10
Nonlinear internal model controller design for wastegate control of a turbocharged gasoline engine $ Zeng Qiu a,n , Jing Sun b , Mrdjan Jankovic c , Mario Santillo c a Department of Electrical Engineering and Computer Science, University of Michigan, Ann Arbor, MI, USA b Department of Navel Architecture & Marine Engineering, University of Michigan, Ann Arbor, MI, USA c Ford Motor Company, Dearborn, MI, USA article info Article history: Received 17 February 2015 Received in revised form 1 September 2015 Accepted 19 October 2015 Keywords: Internal model control Wastegate control Turbocharged gasoline engine Quasi-linear parameter varying abstract Internal Model Control (IMC) has a great appeal for automotive powertrain control in reducing the control design and calibration effort. Motivated by its success in several automotive applications, this work investigates the design of nonlinear IMC for wastegate control of a turbocharged gasoline engine. The IMC design for linear time-invariant (LTI) systems is extended to nonlinear systems. To leverage the available tools for LTI IMC design, the quasi-linear parameter-varying (quasi-LPV) models are explored. IMC design through transfer function inverse of the quasi-LPV model is ruled out due to parameter variability. A new approach for nonlinear inversion, referred to as the structured quasi-LPV model in- verse, is developed and validated. A fourth-order nonlinear model which sufciently describes the dy- namic behavior of the turbocharged engine is used as the design model in the IMC structure. The con- troller based on structured quasi-LPV model inverse is designed to achieve boost-pressure tracking. Fi- nally, simulations on a validated high-delity model are carried out to show the feasibility of the pro- posed IMC. Its closed-loop performances are compared with a well-tuned PI controller with extensive feedforward and anti-windup built in. Robustness of the nonlinear IMC design is analyzed using simu- lations. & 2015 Elsevier Ltd. All rights reserved. 1. Introduction Internal Model Control (IMC), whose diagram is shown in Fig. 1, is a well-established control design methodology with an intuitive control structure (Morari & Zariou, 1989). It incorporates a sys- tem model as an explicit element in the controller so that the control actions are determined based on the difference between the model output and the plant output. It has several desired features and closed-loop properties as established in Morari and Zariou (1989) and Garcia and Morari (1982), such as dual stability criterion, zero offset, and perfect control. The design, analysis, and implementation of IMC for linear systems have been well devel- oped. Rivera, Morari, and Skogestad (1986) showed that process industrial IMCs for many SISO models can lead to PID controllers, occasionally augmented with a rst-order lag. They also demon- strated the superiority of using IMC for PID tuning in terms of closed-loop performance and robustness. The efcacy of IMC for nonlinear systems, however, has been investigated with limited comprehensive results. Economou, Morari, and Palsson (1986) presented an important result of nonlinear IMC, proving that the dual stability criterion, zero offset, and perfect control properties of LTI IMC would carry over to nonlinear cases. The IMC was implemented by nding a nonlinear dynamic inverse, which remained to be the key challenge in ex- tending the IMC design to nonlinear systems. While the invert- ibility condition, inverse structure, and derivation for nonlinear dynamic system inverse were studied (Hirschorn, 1979), the de- rivation of the nonlinear inverse involved higher-order derivatives and caused problems when noises and disturbances were present in the system. In Economou et al. (1986), the nonlinear inverse was derived by exploiting the Hirschorn nonlinear inverse structure and solving it numerically using the contraction principle method or Newton's method. Stability of the IMC structure was discussed under the ideal circumstance that the model was the same as the plant. Henson and Seborg (1991) also exploited the result of Hirschorn nonlinear inverse for nonlinear IMC design. Several as- sumptions were made to calculate the higher-order derivatives. Feedforward/feedback linearization approach was adopted by Calvet and Arkun (1988) to derive the model for the nonlinear Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/conengprac Control Engineering Practice http://dx.doi.org/10.1016/j.conengprac.2015.10.012 0967-0661/& 2015 Elsevier Ltd. All rights reserved. This work is supported by Ford Motor Company. n Corresponding author. E-mail addresses: [email protected] (Z. Qiu), [email protected] (J. Sun), [email protected] (M. Jankovic), [email protected] (M. Santillo). Control Engineering Practice 46 (2016) 105114

Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

Control Engineering Practice 46 (2016) 105–114

Contents lists available at ScienceDirect

Control Engineering Practice

http://d0967-06

☆Thisn CorrE-m

mjanko

journal homepage: www.elsevier.com/locate/conengprac

Nonlinear internal model controller design for wastegate control of aturbocharged gasoline engine$

Zeng Qiu a,n, Jing Sun b, Mrdjan Jankovic c, Mario Santillo c

a Department of Electrical Engineering and Computer Science, University of Michigan, Ann Arbor, MI, USAb Department of Navel Architecture & Marine Engineering, University of Michigan, Ann Arbor, MI, USAc Ford Motor Company, Dearborn, MI, USA

a r t i c l e i n f o

Article history:Received 17 February 2015Received in revised form1 September 2015Accepted 19 October 2015

Keywords:Internal model controlWastegate controlTurbocharged gasoline engineQuasi-linear parameter varying

x.doi.org/10.1016/j.conengprac.2015.10.01261/& 2015 Elsevier Ltd. All rights reserved.

work is supported by Ford Motor Company.esponding author.ail addresses: [email protected] (Z. Qiu), [email protected] (M. Jankovic), [email protected]

a b s t r a c t

Internal Model Control (IMC) has a great appeal for automotive powertrain control in reducing thecontrol design and calibration effort. Motivated by its success in several automotive applications, thiswork investigates the design of nonlinear IMC for wastegate control of a turbocharged gasoline engine.The IMC design for linear time-invariant (LTI) systems is extended to nonlinear systems. To leverage theavailable tools for LTI IMC design, the quasi-linear parameter-varying (quasi-LPV) models are explored.IMC design through transfer function inverse of the quasi-LPV model is ruled out due to parametervariability. A new approach for nonlinear inversion, referred to as the structured quasi-LPV model in-verse, is developed and validated. A fourth-order nonlinear model which sufficiently describes the dy-namic behavior of the turbocharged engine is used as the design model in the IMC structure. The con-troller based on structured quasi-LPV model inverse is designed to achieve boost-pressure tracking. Fi-nally, simulations on a validated high-fidelity model are carried out to show the feasibility of the pro-posed IMC. Its closed-loop performances are compared with a well-tuned PI controller with extensivefeedforward and anti-windup built in. Robustness of the nonlinear IMC design is analyzed using simu-lations.

& 2015 Elsevier Ltd. All rights reserved.

