Вы находитесь на странице: 1из 9

Florian Zurbriggen1

Institute for Dynamic Systems and Control,


Department of Mechanical
and Process Engineering,
ETH Zurich,
Zurich 8092, Switzerland
e-mail: florianz@ethz.ch

Tobias Ott
Institute for Dynamic Systems and Control,
Department of Mechanical
and Process Engineering,
ETH Zurich,
Zurich 8092, Switzerland
e-mail: toott@ethz.ch

Christopher Onder
Institute for Dynamic Systems and Control,
Department of Mechanical
and Process Engineering,
ETH Zurich,
Zurich 8092, Switzerland
e-mail: onder@idsc.mavt.ethz.ch

Lino Guzzella
Institute for Dynamic Systems and Control,
Department of Mechanical
and Process Engineering,
ETH Zurich,
Zurich 8092, Switzerland
e-mail: lguzzella@ethz.ch

Optimal Control of the Heat


Release Rate of an Internal
Combustion Engine With
Pressure Gradient, Maximum
Pressure, and Knock Constraints
In this paper, we present an analysis of the optimal burn rate in an internal combustion
engine (ICE) considering pressure gradient, maximum pressure, and knocking. A zerodimensional model with heat losses is used for that purpose. The working fluids are
assumed to behave like ideal gases with temperature dependent gas properties. In the first
part, it is assumed that the burn rate can be arbitrarily chosen at every time instance in
order to maximize the mechanical work. This leads to an optimal control problem with
constraints. In the second part, a Vibe type burn rate is assumed, where the center of
combustion, the duration and the form factor can be chosen in order to maximize the mechanical work. This Vibe type burn rate is finally compared with the arbitrary combustion
as the benchmark in order to evaluate the potential of the more realistic burn shape.
[DOI: 10.1115/1.4027592]

Introduction

Internal combustion engines are essential in todays transportation. Many investigations have been conducted over the last several decades in order to improve their efficiencies. This paper
focuses on the shape of the combustion and the corresponding
sensitivities to the efficiency.
Ideal engine cycles are analyzed in the classical literature [1,2].
No heat losses are considered here, and compression and expansion are assumed to be isentropic. The maximum efficiency for a
constant configuration can be achieved with the constant volume
cycle.
In Ref. [3], the optimal piston motion is derived applying optimal control theory to a model with friction losses and with a heat
leakage proportional to the temperature difference between working fluid and cylinder walls. Constraints on acceleration, deceleration, and cycle time have also been considered. In Ref. [4], the
effects of finite combustion rates are considered furthermore.
Numerous publications have used the second law of thermodynamics in their analysis. Literature reviews of second-law
analyses applied to internal combustion engines are given in
Refs. [5,6]. The second law of thermodynamics can be used to
determine the maximum possible performance of thermal systems
and to identify the sources for availability destruction. The entropy generation of an adiabatic ICE is minimized in Ref. [7] with
respect to the piston velocity using Pontryagins Minimum Principle. Those authors showed that the maximization of the expansion
work cannot be solved analytically with the same principles. In
Ref. [8] radiative heat transfer is also considered. The effect of the
1
Corresponding author.
Contributed by the Dynamic Systems Division of ASME for publication in the
JOURNAL OF DYNAMIC SYSTEMS, MEASUREMENT, AND CONTROL. Manuscript received
February 1, 2013; final manuscript received April 29, 2014; published online August
8, 2014. Assoc. Editor: Yang Shi.

burn rate parameters is analyzed in Ref. [9]. The importance of


the parameters is shown using an analysis based on the second
law of thermodynamics. A Vibe type combustion shape is applied
for that purpose. Knock, pressure gradient, and maximum pressure
restrictions were not considered.
Several authors have used finite-time thermodynamics to analyze and optimize engine cycles. The effects of internal irreversibility, heat losses, and friction losses on the performance of the
Otto cycle are analyzed in Ref. [10]. A Diesel cycle model is considered in Ref. [11] that includes heat losses, friction losses and
temperature dependent specific heats. In Ref. [12], a Dual cycle
model containing irreversibilities during compression and expansion is analyzed. A constant mass flow rate of fuel to the cylinder
is assumed. The air-standard Dual cycle and the Dual-Atkinson
cycle are compared with each other in Ref. [13]. The effects of
compression ratio on the thermal efficiency and the power output
are analyzed. Temperature dependent properties of the working
fluid, friction losses and heat losses were considered in the model.
A similar analysis is carried out in Ref. [14]. Here, an irreversible
Miller cycle is assumed. The effects of combustion volume,
displacement volume, and compression ratio on the maximum
power output and thermal efficiency are investigated. The performances of air-standard Atkinson and Otto cycles are compared
in Ref. [15].
The shape of the combustion is prescribed for the cycles analyzed in Refs. [1015]. The combustion in the Diesel cycle is
approximated as constant pressure combustion and the combustion in the Otto cycle is approximated as constant volume combustion. The Dual cycle is a combination of the Otto cycle and the
Diesel cycle. Differing compression and expansion ratios are distinctive for the standard Miller and Atkinson cycles.
In Ref. [16], a zero-dimensional model with one zone is used to
analyze the influence of several engine parameters on power and
efficiency. The heat release rate is modeled as a Vibe function.

