Anda di halaman 1dari 11

2015-01-0882

Published 04/14/2015
Copyright 2015 SAE International
doi:10.4271/2015-01-0882
saeeng.saejournals.org

Computing Optimal Heat Release Rates in Combustion Engines


Lars Eriksson and Martin Sivertsson
Linkoping Univ.

ABSTRACT
The combustion process has a high impact on the engine efficiency, and in the search for efficient engines it is of interest to study the
combustion. Optimization and optimal control theory is used to compute the most efficient combustion profiles for single zone model
with heat transfer and crevice effects. A model is first developed and tuned to experimental data, the model is a modification of the well
known Gatowski et al.-model [1]. This model is selected since it gives a very good description of the in-cylinder pressure, and thus the
produced work, and achieves this with a low computational complexity. This enables an efficient search method that can maximize the
work to be developed. First, smooth combustion profiles are studied where the combustion is modeled using the Vibe function, and
parametric optimization is used to search for the optimal profile. Then, the most efficient combustion process with a completely free
combustion is studied with theory and software for optimal control. A parameter study is performed to analyze the impact of crevice
volume and air/fuel ratio . The results show that the losses have a high impact on the behavior, which is natural, and that the crevice
effect has a very distinct effect on the optimal combustion giving a two mode appearance similar to the Seiliger cycle.

CITATION: Eriksson, L. and Sivertsson, M., "Computing Optimal Heat Release Rates in Combustion Engines," SAE Int. J. Engines
8(3):2015, doi:10.4271/2015-01-0882.

1. INTRODUCTION

Gatowski et al.-model [1]. Experimental data are used to tune and


validate the model, in order to get a representative model with good
agreement to real engines.

With the drive and strive for efficient combustion engines several
things are investigated. One thing that has received much attention is
the optimal placement of the combustion in relation to the crank
angle revolution, and where 50% mass fraction burned has been used
in rule of thumbs. Here the investigation will be taken one step
further by looking at the complete combustion trace and asking the
question what is the most efficient combustion trace. This is partly
motivated by the availability to advanced fuel injection systems
where the rate of fuel injection can be controlled which could allow
rate shaping of the combustion. The other motivation is the
fundamental question about how the optimal profile looks like if it
could be selected freely. Rate shaping of the combustion is an
interesting option for optimizing the performance of combustion
engines, and the question studied here is how one can find the optimal
heat release rates and how they vary for varying conditions.

According to theory with idealized cycles the instantaneous burning


Otto cycle is the most efficient, however this is not true when
considering losses such as heat transfer, and the impact of the losses
are studied with the aid of the model.
The main contributions are:

To find the optimum heat release profile a methodology is developed


that uses available theory and methods that allows us to use state of
art optimal control software to solve the optimum combustion trace
for a model. The method developed focuses on the combustion rate
for optimal fuel economy, but the method is general as it can also be
used for emissions provided that there is a model available.

Modification of the Gatowski et al. model [1] that maintains


simplicity within the framework of an ideal gas.

A methodology for computing optimal burning trajectories,


demonstrating that current optimal control software it is feasible
to use for this application.

New knowledge about optimum burn profiles for a variety of


operating conditions.

The contributions are thus both the methodology and the resulting
profiles themselves.

2. SINGLE ZONE MODELING


The reduction of the effects of heat transfer and mass loss on the
cylinder pressure is called heat release analysis and is done within
the framework of the first law of thermodynamics. The data reduction

Focusing on the torque and fuel economy it is well known know that
a single zone model is sufficient for describing the engine cylinder
pressure and thereby the amount of work produced and therefore the
engine efficiency. The model uses the components in the well known
1069