1. Introduction

Internal Model Control (IMC), whose diagram is shown in Fig. 1,is a well-established control design methodology with an intuitivecontrol structure (Morari & Zafiriou, 1989). It incorporates a sys-tem model as an explicit element in the controller so that thecontrol actions are determined based on the difference betweenthe model output and the plant output. It has several desiredfeatures and closed-loop properties as established in Morari andZafiriou (1989) and Garcia and Morari (1982), such as dual stabilitycriterion, zero offset, and perfect control. The design, analysis, andimplementation of IMC for linear systems have been well devel-oped. Rivera, Morari, and Skogestad (1986) showed that processindustrial IMCs for many SISO models can lead to PID controllers,occasionally augmented with a first-order lag. They also demon-strated the superiority of using IMC for PID tuning in terms ofclosed-loop performance and robustness.

[email protected] (J. Sun),(M. Santillo).

The efficacy of IMC for nonlinear systems, however, has beeninvestigated with limited comprehensive results. Economou,Morari, and Palsson (1986) presented an important result ofnonlinear IMC, proving that the dual stability criterion, zero offset,and perfect control properties of LTI IMC would carry over tononlinear cases. The IMC was implemented by finding a nonlineardynamic inverse, which remained to be the key challenge in ex-tending the IMC design to nonlinear systems. While the invert-ibility condition, inverse structure, and derivation for nonlineardynamic system inverse were studied (Hirschorn, 1979), the de-rivation of the nonlinear inverse involved higher-order derivativesand caused problems when noises and disturbances were presentin the system. In Economou et al. (1986), the nonlinear inverse wasderived by exploiting the Hirschorn nonlinear inverse structureand solving it numerically using the contraction principle methodor Newton's method. Stability of the IMC structure was discussedunder the ideal circumstance that the model was the same as theplant. Henson and Seborg (1991) also exploited the result ofHirschorn nonlinear inverse for nonlinear IMC design. Several as-sumptions were made to calculate the higher-order derivatives.Feedforward/feedback linearization approach was adopted byCalvet and Arkun (1988) to derive the model for the nonlinear

Page 2: Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

Fig. 1. Internal model control structure.

Z. Qiu et al. / Control Engineering Practice 46 (2016) 105–114106

plant in IMC. Their approach accounted for the disturbances andinput constraints.

Nonlinear IMC was also investigated in the adaptive controlframework. Hunt and Sbarbaro (1991) used artificial neural net-works for adaptive control of nonlinear IMC. Feasibility of identi-fying the nonlinear model and its inverse by a neural network wasexplored and demonstrated. Boukezzoula, Galichet, and Foulloy(2000) and Xie and Rad (2000) used fuzzy logic to estimate themodel dynamics. The inverse was derived from this fuzzy model.The black-box identifications of neural network and fuzzy logicmade it difficult to incorporate physical knowledge about the plantin the IMC design. In adaptive IMC scheme, using linear models torepresent the dynamics of the nonlinear plant though adaptationhas also been exploited (Brown, Lightbody, & Irwin, 1997; Datta,1998; Shafiq, 2005).

Another possible avenue to exploit the linear IMC design toolsfor nonlinear systems would be through the linear parametervarying (LPV) model. Mohammapour, Sun, Karnik, and Jankovic(2013) applied IMC on a quasi-LPV model with two approaches. Inthe first approach, the IMC controller parameters were scheduledbased on the LPV model parameters which were assumed to beknown in real time and not vary rapidly. In the second approach,the design problem was formulated in the H∞ framework as a setof linear matrix inequalities (LMI). Solving the associated LMIproblem, however, was computationally intensive. Toivonen,Sandström, and Nyström (2003) derived the LPV model based onvelocity-based linearization, then developed the IMC controllerbased on linear IMC theory. It was much less computationallydemanding, but it was only applicable when there were a smallnumber of scheduling parameters.

This paper explores nonlinear IMC for turbocharged gasolineengines driven by the need for developing robust and easy-to-calibrate powertrain control solutions and motivated by severalsuccessful industrial applications. IMC was first applied to turbo-charged diesel engines for automotive applications. Alfieri, Am-stutz, and Guzzella (2009) applied IMC based on the classicalSmith predictor structure to air–fuel ratio control in turbochargeddiesel engines with exhaust gas recirculation. Schwarzmann,Nitsche, and Lunze (2006) treated boost-pressure control of aturbocharged diesel engine with a variable nozzle turbine withIMC. Their IMC utilized a flatness-based approach to design theinverse Q, in which flatness means that the system inputs can beexplicitly expressed in terms of internal system dynamics. In afollow up work, the same group also dealt with a two-stagedturbocharged diesel engine using IMC (Schwarzmann, Nitsche,Lunze, & Schanz, 2006). The inverse Q was designed based ongeometric nonlinear control design method. As turbocharged ga-soline engines are becoming more popular, advanced control de-signs including IMC have been applied to turbocharged gasolineengines for improved performance. Thomasson, Eriksson, Leufven,and Andersson (2009) utilized IMC for PID tuning of wastegatecontrol in turbocharged gasoline engines. Karnik and Jankovic

(2012) later applied IMC directly to wastegate control for a tur-bocharged gasoline engine, motivated by the successful applica-tions on turbocharged diesel engines. They used a first-ordermodel which was simplified from a fourth-order nonlinear modelusing singular perturbation. While the simplicity of the first-ordermodel-based design was an advantage for implementation, itsperformance was limited by the linear approximation, as it is de-fined for a particular operating point.

This work investigates the feasibility, performance, advantages,and limitations of a nonlinear IMC for automotive powertrain-control design, using the fixed geometry turbocharged gasolineengine as a case study. While the nonlinear dynamics of the sys-tem can be sufficiently described by a fourth-order model, in-verting the nonlinear model for the IMC design represents themajor challenge. To facilitate the IMC design, a quasi-LPV model(Rugh & Shamma, 2000) for the nonlinear model is developed.More importantly, the special quasi-LPV model structure is ex-plored, and a structured quasi-LPV model is proposed, which leadsto a feasible nonlinear inverse, referred to as the structured quasi-LPV inverse. The IMC based on the structured quasi-LPV inverse isdeveloped, and its performance is analyzed. Simulation results,using a validated “virtual” plant model, are presented to demon-strate the effectiveness of the proposed design. This work is ap-plicable to IMC with SISO nonlinear models of higher-order and isnot limited by the number of scheduling parameters. The pro-posed IMC was originally presented as a conference paper (Qiu,Sun, Jankovic, & Santillo, 2014), whereas this paper represents anexpanded version. More specifically, the design procedure is dis-cussed in more detail from the stability point of view and ro-bustness analysis is included.

The paper is organized as follows: Section 2 states the problem,presents two main tools used: IMC and LPV. Section 3 presents thenonlinear model for the turbocharged gasoline engine and exploitsquasi-LPV approach to derive its inverse. Section 4 analyzes theIMC implementation results. Section 5 summarizes the paper.