Journal of Dynamic Systems, Measurement, and Control


C 2014 by ASME
Copyright V

Downloaded From: http://asmedigitalcollection.asme.org/ on 12/09/2014 Terms of Use: http://asme.org/terms

NOVEMBER 2014, Vol. 136 / 061006-1

The analyzed parameters are the cylinder volume, the stroke bore
ratio, the compression ratio, the wall temperature, the fuel-to-air
ratio, and the outbreak of the combustion.
In Ref. [17], the influence of variable heat capacities of the
working fluid on the performance of a Diesel heat engine is analyzed. The dependence of the heat capacities on the temperature is
approximated as an affine function. A more realistic approach for
the heat capacities using a polynomial approximation of the heat
capacities is presented in Refs. [18] and [19]. A spark ignition
engine is analyzed in Ref. [18] and a compression ignition engine
is analyzed in Ref. [19].
In Ref. [20], a zero-dimensional, single-zone model is presented
that includes temperature dependent specific heats and heat losses
through the walls. The heat release rate is modeled according to a
modified Vibe function. The influence of engine speed, injection
timing, and compression ratio on the performance is analyzed.
Compared to the literature, we maximize the mechanical work
of an ICE under consideration of the following constraints:
(1) The maximum pressure is restricted to a given value pmax
in order to avoid causing any mechanical damage.
(2) The pressure gradient is limited in order to minimize the
combustion noise.
(3) Knocking has to be avoided.
The mechanical work Wmech is given as

Wmech dW p  dV

can be divided into single-zone or multi-zone models. In this paper, a two-zone model is derived that is also able to predict knocking. The gas properties of the different gases in the cylinder are
polynomials in the temperature T. The instantaneous mixture
depending on the burn progress is considered.
Figure 1 shows the system boundaries and the corresponding
flows of energy and mass. The details of the process simulation
used including heat losses dQw/dt through the walls are given in
Appendix A. The system comprises three states:

The cylinder temperature T(t).


The burn progress xb(t), which is the so-called mass fraction
burnt (MFB).
The knock state I(t) representing the progress of auto
ignition.

For the sake of readability, the time dependency is omitted


throughout the rest of the text. The input to the system is the heat
release rate (HRR) dQb/dt, which is normalized to the total chemical energy content
u

dxb
1
dQb

mf ;1  Hl dt
dt

(2)

where mf,1 is the fuel mass, which is available in the cylinder at


the beginning of the compression, and Hl is the lower heating
value of the fuel.
(1)

Furthermore, we do not restrict the heat release rate. The heat


release rate can be chosen arbitrarily at every instance. This complicates the problem, because the heat release rate is not a function
of a finite number of parameters anymore. As it is explained later
in this text, the problem can be formulated as an optimal control
problem.
A zero-dimensional model, sometimes also referred to as thermodynamic model, is used to describe the thermodynamics in the
combustion chamber [1,2,21,22]. The cylinder volume is assumed
to be ideally mixed in zero-dimensional models. The disadvantage
of such models is that they are not able to capture local phenomena. However, there are several advantages that justify the usage
of such models in the context of this paper. Zero-dimensional
models are computationally efficient. At the same time, they are
meaningful for energy system evaluations and therefore for extensive parametric studies [2,16,20,2325]. Zero-dimensional models

1.1 Outline. Section 2 introduces the system dynamics. The


problem of maximizing the mechanical work under consideration
of the given constraints is then formulated as an optimal control
problem. The optimization problem for maximizing the work
using a Vibe type combustion is formulated as well.
In Sec. 3, we present a methodology to solve the optimal control problem using an adaptive input and reformulating the knock
criterion using a Lagrange multiplier.
In Sec. 4, the results are presented and the various constraints
are discussed.

Problem Description

2.1 System Dynamics. The differential equation for the burn


progress is given as
fb u :

dxb
u
dt

(3)

The dynamics of the cylinder temperature T are derived in


Appendix A.1 and is given as



dT
1
fT T; xb ; u:

 Hl  mf ;1 ufc  mfc;1  ub  mfc;1  u


dt m  cv

dQw
dV
(4)
p

dt
dt
where dxb/dt u is already considered. The mass mfc,1 is the mass
of fresh charge at the beginning of the compression. The temperature T and the pressure p are related through the ideal gas law,
since the volume V is always known (Appendix A.2). The variables ufc and ub represent the mass specific inner energies of the
fresh charge and the burnt gas. The variable cv represents the
specific heat of the total cylinder mass.
The variable I is introduced as the knock state indicating the
auto-ignition progress in the unburnt fuel. The differential equation of I is defined as