1070

Eriksson et al / SAE Int. J. Engines / Volume 8, Issue 3 (June 2015)

also allows a consistency check of the pressure data itself. This model
family can also be inverted to take a heat release trace and deliver a
pressure as output.
A one-zone model offers the benefit of including heat transfer and gas
flow in a simple manner. The combustion process is considered as a
separate heat addition process and the contents of the chamber are
regarded as a single fluid. Straightforward heat transfer and crevice
models can then be used to complete the energy balance.
This paper makes some small extensions the well known model
presented in the paper [1] by Gatowski et al. that develops, tests and
applies a heat release analysis procedure that maintains simplicity
while including the effects of heat transfer and crevice flows. The
heat release model is a one zone description of the cylinder contents.
The thermodynamic properties are represented by a linear
approximation for (T). The cylinder pressure is influenced by
combustion, volume changes, heat transfer to the chamber walls, and
mass leakage.

determine the adiabatic flame temperature. However in single zone


models the working fluid is considered to be homogeneous and the
heat that is released from the combustion is considered to heat up the
cylinder contents so the released heat from the chemical reactions is
represented by Qch in the equations above and this raises the internal
energy of the cylinder contents. The computation of the single zone
heat release is expressed as follows

where the positive flow direction is out of the combustion chamber.


The work is piston work which is equal to

If it is assumed that there is only one flow the mass in the cylinder mc
will decrease with the same rate as the out-flow dmc = dmi, giving

The analysis and the model relies on the the first thermodynamic law
that states

Where h is the specific enthalpy of the flowing fluid, that depends on


the flow direction. The change in sensible internal energy is
calculated by assuming it as a function of the mean charge
temperature only, Us = mcu(T), which is the ideal gas assumption

where the directions of the heat, work, and enthalpy flows are defined
according to Figure 1.

here mc is the charge mass, and cv is the mass average specific heat at
constant volume. The mean temperature is determined by the ideal
gas law,

3. THE BASIC MODEL

(1)

Although this technique for calculating dU is not exact, the mean


temperature found from the ideal gas law is close to the mass
averaged cylinder temperature since the molecular weights of the
reactants and products are nearly identical [1].
Inserting the internal energy differential into the energy equation
gives

Figure 1. Thermodynamic control volume analysis for the combustion


chamber, defining the directions of positive flows. The state is specified by (V,
p, mc).

During the combustion, chemical energy is transformed into thermal


energy of the charge through a change in composition from the
unburned mixture to burnt reaction products. Standard
thermodynamics represents this by considering the internal energy of
the unburned mixture as the sum of a heat of formation at reference
conditions and a sensible energy difference. The internal energies of
both products and reactants are referred to the same datum so one can

In the next step a further simplification of the expression is


performed, in particular dT is expressed as a function of dV, dp and
dm, which is achieved by using the ideal gas law in both its normal
form and differential form.

(here again the assumption of an ideal gas is used, which means that
the change in the gas constant is neglected) rearranging the terms and
multiplying with cv gives

Eriksson et al / SAE Int. J. Engines / Volume 8, Issue 3 (June 2015)

1071

Another component to rewrite is the enthalpy h where the ideal gas


law p v = RT is used together with the definition h(T) = u(T) + p v
= u(T) + RT. Finally all is put together with the temperature
dependence included explicitly in the energy and specific heats
(4)

(2)

in the original paper [1] cv, R, and u - u were expressed using a


model for (T), but here the manipulations are stopped since new
models for cv(T) and u(T) can be used directly in (2). These will be
developed in the next step. Thereafter, the model for the combustion
chamber flow dmc, which is into and out of crevices, will be studied.

Modification 1: New Gas Property Model

The model is tuned to burned gas equilibrium data, obtained from the
CHEPP [6]. For completeness the absolute internal energy is modeled
as second order polynomial in

so the full function for the internal energy becomes is


The resulting model parameters
are given below and the fit to equilibrium data is shown in Figure 2.
The absolute internal energy at T = 300 K is shown below in Figure
3.