2. Control problem and preliminaries

Gasoline engines have been aggressively downsized in an effortto reduce fuel consumption and CO2 emissions. However, thetorque provided by the engine is proportional to the air deliveredto the cylinders. To meet the consumer demands for performanceon the downsized engines, i.e., to maintain the engine outputtorque, turbochargers are widely adopted. They compress the in-take air to increase the density of the engine airflow, thereby in-creasing the torque. The schematic of a turbocharged gasolineengine is shown in Fig. 2. The wastegate is the main actuator tocontrol boost pressure. It affects the engine operation by changingthe rotational speed of the turbine/compressor. The air is com-pressed by the compressor, and passes through an intercooler anda throttle before entering the engine intake port. The engine ex-haust port is connected to the turbine, which is mechanicallyconnected to the compressor. An electric wastegate actuator con-trols the opening of the turbine bypass path in this application(Karnik & Jankovic, 2012), affecting the compressor speed andtherefore the boost pressure.

The turbocharged gasoline engine is expected to produce thedesired engine torque, with higher fuel efficiency, power density,and lower emission (Guzzella & Onder, 2010). To achieve such goal,the desired engine torque is calculated from the driver pedal po-sition. The desired engine torque is then mapped into desired in-take manifold pressure and boost pressure considering the fueleconomy and emission. These two pressures are then trackedthrough throttle and wastegate. This two input two output controlproblem can often be tackled with a decentralized controller:

Page 3: Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

Fig. 2. System schematic of a turbocharged gasoline engine (Buckland, 2009).

Plant ++_Inverse

Model

IMC controller

+

Fig. 3. Equivalent internal model control structure.

Z. Qiu et al. / Control Engineering Practice 46 (2016) 105–114 107

using the throttle to track the intake manifold pressure and usingthe wastegate to track the boost pressure (Karnik, Buckland, &Freudenberg, 2005). In this paper, we will focus on using thewastegate to track the desired boost pressure, and the throttle isconsidered as an exogenous input. Boost pressure set-pointtracking is a critical enabling technology for achieving improvedfuel efficiency, power density, and emission reduction (Guzzella &Onder, 2010). However, operating conditions vary widely in au-tomotive applications. This variation could result in inadequateboost at low speeds and loads, and over-boost situation at highspeeds and loads. The wide range in operating conditions addscomplexity to control design and calibration. IMC with a nonlinearmodel is adopted for its advantages in reducing the design andcalibration effort.

A nonlinear IMC is developed in this paper for the controlproblem of boost-pressure tracking with the electric wastegate asthe actuator. Measurements for boost pressure Pb, temperatures Tb,Ti, Te, throttle opening uth, and engine speed Nen are assumed to beavailable for feedback or feedforward control. The invertibility ofthe nonlinear model is assumed. To proceed with the designprocedure of nonlinear IMC with structured quasi-LPV inverse, thepreliminaries of IMC and quasi-LPV model are presented as below.

2.1. Internal model control

The schematic of a system with IMC is shown in Fig. 1, where P,M, and Q denote the plant, model, and inverse respectively. TheIMC controller contains two elements: the inverse Q and themodel M. Q takes the reference command and the difference be-tween the outputs of the plant and the model as its inputs. If themodel is exact, i.e., M¼P, the IMC becomes a feedforward con-troller. Q is generally designed such that the norm of the differencee between the reference command r and the plant output y isminimized. Thus Q is chosen by solving

e QM rmin min 1 ,1Q Q

2 2∥ ∥ = ∥( − ) ∥( )

subject to the constraint that Q is stable and casual. An absoluteminimum could be reached at Q M 1= − if M is invertible. Howeverthis is not feasible for a non-minimum phase (NMP) model, whichis not stably invertible. An approximate inverse Q can be solvedalgebraically depending on the input r. Therefore, once the linearmodel is derived, the inverse of the model can be derived bysolving the minimization problem (1). For a nonlinear model,however these tools are inapplicable, and effective approaches forinverting nonlinear models are very limited. Therefore, deriving

the model inverse is the main challenge and focus in this work. Forboth linear and nonlinear systems, once the model and its validinverse are derived, IMC control design follows immediately. IMChave three main properties as established for LTI models (Garcia &Morari, 1982) and later extended to nonlinear cases (Economouet al., 1986):

(1) Dual stability property: When the model M is exact, stabilityof both inverse Q and plant P is sufficient for overall systemstability.

(2) Perfect control property: By changing the position of thesignal addition block in the IMC in Fig. 1, one can get an equivalentform of IMC as in Fig. 3, which has the same structure as theclassical control. With this structure, the IMC controller can bepresented by Q MQ1 1( − )− , therefore, the sensitivity function of theIMC control system is MQ MQ PQ1 1 1( − )( − + )− . When MQ¼1, i.e.,Q M 1= − , the sensitivity function is 0, which means “perfect con-trol” is achieved.

(3) Zero offset property: A controller which satisfiesQ M0 0 1( ) = ( )− adds integral action to the controller, and yields zerooffset for a constant reference command r.

2.2. Quasi linear parameter varying model

Quasi-LPV models are LPV models for nonlinear systems wherenonlinearities are hidden through state-dependent parameters, sothat a nonlinear model can be represented by an LPV model andtreated by LPV design techniques (Rugh & Shamma, 2000).

In general, a nonlinear model in the form of

x f x u, 2 = ( ) ( )

can be expressed as an LPV model in the form of

x A p x B p u 3 = ( ) + ( ) ( )

if the model (2) is affine in u and the time varying parametervector p in (3) is allowed to be state-dependent to disguise thenonlinearities (Rugh & Shamma, 2000). For example, the nonlinearmodel

x x x x x x u, sin1 12

1 2 2 1 = + = +

can be expressed in quasi-LPV form as

⎣⎢⎢

⎦⎥⎥

⎡⎣⎢

⎤⎦⎥x A p x Bu

x xx u0

01

,xx

1 1

sin 1

1

= ( ) + = +( )

with⎡⎣⎢⎢

⎤⎦⎥⎥p x , x

x

T

1sin 1

1= ( ) , or

⎣⎢⎢

⎦⎥⎥

⎡⎣⎢

⎤⎦⎥x A p x Bu

x xx u

0

001

,xx

1 2

sin 1

1

= ( ) + =+

+( )

with⎡⎣⎢⎢

⎤⎦⎥⎥p x x , x

x

T

1 2sin 1

1= + ( ) .

Page 4: Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

Z. Qiu et al. / Control Engineering Practice 46 (2016) 105–114108

In the next section, a nonlinear model for the turbochargedgasoline engine is presented. The quasi-LPV approach is exploitedfor representing it in a linear structure to aid deriving an inversefor the nonlinear model in IMC implementation.