Fig. 1 System boundaries (dashed) of the zero-dimensional


model and the corresponding flows of energy and mass

8
1
dI <
fI p; xb : sp
dt :
0

061006-2 / Vol. 136, NOVEMBER 2014

Downloaded From: http://asmedigitalcollection.asme.org/ on 12/09/2014 Terms of Use: http://asme.org/terms

xb < 0:75

(5)

else

Transactions of the ASME

A detailed derivation of the reaction rate s can be found in


Appendix A.4. At the beginning of the compression, the knock
state I(t 0) is set to zero. It is assumed that auto ignition is not
crucial anymore if more than 75% of the fuel is burnt [26].
2.2 Optimal Control Problem of the Arbitrary Heat
Release Rate. Maximizing the mechanical work is formulated as
the following optimal control problem:

(6)
max dW
u

s:t:

dT
fT T; xb ; u
dt
dxb
u
dt

dI
fI p; xb
dt
p  V m  Rxb  T

(7)
/c : /xb 0:5
(8)
(9)

/d : /xb 0:9  /xb 0:05

(10)

(13)

/c ;/d ;mv

u0

(14)

s:t:

I1

(15)

xb t1 It1 0

(16)

max W/c ; /d ; mv

The no-knock condition (15) can be rewritten as


Itf  1

(19)

because the time derivative of the knock state I is nonnegative


(Eq. (5)) and the state itself therefore increases monotonically in
time. This reformulation of the knock criterion is used in Sec. 3.3.
The variable tf represents the time when the high pressure cycle is
over, while t1 represents the start of compression. The index 1 is
used instead of 0 because of the definition of point 1 in the classical literature [1,2]. The constraints on the pressure and on the
pressure gradient are formulated in Eqs. (11) and (12).
The constraint (13) limits the chemical energy released overall
to the available amount of fuel. The constraint (14) can be
explained by the fact that already burnt fuel cannot be converted
back. The pressure p(t1) defines the initial temperature T(t1) (ideal
gas law (10)) and the total cylinder mass m.

/v

(25)
(26)
(27)

Methodology

The continuous-time optimal control problem (6)(17) is solved


using dynamic programming (DP). The algorithm used is
described in Ref. [27]. The following sections present the reformulations of the problem in order to improve the solution. A validation of the reformulations is given and the method used to solve
the Vibe problem (24)(27) is presented.
3.1 Discretization. The problem is discretized using an Euler
forward discretization
X
(28)
dWk
max
uk

s:t:

The solution of the arbitrary combustion rate is a benchmark,


representing the maximum achievable efficiency for a given situation. In order to evaluate a more realistic burn shape, the variable
xb is chosen to be a Vibe function [1]
xb 1  e6:908

(24)

The function W/c ; /d ; mv returns the cycle work when the


Vibe heat release function is applied with the corresponding
values of /c ; /d , and mv.

Optimal Vibe Combustion

// mv 1

p  pmax

dp
 Dpmax
dt
Itf  1

(17)

Throughout the rest of the text, the notation Dpmax is used to


express the maximum allowed pressure gradient

dp
Dpmax : 
(18)
dt max

(23)

The problem of finding the optimal Vibe combustion considering the various constraints can be written as an optimization
problem

xb  1

(12)

(22)

while /d is defined as the angle between 5% and 90% of fuel


burnt

(11)

given

(21)

The three parameters mv, /d , and /0 can be chosen arbitrarily


in order to maximize the mechanical work. The constraints on
maximum pressure gradient, maximum pressure and knock have
to be fulfilled. Below, the center /c and the duration /d are used
to show the results. The variable /c is the angle at which one half
of the fuel is burnt

p  pmax

dp dp
 
dt
dt max

pt1

2.3

dxb @xb @/ @xb


x
dt
@/ @t
@/


// mv 1
0
mv 1 /  /0
 e6:908 /v

6:908  x 
/v
/v

pk1 Fp pk ; xb;k ; uk


xb;k1 Fb xb;k ; uk

(30)

Ik1 FI pk ; xb;k ; Ik

(31)

pk  pmax

(32)

Dpk  Dpmax

(33)

xb;k  1

(34)

uk  0

(35)

IN  1

(36)

xb;1 I1 0
p1

given

(29)

(37)
(38)

(20)
where the pressure gradient is approximated as

where / is the crank angle, /0 is the start of combustion, /v influences the length, and the form factor mv defines the shape of the
combustion. The normalized heat release rate is given as
Journal of Dynamic Systems, Measurement, and Control

Downloaded From: http://asmedigitalcollection.asme.org/ on 12/09/2014 Terms of Use: http://asme.org/terms

Dpk

pk1  pk
h

(39)

NOVEMBER 2014, Vol. 136 / 061006-3

fgk fgt k  h

k 1; 2; N

(40)

The pressure update function (29) is derived by inserting the


ideal gas law (pk  Vk m  Rxb;k  Tk ) into the temperature
update function Tk1 FT(Tk, xb,k, uk), which is the discretization
of Eq. (7). This simplifies the formulations introduced in Sec. 3.2.
3.2 Adaptive Input. The input u as defined in Eq. (8) is
unbounded (u [0; 1)). This complicates the use of dynamic
programming in order to solve the discrete-time optimal control
problem (28)(38). A normalized input u~k 2 0; 1 can be defined
in order to reduce the numerical issues around feasibility boundaries. For this purpose, the pressure state update (as a function of
u~k ) is redefined as the interpolation between the maximum feasible state update (~
uk 1), which is the maximum reachable pressure feasible at the next time step, and the minimum feasible state
update (~
uk uk 0) representing no burn progress

minp
 xb ; pDp ; pmax  u~k 1
(41)
F~p pk ; xb;k ; u~k :
pk1 pk ; xb;k ; uk 0 u~k 0
where pxb is defined as the pressure that is reached in the next time
step (tk1) by burning all the remaining fuel
pxb : pk1 xb;k1 1

