely
affecting compressor mass flow rate. It is desirable to coordinate the dual loop EGR
systems to achieve given EGR concentration target and maximize turbocharger efficiency.
In this work, a Nonlinear MPC (NMPC) is proposed to control the intake manifold pressure
manipulating the throttle, EGR valves of HP and LP loops. For a given preview horizon of
tacking references, the NMPC minimizes tracking error and control effort subject to the
The control of turbocharged gasoline engine air handling system with EGR is a
Multi-Input Multi-Output (MIMO) problem with nonlinear and coupled dynamics. NMPC
has been utilized in [93], to track torque for a turbocharged SI engine where the throttle
131
and wastegate are considered as the control inputs. External EGR control has not been
included in the aforementioned research. While references for the DL-EGR control of SI
engine are limited, the control of the diesel engine air system with dual loop EGR has been
extensively discussed in literature [69] and [94], where the authors successfully
implemented a switched linear MPC with feed-forward and feedback controllers for the
multi-input multi-output nonlinear DL EGR and VGT system. In these papers, multiple
linear models were developed over engine speed and fueling operating regions and a
switched linear MPC approach with quadratic cost function was employed to calculate
control actions.
The EGR control strategy developed for CI engines may not be suitable for SI
engines. The LP EGR transport dynamics considered in previous CI engine research are
highly simplified and usually similar to a first order delay. This approximation may be
tolerable for CI engine operation as the ignition process is more robust against EGR
dilution. Such an assumption for a SI engine can potentially lead to undesirable operation
like misfire and partial combustion due to over dilution of EGR. Hence, it is necessary to
improve the accuracy of transport delay dynamics in the control oriented model of the
airpath as shown in Figure A-16, wherein a NMPC has been utilized with a segmented
boost manifold based transport delay model integrated into the airpath model. This research
employs a similar approach in which a high order transport delay model is utilized to
improve the prediction of LP-EGR fraction from the location of its delivery to the engine
cylinders.
132
Control oriented air path model
The individual components of the air path have been modeled separately and
ultimately integrated to form a control oriented air path as depicted in Error! Reference
Here, 𝑃, 𝑇, 𝐹, 𝑚̇, 𝜔, and 𝑉 are pressure, temperature, mass fraction, mass flow rate,
rotational speed and volume respectively. Subscripts𝐸𝑥ℎ, 𝐵𝑠𝑡, 𝐼𝑚, 𝐻𝑃𝐸, 𝐿𝑃𝐸, 𝐼𝑇𝑉,
𝐶𝑜𝑚𝑝, 𝑇𝑢𝑟𝑏, 𝐼and𝐸 are the exhaust manifold, boost manifold (between compressor and
compressor, turbine, ambient compressor inlet and post-turbine exhaust respectively. The
133
following section is dedicated to description of the component level models used in the air
path model.
The boost manifold is typically the largest control volume in a turbocharged engine air path
and hence consideration of the gas transport dynamics in the boost manifold is critical for
transient control of EGR dilution at the cylinders. This manifold has been traditionally
considered as a lumped volume in the turbocharged diesel engine air path. However, in
typical turbocharged automotive engines the boost manifold is characterized more by thin
long pipes rather than plenums. Therefore, a lumped volume based assumption for the
boost manifold is done at the expense of accuracy. Experimental testing has confirmed the
above mentioned phenomenon as shown in Figure 4 where a step input in EGR valve
position results in a dead time delayed and mixing influenced response from the EGR
measurement sensor installed at the throttle valve. The mixing influence cannot be
attributed purely to mixing in the manifold but also to the filling and emptying dynamics
134
Figure A-17. Step response of oxygen sensor based EGR measurement installed at
The consequence of excessive EGR dilution in diesel engines typically results in increased
smoke and soot emission. However, with excessive EGR dilution, gasoline spark ignited
engines can suffer from very high cyclic variability in torque production. Furthermore, the
three way catalyst can encounter premature damage due to misfires caused by excessive
dilution. The gas transport model chosen here (91) is based on conservation of mass and
𝑅𝑇𝑖 91
𝐹𝑖̇ = (𝐹 𝑚̇ − 𝐹𝑖 𝑚̇𝑂𝑢𝑡 )
𝑃𝑖 𝑉𝑖 𝐼𝑛 𝐼𝑛
where, 𝐹𝑖 is the EGR fraction in the 𝑖 𝑡ℎ control volume 𝑉𝑖 . The number of control volumes
utilized to discretize the boost manifold is the same as the number of fraction states 𝐹𝑖 used
in the transport model. Hence, the order of the transport model is equal to the number of
135
Figure A-18. Comparison of step response of transport delay models with true dead time
delay response
Higher order models show increasing accuracy and similarity to the ‘true’ dead time
delayed response (if pure pipe plug flow is considered) as shown in Figure A-18. A 5th
order transport delay model was chosen for this research as it has significantly higher
accuracy than the 1st order model which is a single lumped volume for the entire boost
136
State space model and control problem formulation
Using the component level models described in the previous section, iso-thermal
pressure dynamics and turbocharger rotor dynamics described in [14] the discrete time
state-space model is derived as given below. Since the goal of this research was to evaluate
control of mass flow rates of the EGR valves, temperature dynamics were ignored to keep
system dimensionality low and allow for more transport delay states.