3. Quasi-LPV model and its inversion

3.1. A nonlinear model for IMC design

Control-oriented models serve the IMC design and im-plementation in two different ways: first, the IMC incorporates asystem model directly in its implementation as shown in Fig. 1;second, the standard IMC design procedure takes an inverse of theprocess model and augments it with a proper filter to avoid non-causal implementation to form the inverse Q.

The nonlinear model for the boost-pressure dynamics of aturbocharged engine presented in this paper is based on the workof Buckland (2009). The nonlinear model has the following statesand one input:

x P P P N u u, , , , ,b i e tT

w= [ ] =

where Pb is the boost pressure, Pi is the intake pressure, Pe is theexhaust pressure, Nt is the turbocharger speed, and the input uw isthe wastegate, which is the fraction of the opening and takes va-lues in the range of 0, 1[ ]. The dynamics of the pressures Pb, Pi, andPe are derived using mass conservation along with isothermalmanifold assumptions, while the dynamics of the turbochargerspeed Nt are derived by a power balance between the turbine andthe compressor as described in Karnik and Jankovic (2012) andBuckland (2009). The equations are summarized as follows:

⎛⎝⎜

⎞⎠⎟

dPdt

RTV

W W

dPdt

RTV

W W

dPdt

RTV

WA F

A FW W

dNdt I N

H H

,

,

1 //

,

1,

4

b b

bc th

i i

ith en

e e

een t w

t

t tt c

= ( − )

= ( − )

= + − −

= ( − )( )

where R is the ideal gas constant, A F/ is the air to fuel ratio, T, V,W,I and H are temperature, volume, mass flow rate, inertia, andpower respectively. The subscript indicates the physical location ofthe variable as in Fig. 2, and b, c, th, i, en, e, t, and w are boost,

Table 1Nomenclature for modeling of turbocharged gasoline engine.

Variables Subscripts

A/F Air to fuel ratio a Ambientcp,(·) Specific heat at constant pressure b Boost

f (·)· Compressor/turbine map c CompressorI Moment of inertia e ExhaustN Rotational speed en EngineP Pressure i IntakeH Power t TurbineR Ideal gas constant th ThrottleT Temperature w Wastegateu Fraction/degree of opening x Turbine exitV VolumeW Mass flow rateϕ(·) Function of pressure ratio

across the componentη Isentropic efficiencyψ Mass flow parameterγ Ratio of specific heats

compressor, throttle, intake, engine, exhaust, turbine, and waste-gate respectively. Modeling of W (mass flow rate) and H (power)are described in detail in Karnik and Jankovic (2012) and Buckland(2009), and the resulting functional expressions are summarizedas follows:

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜⎜

⎞⎠⎟⎟

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

W fPP

N

Wu

RTP

PP

W PVRT

N

Wu

RTP

PP

W fPP

N

T

P

T

H c T WPP

H c T WPP

, ,

sat 0, , 1,

2,

sat 0, , 1,

, ,

, 1 ,

1, 1,

5

c cb

at

thth

bb

i

b

en i enen

i

en

ww

ee

x

e

t te

x

t

e

e

e

t p e e t t t tx

e

c p a a cc

c cb

a

,

1 /

,

1 /

e e

a a

γ ϕ

η

γ ϕ

η ψ ψ

ηψ ψ

=

=( )

=

=( )

=

= = −

= = −( )

γ γ

γ γ

( − )

( − )

where Pa is the ambient pressure, Px is the turbine exit pressure,ϕ(·) is a function of pressure ratio across the component, ψ is amass flow parameter, γ is the specific heat ratio for air, cp,(·) is thespecific heat at constant pressure, η is the isentropic efficiency, uthis the throttle opening, Nen is the engine speed, and usat 0, , 1( )limits u to be in the range 0, 1[ ]. The temperatures Tb, Ti, and Te areassumed to be measured. Typically the temperature sensors have aslow response time and delay, and the measurements are leadfiltered to improve the response time. Therefore, the measurementinaccuracy is not considered in this work. A F/ is the stoichiometricratio of gasoline. uth and Nen are considered as exogenous inputs inthis work and they are measurable. All the variables and subscriptsfor the nonlinear model of turbocharged gasoline engine aresummarized in Table 1.

The nonlinear model is evaluated by comparing its responseswith those of the virtual “plant”, which is a high fidelity Fordproprietary model that has been validated extensively. It includesthe intercooler, throttle, engine, wastegate, turbine, and com-pressor (Buckland, 2009). Responses to a step change in the was-tegate setting from 0.25 to 0.75 at t 5 s= for the nonlinear modeland the virtual “plant” are shown in Fig. 4, confirming that thecontrol-oriented nonlinear model and the “plant” have very si-milar dynamic responses.

3.2. Quasi-LPV turbocharged gasoline engine model

IMC control can be achieved by designing the inverse Q in Fig. 1as the inverse of the model M. For linear models, their inverses canbe achieved by inverting their transfer functions and appending aproper filter to assure causality. To extend this approach to non-linear models, the quasi-linear parameter varying model approachis explored to represent the nonlinear model in a linear structure.

Note that there are an infinite number of quasi-LPV models inthe form of (3) that can match (2), depending on the choice of thevarying parameter p. For the turbocharged gasoline engine system,the physical couplings of state variables are considered and thefollowing structure that leads to the most sparse A, B matrices ischosen:

Page 5: Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

0 5 10100

150

200

Time [sec]

Boo

st P

res

[kP

a]

0 5 10100

150

200

Time [sec]

Inta

ke P

res

[kP

a]

0 5 10100

150

200

Time [sec]

Exh

aust

Pre

s [k

Pa]

0 5 100.5

1

1.5x 104

Time [sec]

Turb

o S

peed

[rad

/sec

]

Nonlinear

Plant

Fig. 4. Comparison of responses of the nonlinear model and the “plant” for a step change in wastegate actuation.

Z. Qiu et al. / Control Engineering Practice 46 (2016) 105–114 109

⎢⎢⎢⎢⎢

⎥⎥⎥⎥⎥

⎢⎢⎢⎢

⎥⎥⎥⎥

A

a a

a a

a a

a a

Bb

x P P P N u u y x

0 0

0 0

0 0

0 0

,

00

0

,

, , , , , . 6b i e tT

w

11 14

21 22

32 33

43 44

3

1

= =

= [ ] = = ( )

The non-zero elements in (6) are defined as follows:

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜⎜

⎞⎠⎟⎟

⎛⎝⎜

⎞⎠⎟

aRTV

WP

RT

Vu

PP

aRTV

WN

RTV N

fPP

N

aRTV

WP

RTV

u A

RT

PP

aRTV

WP

RTV

V NRT

aA F

A FRTV

WP

A FA F

T V N

V T

aRTV

WP

R T

Vf

PP

N

T

aH

I N P I Nc T

WP

aH

I N I Nc T W

bRTV

Wu

RT

VP

PP

sat 0, , 1 ,

, ,

sat 0, , 1,

2,

1 //

1 // 2

,

, ,

1,

1 1,

sat 0, , 1.