GU  0

s:t:

where h is the step size of the discretization (h tk 1 tk) and the


index k is the time index:

(53)

For the sake of simplicity, the state updates for pk and xb,k will
be omitted in the following formulations. The sequence U is
called an input sequence, which is defined as
U fl1 ; l2 ; lN1 g

lk 2 0; 1

(54)

The objective function F U, the negative work, is given as



n X
o

uk lk
F U 
dWk ~
(55)
The state update function of Ik can be rewritten as


Ik1 Ik DFI pk ; xb;k

(56)

which is used in the inequality constraint (53) to formulate the noknock condition

nX
o



uk lk
GU
DFI pk ; xb;k  1~
(57)
One way to deal with inequality (and equality) constraints is
the classical Lagrangian notion (see Ref. [28]). In this case, the
Lagrangian function is
LU; k F U k  GU

(58)

(42)

and pDp is defined as the pressure that is reached at time step k 1


when the pressure gradient at step k is equal to the maximum feasible pressure gradient

where k is a Lagrange multiplier. According to Ref. [28], U  ; k


represents an optimal solution-Lagrange multiplier pair, and U 
an optimal solution to the problem formulated in Eqs. (52) and
(53), if and only if the following four conditions are fulfilled:

pDp : pk1 Dpk Dpmax

GU   0 primal feasibility

(59)

k  0 dual feasibility

(60)

(43)

The problem ((27)(38) can be rewritten using definition (41)


X
max
(44)
dWk

u~k



pk1 F~p pk ; xb;k ; u~k


uk
xb;k1 Fb xb;k ; uk ~

s:t:

(45)
(46)

Ik1 FI pk ; xb;k ; Ik

(47)

u~k 2 0; 1

(48)

IN  1

(49)

xb;1 I1 0

(50)

p1

given

(51)

k  GU  0

3.3.1 Lagrangian Formulation. The optimal control problem


introduced in Eqs. (44)(51) is first formulated as an N 1dimensional optimization problem with an inequality constraint
F U
min
U

~
pk1 Fp pk ; xb;k ; u~k 
xb;k1 Fb xb;k ; uk ~
uk

(52)

complementary
slackness

(61)
(62)

3.4 Numerical Solution of the Arbitrary Combustion. For


the Lagrangian optimality condition (61), the minimization of
LU; k for a constant value of k can be formulated as a twodimensional (2D) optimal control problem
X


(63)
dWk k  DFI pk ; xb;k
min
u~k



pk1 F~p pk ; xb;k ; u~k


uk
xb;k1 Fb xb;k ; uk ~

s:t:

This reformulation guarantees that the constraints (32)(35) are


always fulfilled.
3.3 Eliminating the Knock State by Introducing a
Lagrange Multiplier. The calculation time required for solving
optimal control problems with algorithms such as dynamic programming increases exponentially with the number of states. This
section shows how this problem can be alleviated by eliminating
the knock state.

Lagrangian
optimality

U  argmin LU; k

u~k 2 0; 1
xb;1 0
p1

given

(64)
(65)
(66)
(67)
(68)

The elimination of the knock state is possible because the stateupdate functions (45) and (46) for pk and xb,k are independent of
the knock state Ik. Dynamic programming is used to solve the
problem presented in Eqs. (63)(68). Let U dp k be the associated
optimal input sequence as a function of k.
The value of k* is found using the following algorithm in order
to fulfill the conditions (59), (60), and (62):
(1) Shooting: k 0:
(a)
Calculate U dp k 0


(b)
Check G U dp k 0  0 (condition (59)):

061006-4 / Vol. 136, NOVEMBER 2014

Downloaded From: http://asmedigitalcollection.asme.org/ on 12/09/2014 Terms of Use: http://asme.org/terms

Transactions of the ASME

Table 1 The three cases solved with DP


Case

dp=dtjmax bar= deg

pmax (bar)

Ik jmax ()

(i)
(ii)
(iii)

3
3
3

1
100
100

1
1
1

Table 2

3D 1
3D 2
3D 3
3D 4
3D 5
3D 6
2D

(2)

Results of the comparison

#of mode leval:


time step

g (%)

IN ()

1280000
5120000
10240000
40960000
81920000
163840000
896000

37.47
37.94
38.23
38.47
38.57
38.70
39.13

0.772
0.901
0.880
0.953
0.938
0.939
0.999

If true: k* 0
Else proceed with 2.
Note that the conditions (60) and (62) are fulfilled if
Eq. (59) is true and k 0.
Solve the fixed point problem for k* > 0


(69)
G U dp k 0
Note that the conditions (59), (60), and (62) are fulfilled, if Eq. (69) has a solution for k* > 0.