The output vector 𝑦consists of the intake manifold pressure 𝑃𝐼𝑚 and intake
manifold EGR%𝐹𝐼𝑚 . For gasoline SI engines the intake manifold pressure and EGR%
strongly associated with the engine load and these references are assumed to be available
from an external source (eg. Torque management system, supervisory control). The
remaining states 𝑃𝐵𝑠𝑡 , 𝜔 𝑇𝑢𝑟𝑏𝑜 , 𝐹𝑖 and 𝑚̇𝐿𝑃𝐸 are the compressor outlet pressure,
turbocharger rotational speed, the transport delay states and the LP-EGR mass flow rate
respectively. The input vector 𝑢 consists of the intake throttle and high pressure mass flow
rate 𝑚̇𝐼𝑇𝑉 and 𝑚̇𝐻𝑃𝐸 . The LP-EGR mass flow rate is controlled using input 𝑚̈𝐿𝑃𝐸 , as this
137
allows tuning of the rate of change of the LP-EGR mass flow rate. The model is also
The objective of the control algorithm is to minimize the squared error between a reference
value 𝑦𝑅𝑒𝑓 (𝑘) and the system output 𝑦(𝑘) for the preview horizon. Hence, the optimal
control problem has been designed to be an output tracking controller with the cost function
defined as follows.
where,
3×3
ℝ2×2
>0 , 𝑅 ∈ ℝ>0
𝑄 and𝐻 are the tuning matrices for the penalty on output tracking error within the
prediction horizon and the terminal output error. The tuning of these matrices was done to
place more emphasis on the tracking of 𝐹𝐼𝑚 as the consequence of poor tracking of this
parameter can result in misfires. 𝑅is the tuning matrix for penalty on magnitude of input
min
𝐽(𝑥(𝑘), 𝑈(𝑘))
𝑈(𝑘)
𝑚̇𝐼𝑇𝑉 (𝑖) ≥ 0,
𝑚̇𝐻𝑃𝐸 (𝑖) ≥ 0,
138
𝜔 𝑇𝑢𝑟𝑏𝑜 (𝑖) ≥ 0.
where, 𝑖 = 𝑘, 𝑘 + 1. . , 𝑘 + 𝑁 − 1
All of the constraints are imposed to maintain physical feasibility in the system.
The turbocharger rotational speed has the longest time constant in the system (1.1 seconds)
Hence, for a sample time of 0.1 seconds the horizon length 𝑁 is 11. The numerical values
chosen in the tuning matrices are chosen heuristically with more emphasis on the EGR%
control. Fig. 6 shows the overall schematic representation of Simulation results and
discussion the NMPC control structure. The NMPC problem was simulated using ACADO
toolkit [95].
139
The simulation was performed to study the ability of NMPC to handle the following
The intake manifold pressure reference trajectory was deliberately chosen to have
boosted and un-boosted values so that the interaction between the EGR circuits and the
turbocharger performance could be studied. Arbitrary time varying references were chosen
for EGR%. The simulation was performed for three engine speeds, 1000, 2000 and 3000
RPM with the same reference trajectories. The speeds were chosen because external cooled
EGR has maximum benefit in knock mitigation at low speed high load conditions. In the
It is evident from Figure A-20 and Figure A-21 that in this case the system is able
to track the reference of the intake manifold mass better than the low pressure only and
dual loop cases. However, the intake manifold pressure reference tracking under a
reference value above atmospheric pressure is the poorest in this case as seen between 5
and 7 seconds. This is because of the reduction of mass flow rate through the turbine caused
by the high-pressure EGR system. The HP-EGR circuit effectively behaves like a waste-
gate and hence results in reduced turbine power. As per the perception of the driver, this
140
would seem as poor engine torque response. The controller overshoots the intake throttle
and HP-EGR mass flow rates whenever there is a sudden change in reference. Another
thing to note is that the boost pressure always stays above or equal to the intake manifold
Figure A-20 Pressure states and intake manifold EGR dilution trajectories for HP-EGR
control only.