7

b

b

th

b

b

bth

i

b

b

b

c

t

b

b tc

b

at

i

i

th

b

i

i

th max

b

i

b

i

i

en

i

i

i

en en

i

e

e

en

i

e en en en

e i

e

e

t

e

e

et

e

x

t

e

t

t t e t tp e e

t

et t

c

t t t tp a a c

cc

e

e

w

w

e

ee

x

e

11

14

21

22

32

33

43 ,

44 2 2 ,

3

γϕ

ϕ γ

η

η ψ

ηψ

γ ϕ

= − = − ( )

= =

= =( )

= − = −

= + = +

= − = −

= =

= − = −

= −( )

= −( )

3.3. Structured quasi-LPV inverse

Given that the parameters defined by (7) are varying fast dur-ing transients, treating the parameters as frozen and deriving thetransfer function of (6) will not be effective for deriving the in-verse. Indeed, by numerical simulations it is confirmed that thetransfer function inverse does not represent the nonlinear modelinverse. In this section, the special form of the quasi-LPV structureof (6) is explored to derive its inverse model in an effort to

minimize the approximation error.(1) Quasi-LPV model inverse structure: Exploiting the sparsity of

the A, B matrices of model (6), the quasi-LPV model is expressed asan integration of several first-order sub-models. With this veryspecial structure of the nonlinear model, the inverse can be pur-sued by deriving the inverse of multiple first order nonlinearmodels, which will involve limited approximation.

Given the sparse matrices A, B in the form of (6), the followingfirst-order sub-models , , ,1 2 3 4Σ Σ Σ Σ are defined as

x a x a x x x

x a x a x x x

x a x a x b u x x u

x a x a x x x

: ,

: ,

: , ,

: . 8

1 1 11 1 14 4 1 1 4

2 2 22 2 21 1 2 2 1

3 3 33 3 32 2 3 3 3 2

4 4 44 4 43 3 4 4 3

Σ Σ

Σ Σ

Σ Σ

Σ Σ

= + ⟹ = ( )

= + ⟹ = ( )

= + + ⟹ = ( )

= + ⟹ = ( ) ( )

Further expressing Σ3 to be

x u x u, ,3 2 31 2 32Σ Σ Σ( ) = ( ) + ( )

we can show that the input–output relation of the quasi-LPVmodel can be expressed as a composition of these sub-models:

y x x u x u, , 91 1 4 3 2 1 1 4 31 2 1 32Σ Σ Σ Σ Σ Σ Σ Σ Σ= = ( ( ( ( ) ))) = ( ( ( ( )) + ( ))) ( )

whose block diagram representation is shown in Fig. 5.Expressing the input u in terms of the output y based on (9),

u y y , 10321

41

11

31 2Σ Σ Σ Σ Σ= ( ( ( )) − ( ( ))) ( )− − −

which can be viewed as an inverse model of (9). Fig. 6 shows theblock diagram representation of the inverse of the quasi-LPVmodel through the integration of several inverse models of first-

order blocks as expressed in (10), in which 11Σ − , 4

1Σ − , and 321Σ − are

approximate inverses of Σ1, Σ4, and Σ32. The derivations of the

11Σ − , 4

1Σ − , and 321Σ − are explained in the following section. x x1 4– in

Fig. 6 are approximations of x1–x4 in Fig. 5.(2) First-order quasi-LPV inverse: We now derive the inverse of

the first-order quasi-LPV model and define its property in order toderive the representation for the inverse model (10).

For a first-order quasi-LPV model:

x ax bu: , 11iΣ = + ( )

define

Page 6: Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

Fig. 5. Interconnection of the first-order quasi-LPV sub-models for the fourth-orderturbocharged gasoline engine LPV model.

Fig. 6. Interconnection of first-order quasi-LPV sub-models for inverse of the LPVmodel shown in Fig. 5.

Fig. 7. Structure for validation of first-order inverse.

Z. Qiu et al. / Control Engineering Practice 46 (2016) 105–114110

⎧⎨⎩⎫⎬⎭u

su

11

,12τ

¯ =+ ( )

wheres

11

{ }τ +

denotes the first-order filter with a transfer function

s1

1τ +. Then, the following lemma gives the representation of u.

Lemma 1. Let a, b be time varying parameters with a b, ∈ ∞, b 0≠ ,and b 1∈ . Then, for any u, u given by (12) can be expressed in termsof the state x as

⎛⎝⎜⎜

⎛⎝⎜

⎞⎠⎟

⎞⎠⎟⎟

ub

x z

z zb

bb

ab

x

1,

1 1,

132

τ

τ τ

¯ = −

= − + −

+( )

where x is given by (11).

Proof. Note that

⎜ ⎟⎧⎨⎩

⎫⎬⎭⎧⎨⎩

⎫⎬⎭⎛⎝

⎞⎠u

su

sxb

axb

11

11

.14τ τ

¯ =+

=+

( )

Since

⎜ ⎟⎛⎝

⎞⎠

ddt

xb

xb

bxb

,152=

( )

we have, from (14), that

⎪ ⎪ ⎪ ⎪⎪ ⎪

⎪ ⎪

⎪ ⎪ ⎪ ⎪⎪ ⎪

⎪ ⎪

⎧⎨⎩

⎫⎬⎭

⎛⎝⎜

⎛⎝⎜

⎞⎠⎟

⎞⎠⎟

⎧⎨⎩

⎫⎬⎭

⎧⎨⎩

⎫⎬⎭

⎛⎝⎜⎜

⎞⎠⎟⎟

⎧⎨⎩

⎫⎬⎭

⎡⎣⎢⎢

⎛⎝⎜

⎞⎠⎟

⎤⎦⎥⎥

us

ddt

xb

bxb

axb s

sxb

s bbb

ab

x

11

11

1

11

1.

16

2

2

τ ττ

τ

τ τ

¯ =+

+

− =+

+

−+

+( )

Note that the time invariant operator s 1τ{ + } can now becanceled with

s1

1{ }

τ +since there is no time varying signal in be-

tween. Let

⎪ ⎪

⎪ ⎪⎧⎨⎩

⎫⎬⎭

⎡⎣⎢⎢

⎛⎝⎜

⎞⎠⎟

⎤⎦⎥⎥z

s bbb

ab

x1

11

,17

2τ τ=

+−

+

( )

then the dynamics from output x to u can be represented by a first-order LPV model

⎛⎝⎜⎜

⎛⎝⎜

⎞⎠⎟

⎞⎠⎟⎟

ub

x z

z zb

bb

ab

x

1,

1 1.