3.5 Evaluation of the Reformulation. This section compares


the results obtained with the proposed methodology, which is in
the following labeled as 2D indicating the two states, with the solution of the initial three-dimensional (3D) problem (28)(38), if
it is solved directly using dynamic programming. In the following,
the latter is called 3D. The case (iii) defined in Table 1 is analyzed
in order to evaluate the reformulation. The operating point used is
the same as the one in Sec. 4.1. For the DP used to solve 3D, six
different cases are considered for increasing numbers of grid
points. Table 2 shows the associated number of model evaluations
per time step, the efficiency of the high pressure cycle and the
final knock state. In the 2D column, the number (#) of model evaluations per time step accounts for the number of iterations used
for the fixed-point problem.
The solution of 3D can hardly reach the optimal solution even
for a very fine discretization. The final knock state of the optimal
solution would for the considered operating point be IN 1, since
the constraint (53) is active in this case. The final knock state IN
of the method 2D is therefore found by numerically solving the
fixed point problem (69). The corresponding final error of IN,
which is in this case smaller than 0.001 (see Table 2), is a consequence of the chosen tolerance of the numerical solver of the fixed
point problem.
Eliminating the knock state significantly reduces the calculation
effort, which on the other hand allows a finer discretization of the
other two states and the input.
3.6 Numerical Solution of the Vibe Combustion. The Vibe
combustion described in Sec. 2.3 is evaluated at a precise discretization of the variables /c ; /d , and mv in order solve the problem
(24)(27). Figure 2 shows an example of the efficiencies as a
function of the center and the duration of combustion for the
unconstrained case and a constant mv 4. The same operating
point as in Sec. 4.1 (2000 rpm and p1 2 bar) is considered. The
maximum efficiency is reached if the center of combustion is at

Fig. 2 Efficiencies of Vibe combustion depending on center


and duration of combustion if the constraints (24)(27) are not
considered

approximately 8 deg after top dead center (TDC), which is often


mentioned in literature as a reasonable value for the torque maximizing center of combustion (see e.g., Ref. [2]).

Results and Discussions

This section first investigates the influence of the constraints on


the arbitrary combustion. Second, the sensitivity of the Vibe parameters on the maximum possible work is analyzed. Finally, the
realistic Vibe type combustion is compared to the benchmark,
namely the best possible arbitrary combustion. The parameters
used can be found in Appendix B.
4.1 Influence of the Different Constraints on the Arbitrary
Combustion. In the following, the influence of the various constraints is analyzed using an example. Three different cases are
compared:
(i) Only the pressure gradient is considered.
(ii) Pressure gradient and maximum pressure are considered.
(iii) Pressure gradient, maximum pressure and knocking are
considered.
The constraints for all of these three cases are listed in Table 1.
Figure 3 shows the corresponding results for an engine speed of
2000 rpm and an initial pressure p1 2 bar. The maximumpressure constraint introduced in (ii) leads to a later center of combustion. The center of combustion is also delayed when a limit of
the auto-ignition progress is introduced (case (iii)). These effects
express themselves in decreasing efficiencies (see Table 3).
Figure 3 shows that the heat release rate always stays at the
maximum feasible limit, because one constraint is always active.
The solution for this case can be summarized as follows:

Burning as fast as allowed regarding the pressure gradient


and maximum pressure constraints.
The start of combustion is shifted in order to maximize the
mechanical work and, if it occurs, to eliminate knocking.

This interpretation of the results motivates the analysis presented in Sec. 4.3, where a more realistic burn rate with fewer
degrees of freedom is compared to that obtained with the arbitrary
combustion.
4.2 Sensitivity on the Vibe Parameters. The same operating
point as in Sec. 4.1 above is considered, namely an engine speed
of 2000 rpm and a pressure p1 2 bar.
Figure 4 shows the sensitivities of the Vibe parameters on the
maximum possible efficiency, illustrated if one parameter is kept
constant in each case. The form parameter mv shows the least

Journal of Dynamic Systems, Measurement, and Control

Downloaded From: http://asmedigitalcollection.asme.org/ on 12/09/2014 Terms of Use: http://asme.org/terms

NOVEMBER 2014, Vol. 136 / 061006-5

Fig. 3 Results of DP comparing the influence of the constraints


Table 3

g (%)

The efficiencies for the three cases


(i)

(ii)

(iii)