141
Figure A-21 Input trajectories for HP-EGR control only.
142
Low-pressure EGR control only
The most significant difference between this case and the HP-EGR case is that the intake
manifold pressure reference tracking is significantly better as seen in Figure A-22 and
Figure A-23. as all of the exhaust mass flow passes through the turbine resulting in higher
turbine power to drive the compressor to higher boost levels. However, the reference
tracking for EGR mass in the intake manifold is worse with oscillation at the EGR%
reference changes at 5 and 8 seconds. Due to the transport delay model being
incorporated into the NMPC, there is also some lead action visible at 0.5 seconds on the
LP-EGR mass flow rate to meet the EGR% reference change at 1 second. The LP-EGR
mass flow rate has an immediate effect on only the EGR% in the first section of the boost
manifold which is far upstream of the intake manifold. The dilution dynamics in every
consequent air-path section diminishes the effect of this input on the intake manifold
EGR%. Hence, the tuning weight in matrix R associated with the rate of change of LP-
EGR mass flow rate had to be modified to reduce oscillations in the LP-EGR mass flow
rate.
143
Figure A-22 Pressure states and intake manifold EGR dilution trajectories for LP-EGR
control only.
144
Figure A-23 Input trajectories for LP-EGR control only.
145
Dual-loop EGR control
In this case the intake manifold pressure and EGR mass reference tracking appear
to be reasonably better compared to the LP-EGR only case as shown in Figure 9a. and 9b.
Since the engine is operated throttled up to 5 seconds the NMPC commands only HP-
EGR up to 4 seconds. At the transition from throttled to boosted operation LP-EGR flow
is initiated to maximize turbine power. As the intake manifold pressure level increases
due to turbo spool-up, a sharp HP-EGR mass flow rate spike is commanded at 5 and 8
seconds to increase intake manifold pressure further while maintaining the desired
EGR% reference. The penalty on rate of change of LP-EGR mass flow rate was also
relaxed compared to the LP- EGR only case. It is noteworthy that the commanded LP-
EGR mass flow rate shows lesser oscillations despite the relaxed penalty indicating that
NMPC effectively uses HP-EGR to ‘supplement’ the LP-EGR during boosted operation.
146
Figure A-24 Pressure states and intake manifold EGR dilution trajectories for Dual-loop
EGR control.
147
Figure A-25 Input trajectories for Dual-loop EGR control.
148
Additional comparison between different cases
The intake manifold EGR% for different cases at 2000 RPM is shown in Figure
A-25. Even though the HP-EGR case settles to the reference at 5 seconds, there is an
error associated with the first order filling dynamics of the intake manifold visible
between 4.9 and 5 seconds. This error is diminished for the LP-EGR and DL-EGR case
as the primary EGR actuator in these cases is the LP-EGR mass flow rate which is farther
Figure A-26 Transient intake manifold EGR dilution for multiple EGR architectures
The performance of the NMPC for dual and single loop EGR architectures has been
summarized in TABLE A-2. At 1000 and 2000 RPM the DL-EGR shows reduced RMSE
149
and Max errors compared to the single loop architectures. However, at 3000 RPM the
DL-EGR showed poorer performance than the single loop systems. This can be attributed
and LP EGR mass flow rate to track the intake manifold pressure and intake manifold
balancing between the EGR loops whilst minimizing turbocharger lag. The NMPC
150
guarantees constraints and hence physically feasible values for the control inputs. It also
demonstrates the ability to consider the transport delay dynamics in the LP-EGR loop due
to the utilization of a multi-segment transport delay model. For the lower engine speeds,
and turbine bypass valve, engine cylinder models in addition to further refinement of the
151
REFERENCES
[1] Lecointe, B. and Monnier, G., "Downsizing a Gasoline Engine Using Turbocharging
https://doi.org/10.4271/2003-01-0542.