182

τ

τ τ

¯ = −

= − + −

+ □( )

Remark 1. Treat x as the input, u as the output, and a, b, and τ asthe parameters, the BIBO stability of the first order system (13) canbe easily established given that the system has a single frozen-

time pole at 1−τ, and

b1τ,

bb

b

ab

1 12( − + )

τ τ

are bounded.

Remark 2. Note that since u u¯ ≈ for small τ, one can treat (13) asan approximate inverse model of (11). Moreover, u u τ| ¯ − | ∝ ( ),namely, the inverse model error can be made arbitrarily small witha properly chosen τ.

Remark 3. Lemma 1 assumes a continuous-time implementationof the inverse of (11). When (13) is discretized for real engineimplementation, its BIBO stability remains due to Remark 1. If thedelay caused by discretization is small, its performance will not besubstantially affected.

Now the approximate inverse model i1Σ − is given by (13). For

simplicity, one can drop the b b/ 2 term if the parameter variation issubstantially slower than the system dynamics. However, this isnot the case in this application. b b/ 2 as in (7) includes the states.Therefore, b b/ 2 is not slower than the system dynamics. The fol-lowing simulation also verified that b b/ 2 should not be omitted: Tovalidate the first-order sub-model inverse, the two systems Σi and

i1Σ − are connected in cascade as shown in Fig. 7. According to

Remark 2, the output v in Fig. 7 should be close to the input v. Anumerical analysis of the inverse performance is shown in Fig. 8. Itis obvious that the inverse incorporating the b b/ 2 term withsmaller time constant τ has better accuracy. Therefore, the inverseincorporating b b/ 2 is adopted in the subsequent derivation. Thetime constant τ is the tuning parameter for the IMC design. Fig. 8indicates that smaller time constants lead to a better inverse, asexpected.

(3) Structured quasi-LPV inverse: Representing each first-ordermodel inverse in (10) with (13), an inverse model for the nonlinearmodel is derived, which will be referred to as the structured quasi-LPV inverse.

To incorporate the quasi-LPV inverse model in the IMC struc-ture, two implementable configurations are possible as shown inFigs. 9a and b. Firstly, one can use the states in the nonlinearmodel Σ to schedule the parameters in the inverse model 1Σ− (asin Fig. 9a). Since the states used for parameter scheduling areexternal to 1Σ− , Fig. 9a is referred to as the externally scheduledquasi-LPV inverse. Secondly, since all the states in the originalquasi-LPV model are explicit in its inverse structure, one can de-rive parameters used in A and B (defined in (6)) for 1Σ− based onthe internal states in 1Σ− (as in Fig. 9b), which is referred to as theinternally scheduled quasi-LPV inverse.

While the externally scheduled quasi-LPV inverse looks ap-pealing at first, its utility is ruled out after more in-depth analysisand simulation. The dashed line in Fig. 9a which represents thegain scheduling signal forms a feedback loop, which causes in-stability. In this work, the proposed IMC controller uses the in-ternally scheduled quasi-LPV inverse model for its implementa-tion. It should be noted that the internally scheduled quasi-LPVinverse is not possible for inverse LPV model derived by general

Page 7: Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

Fig. 8. Analysis of first-order inverse 11Σ− with and without b (b is derived from

numerically differentiating b).

Fig. 10. Parameter scheduling relationship in the internally scheduled quasi-LPVinverse (blue solid lines indicate that the states are actually used for scheduling;red dotted lines represent the use of the steady state value generated from steadystate mapping in scheduling). (For interpretation of the references to color in thisfigure caption, the reader is referred to the web version of this paper.)

Z. Qiu et al. / Control Engineering Practice 46 (2016) 105–114 111

system inverting methodologies, unless the states are preserved inthe inverse model. With the structured quasi-LPV inverse modelshown in Fig. 6, it is true that all states are explicitly preserved inthe inverse model, thereby making the internally scheduled quasi-LPV inverse implementation possible.

(4) Stability of the structured quasi-LPV inverse: Even thougheach subsystem in the structured quasi-LPV model is BIBO stable,the overall inverse 1Σ− , with internally scheduled parameters, stillhas stability issues due to the feedback loop introduced by thosestate-dependent parameters.

Since the state-dependent parameters are the root cause for theinstability phenomenon in the structured quasi-LPV inversemodel, an in-depth analysis of the parameter scheduling is carriedout in the inverse model. Fig. 10 shows the scheduling of thevarying parameters in each sub-model for the internally scheduledquasi-LPV inverse. One can see that some of the interconnections,shown by the blue solid lines in Fig. 10 from the states x1 and x2, donot introduce additional feedback loops except those within theirown sub-models. Others, that are shown by red dotted line for theparameters scheduled based on x3 and x4, form additional feed-back loops within the structured quasi-LPV inverse. Meanwhile,the errors from 1

1Σ − , 41Σ − , and 32

1Σ − propagate within the structure,leading to complicated dynamic responses. To construct a stableinverse, the scheduling signals x3 (approximate exhaust pressure)and x4 (approximate turbo speed) are replaced by their steady-state values. The steady-state maps are generated with respect todifferent engine speed and throttle opening.

Remark 4. Note that, based on the dual stability property of IMC(Garcia & Morari, 1982), the stability of the closed-loop system canbe assured if P and Q are stable and M¼P. The stability of thequasi-LPV inverse Q can be established given that it is made up bystable sub-models Σi's and i

1Σ − 's through feedforward connec-tions. All feedback loops in Q are eliminated after replacing the

Fig. 9. Validation structures: (a) externally scheduled quasi-LPV inverse validation

gain-scheduling elements that form feedback loops using theirsteady state values. Therefore, the stability of the closed-loopsystem with IMC can be assured using Q 1Σ= − given in Fig. 10.

(5) Structured quasi-LPV inverse model validation: The inversemodel shown in Fig. 10, with some of the state-dependent para-meters replaced by steady-state mapping values, is validatedthrough simulation. The validation is performed by connecting theinverse to the original model as in Fig. 9b and the validated resultsare shown in Fig. 11.

Note that the quality of the inverse model depends on thetuning parameters, which are the time constants τ as in each first-order LPV sub-model inverse 1

1Σ − , 41Σ − , and 32

1Σ − as shown in Fig. 6.Theoretically if the time constants are small, the output has fasterresponses but also potential oscillations during transients. If thetime constants are large, the transient response will be slow. Tworesults with different tuning parameters are shown in Fig. 11,which validates the inverse model. Inverse 1 has the time con-stants 0.1 s, 0.04 s, and 0.02 s in 1

1Σ − , 41Σ − , and 32

1Σ − , respectively.