40.93

40.33

39.13

influence over a broad range, where the maximum possible efficiency is almost constant. But small values of mv lead to high heat
release rates at the beginning of the combustion, which delays the
combustion in order to meet the maximum pressure and pressure
gradient constraints. This delay of the combustion results in a
reduced efficiency. The first two subplots in Fig. 4 show that the
duration and the center of the combustion have the higher influence on the maximum possible efficiency.
Figure 5 shows the maximum possible efficiencies as a function
of /c and /d , limited by the three constraints. The form factor is
kept constant as mv 4, which is a reasonable value (see Fig. 4).
Depending on the operating point, different constraints may be
active. For the operating point shown, the knock constraint and
the pressure gradient constraint are active if the best possible center and duration of combustion are chosen.
It should be mentioned that the best efficiency can only be
reached if the center and the duration of the combustion can be
chosen arbitrarily.
4.3 Comparison Between Vibe and Arbitrary Combustion.
The wide range of operating points of the internal combustion
engine is investigated in this section. The Vibe combustion with
optimized parameters is benchmarked against the optimal arbitrary combustion. Figure 6 shows the comparison for 2000 rpm
and 5000 rpm, whereas the load is varied from pmi 5 bar to
pmi 25 bar. This variation of operating points represents slow
and fast engine speeds, as well as low and high torques. The constraints used are given in case (iii) of Table 1. The indicated mean
effective pressure for an engine with the displacement volume Vd
is given as

Fig. 4 Maximum possible efficiencies depending on the parameters mv, /c , and /d . The gray area denotes the region,
where no feasible configuration is possible due to knocking
and pressure-gradient limitations (see Fig. 5).

dW
pmi

Vd

(70)

Only the high pressure cycle is considered here. The indicated


mean effective pressure is the engine work of one cycle normalized by the displacement volume. The indicated mean effective
pressure is proportional to the engine torque.
For low loads, the efficiency of the benchmark (arbitrary combustion) can almost be reached by the Vibe combustion. For the

061006-6 / Vol. 136, NOVEMBER 2014

Downloaded From: http://asmedigitalcollection.asme.org/ on 12/09/2014 Terms of Use: http://asme.org/terms

Transactions of the ASME

dU dQb dQw
dV

m_ in  hin  m_ out  hout



p
dt
dt
dt
dt

(A1)

where the change of the inner energy dU/dt is given by the heat
released in the burnt fuel dQb/dt, the heat transfer through the
walls dQw/dt, the mechanical power p  dV=dt, and the enthalpies hin and hout of the appropriate mass flows m_ in and m_ out
through the system boundaries. The change of inner energy is
!
dU d
d X
m  u
mi  ui
dt
dt
dt
i
X dmi
dT

(A2)
 ui m  cv 
dt
dt
i
Fig. 5 Efficiencies depending on center and duration of combustion for a form factor mv 5 4

The index i in mi, ui, and cv,i stands for the various components
in the cylinder. Variables without an index refer to the total cylinder mass m, and u is the mass-specific inner energy and cv is the
specific heat. There are two gases assumed to be in the cylinder,
namely burnt gas (with mass mb) and fresh charge (with mass
mfc). The related time derivatives are proportional to the (normalized) heat release rate dxb/dt
dmb
dxb
mfc;1 
dt
dt
dmfc
dxb
mfc;1 
dt
dt

Fig. 6 Comparison between arbitrary and Vibe combustion for


two engine speeds 2000 rpm and 5000 rpm

case of higher loads, the Vibe combustion is worse, especially at


low speeds. The difference in efficiency is then around 1%, which
is still not very much. This small difference indicates that the efficiency potential for a given configuration can in large part be
exploited if the duration and the center of combustion can be
chosen arbitrarily.

Conclusion

This paper investigates optimal heat release rates in internal


combustion engines. The mechanical work is maximized considering constraints on the maximum pressure, on the maximum
pressure gradient and on knocking. The heat release rate is
assumed as the input variable. A zero-dimensional model with
heat losses and temperature dependent gas properties is used. The
influence of the various constraints on the maximum possible efficiency is analyzed as well.
The best possible arbitrary combustion is finally compared with
a more realistic Vibe combustion, where center, duration and
form of combustion are chosen such as to maximize the mechanical work under consideration of the same constraints as above.
The Vibe-combustion efficiency turns out to be close to the efficiency of the arbitrary combustion over a broad range of operating
points if center and duration can be chosen independently.

Appendix A: Process Simulation Model

(A3)
(A4)

where mfc,1 is the mass of fresh charge at the beginning of the


compression. The gas properties are approximated as polynomials
in T. The coefficients for nitrogen N2 and water vapor H2O are
taken from Ref. [29]. The coefficients for the fuel C8H16, oxygen
O2 and carbon dioxide CO2 are taken from Ref. [30].
The approach chosen for the heat losses dQw/dt is described in
Appendix A.3. The mass flows are assumed to be zero, because
only the high pressure cycle is analyzed in this paper, i.e., the period when the valves are closed and the engine is fired. For the
sake of simplicity, both valves are assumed to be closed between
180 deg crank angle before and after TDC.
Combining Eqs. (2), (A1)(A4) leads to the differential equation of the cylinder temperature T


dT
1
dxb 


 Hl  mf ;1 ufc  mfc;1  ub  mfc;1
dt m  cv
dt

dQw
dV
(A5)
p

dt
dt

A.2 Piston Movement


Since the engine is assumed to be operated at a constant speed
x, the crank angle / is given as
/ x  t /0
The piston position s [0; H] and movement are given as


1
1 r
s  H  1  cos /   sin2 /
2
2 l


ds 1
r
 H  x  sin /  sin /  cos /
dt 2
l

(A6)

(A7)

(A8)

The first law of thermodynamics for the system shown in Fig. 1


is given as

where H is the engine stroke, r is the crank radius, and l is the


length of the connecting rod. The cylinder volume V and the
corresponding time derivative are


s
1

(A9)
V Vd 
H e1

Journal of Dynamic Systems, Measurement, and Control

NOVEMBER 2014, Vol. 136 / 061006-7

A.1 Energy Conservation

Downloaded From: http://asmedigitalcollection.asme.org/ on 12/09/2014 Terms of Use: http://asme.org/terms

dV Vd ds


dt
H dt

Name

where Vd is the displacement volume of the cylinder and  is the


compression ratio
e

Vd Vc
Vc

(A11)

and Vc is the clearance volume.