[2] Lake, T., Stokes, J., Murphy, R., Osborne, R. et al., "Turbocharging Concepts for
https://doi.org/10.4271/2004-01-0036.
[3] Zaccardi, J., Pagot, A., Vangraefschepe, F., Dognin, C. et al., "Optimal Design for a
https://doi.org/10.4271/2009-01-1794.
economy
[5] Shekaina, J., “Reduction Of Turbo Lag In Crdi Passenger Car And Afr Tuning Using
University.
152
[6] Jung, J., Park, C., and Bae, C., "Effects of High-Response TiAl Turbine Wheel on
[7] Jung, J., Oh, H., and Bae, C., "Characteristics of Turbocharger with TiAl Turbine
Wheel in a Downsizing GDI Engine," SAE Int. J. Fuels Lubr. 6(3):579-588, 2013,
https://doi.org/10.4271/2013-01-2499.
[8] Steidten, T., Rauscher, M., Muenz, S.: Measures to improve the transient behavior of
[10] Park, S., Matsumoto, T., and Oda, N., "Numerical Analysis of Turbocharger
https://doi.org/10.4271/2010-01-1226.
[11] Siokos, K., Koli, R., “Assessment of simulation-based calibration for fuel
[12] Siokos, K., Koli, R., Prucka, R., Schwanke, J. et al., "Assessment of Cooled Low
[13] Ashok, B., Denis, S., Kumar, C., “A review on control system architecture of a SI
[14] Cholvin, R., “Turbocharger Controls”, SAE Summer Meeting Atlantic City, June,
1962
153
[15] Ementhal, et al, “Turbocharging Small Displacement Spark Ignition Engines for
[16] Karnik, A., Buckland, J., and Freudenberg, J., “Electronic Throttle and Wastegate
[17] Lauber, K., Floquet, T., Colin, Y., “Nonlinear modelling and control approach for
Industrial Electronics
[18] Moulin, P., Chauvin, J., Youssef, B., “Modelling and control of the air system of
[19] Arnold, S., Groskreutz, M., Shahed, S., and Slupski, K., "Advanced Variable
Geometry Turbocharger for Diesel Engine Applications," SAE Technical Paper 2002-
[20] Hawley, J., Wallace, F., Cox, A., “Variable geometry turbocharging for lower
[21] Filipi, Z., Wang, Y., and Assanis, D., "Effect of Variable Geometry Turbine
(VGT) on Diesel Engine and Vehicle System Transient Response," SAE Technical
154
[22] Brace, C., Cox, A., Hawley, J., Vaughan, N. et al., "Transient Investigation of
Two Variable Geometry Turbochargers for Passenger Vehicle Diesel Engines," SAE
[23] Petitjean, D., Bernardini, L., Middlemass, C., and Shahed, S., "Advanced
[24] Shimizu, K., Sato, W., Enomoto, H., and Yashiro, M., "Torque Control of a Small
[25] Andersen, J., Karlsson, E., and Gawell, A., "Variable Turbine Geometry on SI
01-0020.
[26] Tang, H., Pennycott, A., Akehurst, S., Brace, C., “A review of the application of
[27] Martinez, R., Pesiridis, A., Yang, Y., “Overview of boosting options for future
https://doi.org/10.1007/s11431-010-4272-1
[28] Katrasnik, T., Rodman, S., Trenc, F., Hribernik, A., Medica, V., “Improvement
155
[29] Takaki, D., Tsuchida, H., Kobara, T., Akagi, M. et al., "Study of an EGR System
[30] Potteau, S., Lutz, P., Leroux, S., Moroz, S. et al., "Cooled EGR for a Turbo SI
Engine to Reduce Knocking and Fuel Consumption," SAE Technical Paper 2007-01-
[32] Wei, H., Zhu, T., Shu, G., “Gasoline engine exhaust gas recirculation – A
[33] Alger, T., Gingrich, J., Roberts, C., “Cooled exhaust-gas recirculation for fuel
[34] Cha, J., Kwon, J., Cho, Y. et al., “The effect of exhaust gas recirculation (EGR)
https://doi.org/10.1007/BF03185686
[35] Grandin, B., Ångström, H., Stålhammar, P., and Olofsson, E., "Knock
156
[36] Kaiser, M., Krueger, U., Harris, R., and Cruff, L., "“Doing More with Less” - The
Fuel Economy Benefits of Cooled EGR on a Direct Injected Spark Ignited Boosted
0589.