Inverse 2 has the time constants 0.05 s, 0.04 s, and 0.02 s in 11Σ − ,

41Σ − , and 32

1Σ − , respectively. Inverse 1, which has the larger timeconstant, is more damped than inverse 2 and has little overshoot,which matches the intuition. The time constant for 1

1Σ − is chosento be larger than the others, considering the error in its output x4propagates to x3 and u, as shown in Fig. 10.

4. Performance evaluation of IMC on turbocharged gasolineengine

With the fourth-order nonlinear model and the model inversedeveloped, IMC can be designed by applying the nonlinear modeland its inverse into Fig. 1. The IMC is implemented in continuous-time domain, and is applied to the virtual “plant”, which is a va-lidated Ford proprietary model. The structure of the resultingnonlinear IMC system is shown in Fig. 12. The tuning parametersare the time constants in each first-order LPV sub-model inverse

structure and (b) internally scheduled quasi-LPV inverse validation structure.

Page 8: Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

Fig. 11. Validation of structured quasi-LPV inverse.

Fig. 12. IMC structure with structured quasi-LPV inverse.

4 5 6 7 8 9 10 11 12

140

160

180

200

Time [sec]

Boo

st P

res

[kP

a]

IMC responsePI responsereference

4 5 6 7 8 9 10 11 120

0.5

1

Was

tega

te [−

]

Time [sec]

IMCPI

1.5 2 2.5 3 3.5 4 4.5

100

150

200

250

300

Time [sec]

Boo

st P

res

[kP

a]

IMC responsePI responsereferenceengine speed/10 [rpm]

1.5 2 2.5 3 3.5 4 4.5

0

0.2

0.4

Was

tega

te [−

]

Time [sec]

IMCPI

Fig. 13. Simulation results. (a) IMC performance: constant engine speed. (b) IMCperformance: varying engine speed.

Z. Qiu et al. / Control Engineering Practice 46 (2016) 105–114112

11Σ − , 4

1Σ − , and 321Σ − , as shown in Fig. 6. The time constants are

chosen at 0.1 s, 0.04 s, and 0.02 s, which are the same as in inverse1. The pressure and temperature sensors are assumed to be ac-curate. Performance of the resulting control system is evaluated inthis section together with the robustness analysis with respect todifferent operating conditions and measurement noises.

4.1. Performance evaluation

To evaluate the performance of IMC, some features have to beconsidered (Moulin & Chauvin, 2011):

� The overshoot of Pb has to be minimized to avoid throttle re-closing.

� The pressure oscillation of Pb while tracking a step function isundesirable because they could generate torque oscillations thatare noticeable to the driver.

IMC is compared with a well-tuned PI controller with extensivefeedforward and anti-windup built in, which is referred to as PIcontrol in the context. The system response and control input arecompared in two cases: constant engine speed of 3000 rpm (as inFig. 13a); varying engine speed (as in Fig. 13b). Here varying enginespeed rises from 1500 to 3000 rpm gradually. The throttle openingis 45° in both simulations. In Fig. 13a, it can be observed that IMCachieved a faster reference tracking than PI with less overshoot oroscillation. In Fig. 13b, the IMC response does not overshoot, evenwithout incorporating an explicit anti-windup strategy.

Overall, the nonlinear IMC for the wastegate control of a

turbocharged gasoline engine shows promising performance. Itshows good reference tracking, no steady-state error, no need for aseparate anti-windup design, and intuitive tuning. Its performancematches, and in cases exceeds, that of a well-tuned PI control withextensive feedforward and anti-windup built in.

4.2. Performance in the presence of disturbances

In real applications of boost-pressure control of a turbochargedgasoline engine, the reference and operating points vary. Moreanalysis is performed herein to evaluate the system performancesensitivity with respect to the variation of operating conditions(engine speed and throttle opening). Note that only one set oftuning parameters is used for all the tests.

First, the impact of the engine speed Nen is considered. Two setsof tests are performed: the same Pb reference step at differentengine speed Nen (as in Fig. 14a), constant Pb reference with a stepchange in Nen (as in Fig. 14b), in which case the variation in Nen can

Page 9: Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

7 8 9 10 11 12120130140150160170

Time [sec]

Pre

s [k

Pa]

7 8 9 10 11 120

0.5

1

Was

tega

te

Time [sec]

2000rpm2500rpm3000rpmreference

11.5 12 12.5 13 13.5160

170

180

190

Time [sec]

Pre

s [k

Pa]

11.5 12 12.5 13 13.5

0

0.2

0.4

Was

tega

te [−

]

Time [sec]

referenceIMC response

IMC

Fig. 14. IMC robustness evaluation with respect to engine speed. (a) IMC robust-ness: different engine speed. (b) IMC robustness: varying engine speed, constantreference.

7 8 9 10 11 12120

130

140

150

Time [sec]

Pre

s [k

Pa] 20 deg

45 deg70 degreference

7 8 9 10 11 120

0.5

1

Was

tega

te [−

]

Time [sec]

11.5 12 12.5 13 13.5160

170

180

190

Time [sec]

Pre

s [k

Pa]IMC responsereference

11.5 12 12.5 13 13.5

0

0.2

0.4

0.6

Was

tega

te

Time [sec]

IMC

Fig. 15. IMC robustness evaluation with respect to throttle opening. (a) IMC ro-bustness: different throttle opening. (b) IMC robustness: step change in throttleopening, constant reference.

Z. Qiu et al. / Control Engineering Practice 46 (2016) 105–114 113

be viewed as a disturbance. The step change in Nen is from 2500 to3000 rpm. The throttle opening is 45° in all simulations. The re-sults show that IMC performs well at all engine speeds, and itrejects the disturbance in Nen.

Next, the impact of the throttle opening uth is considered. Twosets of tests are performed: the same Pb reference step at differentthrottle opening uth (as in Fig. 15a), constant Pb reference with astep change in uth (as in Fig. 15b), in which case the variation in uthcan be viewed as a disturbance. The step change in uth is from 45to 30°. The engine speed is 3000 rpm in all simulations. The resultsshow that IMC performs well at all throttle openings, and it rejectsthe disturbance in uth.

4.3. Discussion of applying IMC on turbocharged gasoline enginewith CBV and EGR

A compressor bypass valve (CBV) is a valve that releases theboost pressure when the boost pressure is too high. It is commonlypresent in a turbocharged gasoline engine to prevent compressorsurge. The proposed nonlinear IMC that is applicable to an enginewith CBV. Generally, when the CBV dumps the boost pressure, bydefault the wastegate is set to be open, and the nonlinear IMC isdisengaged. Therefore, the addition of CBV does not affect the

nonlinear IMC design.Exhaust gas recirculation (EGR) is a technique that involves a

valve that recirculates a portion of exhaust gas back to the enginecylinders, which dilutes the air and reduces in-cylinder tempera-ture. It can help reduce the NOx emission. The engine discussed inthis work has no EGR system. For an engine with EGR, the state-space equations (4) will be changed due to the coupling betweenthe boost pressure and exhaust pressure dynamics with the pre-sence of EGR. Therefore, the nonlinear model has to be revisited,and the parameters in the quasi-LPV model will change accord-ingly. The structured quasi-LPV inverse has to be reanalyzed. Ap-plication of nonlinear IMC to a turbocharged gasoline engine withEGR will be considered in future work.