A.3 Heat Losses


The heat losses are modeled as
dQw
a  Aw  T  Tw
dt

a 130  d0:2  p0:8  T 0:53  C1   0:8

(A13)

C2 Vd  T1
 cm

 p  pd
C1 p1  V1

(A14)

The variable d represents the cylinder bore, p is the cylinder


pressure, Vd is the displacement volume, and C1 and C2 are constants. The variable pd is the pressure trajectory in the dragged
case. The mean piston speed cm for an engine speed x is given as
xH
p

(A15)

The ratio of the characteristic velocity  to the mean piston


speed cm is usually between 0 and 3. In this case, it is assumed as
 1:5  cm

(A16)

A.4 Knock Prediction


A phenomenological model is used to predict the occurrence of
knock [1,2]. The ignition delay s for a constant pressure pu and a
constant temperature Tu of the unburnt fuel is given by the following Arrhenius function:

ON 3:402 1:7 3800=Tu
pu  e
s 0:01768 
100

(A17)

where ON is the octane number of the fuel. In the case of timevarying pressure pu and temperature Tu, the ignition delay tign is
given by the condition
tign

1
 dt 1
sp
u ; Tu
t0

Octane number
Specific heat ratio
for knock prediction
Compression ratio
ratio crank radius
to rod length
Cylinder bore
Stroke
Woschni constant 1
Woschni constant 2
Lower heating value

Parameters of the model


Param.

Value

Units

ON

98

()

j


1.3
10

()
()

r/l
d
H
C1
C2
Hl

0.3
85
66
2.742
3.24  103
43

()
(mm)
(mm)
()
()
MJ=kg

(A12)

where Tw is the cylinder wall temperature (assumed to be constant), T is the cylinder temperature, Aw is the in-cylinder surface
area, and the heat-transfer coefficient a is defined as proposed by
Woschni [2]

cm

Table 4

(A10)

(A18)

The pressure is assumed to be equally distributed in the cylinder (pu p). The progression of the temperature Tu in the unburnt
zone can be approximated by an isentropic compression
 j1
pu j
(A19)
Tu T1 
p1
where T1 and p1 are the temperature and pressure at the beginning
of the compression, and j is the specific heat ratio, which is
assumed to be constant for the unburnt fuel. The ignition delay s
can be expressed as a function with pu, (which is equal to p) as the
only argument if Eq. (A19) is inserted in Eq. (A17).

Appendix B: Parameters
The values used for the different parameters are summarized in
Table 4.