[37] Kiwan, R., Stefanopoulou, A., Martz, J., Surnilla, G. et al., "Effects of Differential
https://doi.org/10.4271/2017-01-0531.
[38] Kiwan, R., Stefanopoulou, A., Martz, J., “Effects of differential pressure
[39] Liu, F. and Pfeiffer, J., "Estimation Algorithms for Low Pressure Cooled EGR in
https://doi.org/10.4271/2015-01-1620.
[40] Koli, R., Siokos, K., Prucka, R., Jade, S. et al., "A Control Algorithm for Low
Pressure - EGR Systems Using a Smith Predictor with Intake Oxygen Sensor
01-0612.
University, 2017.
157
[42] Guzzela, L. and Onder, C., “Introduction to Modeling and Control of Internal
[43] Chung, J., Lee, S., “Measurement of Inertia of Turbocharger Rotor in a Passenger
[44] Leufven, O., and Eriksson, L., “Surge and Choke Capable Compressor”, Model
[45] Leufven, O., and Eriksson, L., “A surge and choke capable compressor flow
[46] Leufven, O., and Eriksson, L., “Measurement, analysis and modeling of
centrifugal compressor flow for low pressure ratios”, International Journal of Engine
Research, 2016
[47] Nelson, S., Filipi, Z., Assanis, D., “The Use of Neural Nets for Matching Fixed or
https://doi.org/10.4271/1999-01-0908.
[49] El Hadef, J., Colin, G., Chamaillard, Y., and Talon, V., "Physical-Based
158
[50] El Hadef, J., Janas, P., Colin, G., Talon, V. et al., "Geometry-Based Compressor
https://doi.org/10.4271/2013-01-0933.
[51] Jensen, J., Kristensen, A., Sorenson, S., Houbak, N. et al., "Mean Value Modeling
https://doi.org/10.4271/910070.
[52] Sorenson, S., Hendricks, E., Magnusson, S., and Bertelsen, A., "Compact and
Accurate Turbocharger Modelling for Engine Control," SAE Technical Paper 2005-
[53] Eriksson, L., “Modeling and Control of Turbocharged SI and DI Engines”, Oil &
Gas Science and Technology – Rev. IFP, Vol. 62 (2007), No. 4, pp. 523-538
[54] Zhu, Q., “A Study Model Predictive Control for Spark Ignition Engine
[55] Lee, J., “Model Predictive Control: Review of the Three Decades of
[56] Rawlings, J., Muske, R., “The Stability of Constrained Receding Horizon
Control”, IEEE Transactions On Automatic Control, Vol. 38, No. 10, October 1993
[57] Colin, G., Chamaillard, Y., Charlet, A., Bloch, G. et al., "Linearized Neural
159
[58] Colin, G., Chamaillard, Y., Bloch, G., Corde, G., “Neural Control of Fast
[59] Le Berr, F., Miche, M., Le Solliec, G., Lafossas, F. et al., "Modelling of a
01-3264.
[60] Ericsson, G., Angstrom, H., and Westin, F., "Optimizing the Transient of an SI-
Engine Equipped with Variable Cam Timing and Variable Turbine," SAE Int. J.
[61] “Exhaust Gas Recirculation Control for Large Diesel Engines - Achievable
[62] Hegarty, K., Dickinson, P., Cieslar, D., and Collings, N., “Fast O2 Measurement
using Modified UEGO Sensors in the Intake and Exhaust of a Diesel Engine,” SAE
[63] Welling, O. and Collings, N., “UEGO Based Measurement of EGR Rate and
doi:10.4271/2011-01-1289.
publications
160
[66] Nishio, Y., Hasegawa, M., Tsutsumi, K., Goto, J. et al., "Model Based Control for
Dual EGR System with Intake Throttle in New Generation 1.6L Diesel Engine," SAE
[67] Fu, J. and Kurihara, N., "Intake Air Control of SI Engine Using Dead-Time
3267.