The sensitivity analysis performed on the model in [2] shows that the
model for the gas properties is the most important parameter in the
Gatowski et al. model. This motivated the study and search for an
alternative model for in [2]. Further publications that highlights the
importance of the gas properties are [3, 4], whereof the first presents
a high polynomial degree model for cv and the latter modifies the
Gatowski approach. With these previous approaches it is still difficult
to include the crevice effect and maintain simplicity in the equations,
as commented on in [2].
Here the goal is to improve the property model while giving a simple
calculation scheme for the both the analysis procedure and the
crevice model. This is achieved by staying with an ideal gas model
and exchanging the linear (T) model with a higher order polynomial
model for the internal energy u(, T) and cv(, T). Looking at
thermodynamic data for one sees that the air to fuel ratio shifts the
properties. In [5] it was implemented so that the offset of gamma was
allowed to shift. This shift is incorporated directly in the method here
where is included as a parameter in the property models.
In a single zone model the absolute level of the energy is not needed
as the process mainly deals with changes and differences in
properties. With a two zone model the absolute value would be
needed to determine the combusted gas temperature, so there is also a
component for this in the model here.
For lean conditions > 1 and at high pressures where there is little
dissociation and the internal energy and specific heat can be captured
well with the following structure of the polynomial in temperature
and mixture strength

Figure 2. Model and its validation for the temperature range 300-2000 K,
(solid lines p=1 atm, dashed p=10 atm, dotted p=100 atm). At high
temperatures T>1400 K and low pressures, there is dissociation but this is
neglected in the model. The model has thick dots, while the non dissociation
data has thin dots.

Furthermore for lean conditions molecular weight is also almost


constant, with M 0.0286 kg/mol giving R 290 J/kg K.
Another modification is to allow the mass in the cylinder to vary and
use the varying mass for temperature determination.

(3)

where
, with T given in K. This gives an internal energy
that function that is identical to 0 at T=300 K, differentiating this
gives

Crevice Effect
Crevice walls are cold and narrow and the volume is small so gases
in the crevice is assumed to be close to the wall temperature. Due to
these facts the crevices may contain substantial amounts of gas

1072

Eriksson et al / SAE Int. J. Engines / Volume 8, Issue 3 (June 2015)

during the period when the cylinder pressure is very high. Some gas
blows by the top ring into the volumes behind the rings and some of
this ends up in the crankcase.
A simple model which does not account for blowby is to consider a
single aggregate crevice volume with the same temperature as the
wall and the same pressure as the cylinder, i.e.
. In this
model it is assumed that there is no blow-by past the piston rings so
all mass stays in the combined volume of the cylinder and the
crevice. See Figure 4 for a sketch of the situation, illustrating that the
considered volume is above and behind the uppermost piston ring. In
the model here the single mass flux term considered, is the flow of
gas into and out of the crevice regions. The total trapped mass is
divided into the cylinder mass and the crevice mass

This gives,

which inserted into (2) gives

(6)

Where

(5)
(7)

Here it is worth to note that it is the difference in internal energy that


is evaluated and therefore only the sensible part is needed, u(T)
- u(T) = us(T) - us(T). The internal energy and specific heat depend on
, and this dependence is omitted here in order to keep the
nomenclature simple, this is done since is assumed to be constant
throughout the cycle.
As a side note on the importance of the crevice. The percentage of the
original charge present in the crevice at the end of the combustion
can approach 10%.

Modification 2: Using Internal Energy Directly

Figure 3. Model for the absolute internal energy at 300 K, as a function of . Solid - data,
dots - Model. Data is generated by CHEPP [6].