References
[1] Heywood, J. B., 1988, Internal Combustion Engine Fundamentals, McGrawHill, International, New York.
[2] Pischinger, R., Kell, M., and Sams, T., 2009, Thermodynamik Thermodynamik
der Verbrennungskraftmaschine, 3rd ed., Springer, Bonn.
[3] Mozurkewich, M., and Berry, R. S., 1982, Optimal Paths for Thermodynamic
Systems: The Ideal Otto Cycle, J. Appl. Phys., 53(1), pp. 3442.
[4] Hoffmann, K. H., Watowich, S. J., and Berry, R. S., 1985, Optimal Paths for
Thermodynamic Systems: The Ideal Diesel Cycle, J. Appl. Phys., 58(6),
pp. 21252134.
[5] Caton, J. A., 2000, A Review of Investigations Using the Second Law of Thermodynamics to Study Internal-Combustion Engines, SAE Paper No. 2000-011081.
[6] Rakopoulos, C., and Giakoumis, E., 2006, Second-Law Analyses Applied to
Internal Combustion Engines Operation, Prog. Energy Combust. Sci., 32(1),
pp. 247.
[7] Teh, K.-Y., and Edwards, C. F., 2008, An Optimal Control Approach to
Minimizing Entropy Generation in an Adiabatic Internal Combustion Engine,
J. Dyn. Syst., Meas., Control, 130(4), p. 041008.
[8] Ge, Y., Chen, L., and Sun, F., 2012, Optimal Path of Piston Motion of Irreversible Otto Cycle for Minimum Entropy Generation With Radiative Heat Transfer
Law, J. Energy Inst., 85(3), pp. 140149.
[9] Caton, J. A., 2000, The Effect of Burn Rate Parameters on the Operating
Attributes of a Spark-Ignition Engine as Determined From the Second Law of
Thermodynamics, Spring Technical Conference of the ASME ICE Division.
[10] Ge, Y., Chen, L., and Sun, F., 2008, Finite-Time Thermodynamic Modeling
and Analysis of an Irreversible Otto-Cycle, Appl. Energy, 85(7), pp. 618624.
[11] Ge, Y. L., Chen, L. G., Sun, F. R., and Wu, C., 2007, Performance of Diesel
Cycle With Heat Transfer, Friction, and Variable Specific Heats of Working
Fluid, J. Energy Inst., 80(4), pp. 239242.
[12] Ozsoysal, O. A., 2009, Effects of Combustion Efficiency on a Dual Cycle,
Energy Convers. Manage., 50(9), pp. 24002406.
[13] Gahruei, M. H., Jeshvaghani, H. S., Vahidi, S., and Chen, L., 2013,
Mathematical Modeling and Comparison of Air Standard Dual and DualAtkinson Cycles With Friction, Heat Transfer and Variable Specific-Heats of
the Working Fluid, Appl. Math. Modell., 37(1213), pp. 73197329.
[14] Ebrahimi, R., and Hoseinpour, M., 2013, Performance Analysis of Irreversible
Miller Cycle Under Variable Compression Ratio, J. Thermophys. Heat Transfer, 27(3), pp. 542548.
[15] Hou, S.-S., 2007, Comparison of Performances of Air Standard Atkinson and
Otto Cycles With Heat Transfer Considerations, Energy Convers. Manage.,
48(5), pp. 16831690.
[16] Descieux, D., and Feidt, M., 2007, One Zone Thermodynamic Model Simulation of an Ignition Compression Engine, Appl. Therm. Eng., 27(8),
pp. 14571466.
[17] Zhao, Y., and Chen, J., 2007, Optimum Performance Analysis of an Irreversible Diesel Heat Engine Affected by Variable Heat Capacities of Working
Fluid, Energy Convers. Manage., 48(9), pp. 25952603.
[18] Abu-Nada, E., Al-Hinti, I., Akash, B., and Al-Sarkhi, A., 2007,
Thermodynamic Analysis of Spark-Ignition Engine Using a Gas Mixture
Model for the Working Fluid, Int. J. Energy Res., 31(11), pp. 10311046.
[19] Sakhrieh, A., Abu-Nada, E., Al-Hinti, I., Al-Ghandoor, A., and Akash, B.,
2010, Computational Thermodynamic Analysis of Compression Ignition
Engine, Int. Commun. Heat Mass Transfer, 37(3), pp. 299303.
[20] Menacer, B., and Bouchetara, M., 2013, Numerical Simulation and Prediction
of the Performance of a Direct Injection Turbocharged Diesel Engine, Simulation, 89(11), pp. 13551368.
[21] Merker, G. P., Schwarz, C., Stiesch, G., and Otto, F., 2006, Simulating Combustion, Springer-Verlag, Berlin, Heidelberg.

061006-8 / Vol. 136, NOVEMBER 2014

Downloaded From: http://asmedigitalcollection.asme.org/ on 12/09/2014 Terms of Use: http://asme.org/terms

Transactions of the ASME

[22] Stiesch, G., 2003, Modeling Engine Spray and Combustion Processes,
Springer-Verlag Berlin, Heidelberg.
[23] Scappin, F., Stefansson, S. H., Haglind, F., Andreasen, A., and Larsen, U.,
2012, Validation of a Zero-Dimensional Model for Prediction of NOx and
Engine Performance for Electronically Controlled Marine Two-Stroke Diesel
Engines, Appl. Therm. Eng., 37, pp. 344352.
[24] Payri, F., Olmeda, P., Martin, J., and Garcia, A., 2011, A Complete 0D Thermodynamic Predictive Model for Direct Injection Diesel Engines, Appl.
Energy, 88(12), pp. 46324641.
[25] Asay, R. J., Svensson, K. I., and Tree, D. R., 2004, An Empirical, MixingLimited, Zero-Dimensional Model for Diesel Combustion, SAE Paper 200401-0924.

[26] Franzke, D. E., 1981, Beitrag zur Ermittlung eines Klopfkriteriums der ottomotorischen Verbrennung und zur Vorausberechnung der Klopfgrenze, Ph.D.
thesis, TU M
unchen.
[27] Sundstr
om, O., and Guzzella, L., 2009, A Generic Dynamic Programming
Matlab Function, Control Applications, (CCA) Intelligent Control, (ISIC),
2009 IEEE, St. Petersburg, Russia, July 810, pp. 16251630.
[28] Bertsekas, D. P., 1995, Nonlinear Programming, Athena Scientific, Belmont,
MA.
[29] Moran, M. J., and Shapiro, H. N., 1995, Fundamentals of Engineering Thermodynamics, Wiley, New York.
[30] 2004, Thermodynamic Tables: Hydrocarbons and Non-Hydrocarbons,
http://www.trc.nist.gov/tables/trctables.htm

Journal of Dynamic Systems, Measurement, and Control

NOVEMBER 2014, Vol. 136 / 061006-9

Downloaded From: http://asmedigitalcollection.asme.org/ on 12/09/2014 Terms of Use: http://asme.org/terms

Вам также может понравиться