5. Conclusions

A nonlinear IMC design for the turbocharged gasoline engine ispresented. A nonlinear fourth-order dynamics model is adopted inthe controller. The challenges for inverting the nonlinear modelare addressed by: (1) representing the nonlinear dynamics with aquasi-LPV model, (2) exploring the special quasi-LPV model

Page 10: Control Engineering Practiceracelab/static/Webpublication/2016-IFACCEP-ZengQ.pdfAnother possible avenue to exploit the linear IMC design tools for nonlinear systems would be through

Z. Qiu et al. / Control Engineering Practice 46 (2016) 105–114114

structure, (3) using inverse of the simple first-order quasi-LPVmodel, and (4) assuring the stability of the inverse model byeliminating the internal loops. The simulation results, validatedusing a proprietary model as the “plant”, demonstrate the validityof the proposed approach and establish the performance and ro-bustness of the closed-loop system. The proposed nonlinear IMCpresents a more implementable and less computationally ex-pensive design than the other nonlinear IMC methodologies. It ishowever presented in the context of the boost pressure control ofa turbocharged gasoline engine, and has not been generalized yet.Future work in this direction includes applying the nonlinear IMCto the boost pressure control of a plant with EGR and a real plant,expanding this specific nonlinear IMC design into a more generalmethodology, and developing a systematic approach for tuningthe time constants.

References

Alfieri, E., Amstutz, A., & Guzzella, L. (2009). Gain-scheduled model-based feedbackcontrol of the air/fuel ratio in diesel engines. Control Engineering Practice, 17(December (12)), 1417–1425.

Boukezzoula, R., Galichet, S., & Foulloy, L. (2000). Fuzzy nonlinear adaptive internalmodel control. European Journal of Control, 7(December (5)), 541–556.

Brown, M. D., Lightbody, G., & Irwin, G. W. (1997). Nonlinear internal model controlusing local model networks. Control Theory and Applications, IEE Proceedings,144(November (6)), 505–514.

Buckland, J. H. (2009). Estimation methods for turbocharged spark ignition engines(Ph.D. dissertation). Ann Arbor, MI, USA: Department of Electrical Engineeringand Computer Science, University of Michigan.

Calvet, J. P., & Arkun, Y. (1988). Feedforward and feedback linearization of nonlinearsystem and its implementation using internal model control (IMC). Industrial &Engineering Chemistry Research, 27(10), 1822–1831.

Datta, A. (1998). Adaptive internal model control. London: Springer.Economou, C. G., Morari, M., & Palsson, B. O. (1986). Internal model control: Ex-

tension to nonlinear systems. Industrial & Engineering Chemistry Process Designand Development, 25(2), 403–411.

Garcia, C. E., & Morari, M. (1982). Internal model control. A unifying review andsome new results. Industrial & Engineering Chemistry Process Design and De-velopment, 21(2), 308–323.

Guzzella, L., & Onder, C. H. (2010). Introduction to modeling and control of internalcombustion engine systems (2nd ed.). Berlin, Heidelberg: Springer.

Henson, M. A., & Seborg, D. E. (1991). An internal model control strategy for non-linear systems. AIChE Journal, 37(July (7)), 1065–1081.

Hirschorn, R. M. (1979). Invertibility of nonlinear control systems. SIAM Journal onControl and Optimization, 17(2), 289–297.

Hunt, K. J., & Sbarbaro, D. (1991). Neural networks for nonlinear internal modelcontrol. Control Theory and Applications, IEE Proceedings D, 138(September (5)),431–438.

Karnik, A., & Jankovic, M. (2012). IMC based wastegate control using a first-ordermodel for turbocharged gasoline engine. In American control conference (ACC)(pp. 2872–2877), June.

Karnik, A. Y., Buckland, J. H., & Freudenberg, J. S. (2005). Electronic throttle andwastegate control for turbocharged gasoline engines. In American control con-ference (Vol. 7, pp. 4434–4439), June.

Mohammapour, J., Sun, J., Karnik, A., & Jankovic, M. (2013). Internal model controldesign for linear parameter varying systems. in American control conference(ACC) (pp. 2409–2414), June.

Morari, M., & Zafiriou, E. (1989). Robust process control. Englewood Cliffs, NewJersey: Prentice Hall.

Moulin, P., & Chauvin, J. (2011). Modeling and control of the air system of a tur-bocharged gasoline engine. Control Engineering Practice, 19(March (3)),287–297.

Qiu, Z., Sun, J., Jankovic, M., & Santillo, M. (2014). Nonlinear internal model con-troller design for wastegate control of a turbocharged gasoline engine. InAmerican control conference (ACC) (pp. 214–219), June.

Rivera, D. E., Morari, M., & Skogestad, S. (1986). Internal model control: PID con-troller design. Industrial & Engineering Chemistry Process Design and Develop-ment, 25(1), 252–265.

Rugh, W. J., & Shamma, J. S. (2000). Research on gain scheduling. Automatica, 36(October (10)), 1401–1425.

Schwarzmann, D., Nitsche, R., & Lunze, J. (2006). Diesel boost pressure control usingflatness-based internal model control. SAE technical paper, 2006-01-0855.

Schwarzmann, D., Nitsche, R., Lunze, J., & Schanz, A. (2006). Pressure control of atwo-stage turbocharged diesel engine using a novel nonlinear IMC approach. InComputer aided control system design, 2006 IEEE international conference oncontrol applications, 2006 IEEE international symposium on intelligent control,2006 IEEE (pp. 2399–2404), October.

Shafiq, M. (2005). Internal model control structure using adaptive inverse controlstrategy. ISA Transactions, 44(July (3)), 353–362.

Thomasson, A., Eriksson, L., Leufven, O., & Andersson, P. (2009). Wastegate actuatormodeling and model-based boost pressure control. In IFAC workshop on engineand powertrain control, simulation and modeling (Vol. 41).

Toivonen, H. T., Sandström, K. V., & Nyström, R. H. (2003). Internal model control ofnonlinear systems described by velocity-based linearizations. Journal of ProcessControl, 13(April (3)), 215–224.

Xie, W. F., & Rad, A. B. (2000). Fuzzy adaptive internal model control. IEEE Trans-actions on Industrial Electronics, 47(February (1)), 193–202.