[68] Siokos, K., Koli, R., Prucka, R., Schwanke, J., Jade, S.,“Physics-Based Exhaust
Manifold Pressure and Temperature Estimation for Low Pressure EGR Control in
[69] Zheng, M., Reader, G., Hawley, J., “Diesel engine exhaust gas recirculation––a
review on advanced and novel concepts”, Vol 45, Issue 6, Energy conversion and
Management, 2004
“Multivariable Control of Dual Loop EGR Diesel Engine with a Variable Geometry
[71] Wang, S., “Model Based Combustion Phasing Control for High Degree of
[72] Baratta, M., Spessa, E., and Mairone, P., “Numerical investigation of turbolag
reduction in hd cng engines by means of exhaust valve variable actuation and spark
timing control”, International Journal of Automotive Technology, Vol. 11, No. 3, pp.
289−306 (2010)
161
[73] Ferreau, H., Ortner, P., Langthaaler, P., del Re, L., Diehl, M., “Predictive control
of a real-world diesel engine using an extended online active set strategy”, Annual
[74] Ortner, P., Langthaaler, P., Ortiz, J., del Re, L., “MPC for a diesel engine air path
using an explicit approach for constraint systems”, in Proc. Of 2006 IEEE International
2006.
[75] Garcı´a-Nieto, S., Martı´nez, M., Blasco, X., Sanchis, J., “Nonlinear predictive
control based on local model networks for air management in diesel engines” Control
[76] Emekli, Aksun, “Explicit MIMO Model Predictive Boost Pressure Control of a
[77] Wiese, A., Stefanopoulou, A., Karnik, A., Buckland, J., “Model Predictive
Control for Low Pressure Exhaust Gas Recirculation with scavenging” 2017
[78] Hu, Y., Chen, H., Ping, W., Chen, H., Ren, L., “Nonlinear model predictive
passenger vehicle”, Mechanical Systems and Signal Processing, Vol. 109, 2018.
[79] Zhu, Q., Onori, S., and Prucka, R., “Nonlinear economic model predictive control
162
[80] Zhu, Q., Onori, S., and Prucka, R., “Pattern recognition technique based active set
QP strategy applied to MPC for a driving cycle test”, in Proc. of 2015 American Control
[81] Zhu, Q., Onori, S., and Prucka, R., “Nonlinear economic model predictive control
[82] Bemporad, A., Bernardini, D., Long, R., and Verdejo, J., "Model Predictive
[83] Zhu, Q., Koli, R., Feng, L., Onori, S., Prucka, R., “Nonlinear Model Predictive Air
Path Control for Turbocharged SI Engines with Low Pressure EGR and a Continuous
[84] Nicolau, G., Scattolini, R., Siviero, C., “Modelling the volumetric efficiency of ic
[85] Wu, B., Filipi, Z., Assanis, D., Kramer, D. et al., "Using Artificial Neural Networks
for Representing the Air Flow Rate through a 2.4 Liter VVT Engine," SAE Technical
network.html
163
[87] Lee, T., Filipi, Z., “Nonlinear model predictive control of a dual-independent
variable valve timing engine with electronic throttle control”, Proceedings of the
[88] Vahidi, A., Stefanopoulou, A., “Current Management in a Hybrid Fuel Cell
[89] Thomasson, A., Llamas, X., and Eriksson, L., "Turbo Speed Estimation Using
https://doi.org/10.4271/2017-01-0591.
[90] Antoniewicz, K., Jasinski, M., Kazmierkowski, M., Malinowski, M., “Model
Shunt Active Power Filter”, IEEE transactions on industrial electronics, vol. 63, no. 8,
august 2016
[91] Hoffman, K., Heßeler, F., Abel, D., “Rapid Control Prototyping with Dymola and
Matlab for a Model Predictive Control for the air path of a boosted diesel engine”, E-
COSM - Rencontres Scientifiques de l'IFP - 2-4 Octobre 2006, Proceedings, pp. 33-40
https://www.gmperformancemotor.com/parts/19328837.html
[93] X. Zhou, Y. Li, Y. Hu, and H. Chen, “Torque tracking control of turbocharged
gasoline engine using nonlinear MPC”, In 2015 European Control Conference, pp.
2958-2963, (2015).
164
[94] H. Borhan, G. Kothandaraman, and B. Pattel, “Air handling control of a diesel
engine with a complex dual-loop EGR and VGT air system using MPC”, In 2015
[95] B. Houska, H.J. Ferreau, and M. Diehl. ACADO Toolkit – An Open Source
165