Without blow-by, mtot is constant which gives the following relation


for mass transfers between crevice and cylinder

In the original paper the linear model for (T) was used to determine
the internal energy difference, which resulted in a logarithm of a
quotient. With the new property model it is possible to use the
internal energy model (3) directly

Modification 3: Allowing a Variation in Mass for the


Temperature Calculation
In the original model [1] the cylinder gas temperature is determined
from the ideal gas law (1) but it is not explicitly stated how mc is
calculated there, so many implementations use the following
expression

Which only requires inlet valve closing conditions to be known and


does not need mc or R. However this model corresponds to the
assumption that the cylinder mass is constant during the cycle. Here
the total mass (5) is considered constant which is determined at intake
valve closing to
Figure 4. Sketch of the crevice model with the volume above and behind the top ring. The
pressure in the crevice is the same as in the cylinder, while the gases that enter the
crevices quickly become cooled and assume the same temperature as the wall
temperature. The mass in the crevice is therefore determined from the ideal gas law with
the aggregated crevice volume, cylinder pressure, and wall temperatue.

For every p and V the cylinder mass can now be calculated as

Eriksson et al / SAE Int. J. Engines / Volume 8, Issue 3 (June 2015)

1073

Woschni's Heat-Transfer Correlation


One of the most often used correlations is the form proposed in
Woschni [7] that is based on a correlation of the form Nu = 0.035
Rem. There it is shown that m = 0.8 produced a good fit to
measurement

and then ideal gas law (1) can be used to calculate the temperature.

Heat Transfer
Heat transfer by convection is the transfer of energy between a fluid
and a solid surface. The first phenomena is the diffusion or
conduction of energy through the fluid because of the presence of a
temperature gradient within the fluid. The diffusion and conduction is
a molecular transport phenomena with a rate controlled by the
thermophysical properties of the substance as well as the thermal
environment. The second is the transfer of energy within the fluid due
to the movement of the fluid from one thermal environment,
temperature field, to another. This phenomenon is associated with the
macroscopic characteristics, the movement or flow of the fluid, as
well as the thermophysical characteristics of the fluid and the thermal
characteristics of the solid.

where h is the heat transfer coefficient, k is the gas thermal


conductivity, B is cylinder bore, w is the characteristic speed, an is
the gas viscosity. Solving the equation above for the heat transfer
coefficient yields

After substituting the ideal gas law for the density, , and the assumed
temperature scaling for viscosity and conductivity, the heat transfer
coefficient is written as

The magnitude of the rate of energy transfer by convection, which


occurs in a direction perpendicular to the surface fluid interface,
is obtained by use of an expression referred to as Newton's law of
cooling

where A is the surface area of the body which is in contact with the
fluid, T is the temperature difference between the bulk gas
temperature T and a surface averaged wall temperature Tw, and h is
the convection heat transfer coefficient. The most important task is to
accurately predict the magnitude of the convection heat transfer
coefficient. Since this quantity is a composite of both microscopic
and macroscopic phenomena, many factors must be taken into
consideration. For many flow geometries, h, is given by the relation

which sometimes is rewritten to,

When using these heat-transfer correlations the critical choices to be


made are (1) the velocity to be used in the Reynolds number; (2) the
gas temperature at which to evaluate the gas properties in the
equation above; and (3) the gas temperature to be used in the
heat-transfer equation.
There exists several correlations for calculating the instantaneous
heat transfer where most are a Nusselt-Reynolds number similar to
those used in pipe flow, differing primarily in the way the Reynolds
and Nusselt numbers are defined. The lack of any generally accepted
engine heat transfer model attests to the uncertainty of this aspect of
heat release analysis.

(8)

Woschni chose the characteristic speed to be,

where the variables and units are:


pf
pm
Up

Pressure for firing cycle


Pressure for motored cycle
Mean piston speed

[bar]
[bar]
[m/s]

with the parameters


C1 = 6.18
C1 = 2.28
C1 = 2.28

C2 = 0
C2 = 0
C2 = 3.24 103

During gas exchange


During compression
During comb. and exp.

Now turning to (8), with m = 0.8 selected, since it replicates turbulent


flow in pipes, gives T0.546, but in the original work the temperature
dependence is set to T0.53 and the following expression is given

with units hc [kcal/m2 C], B [m], p [kp/cm2], T [K], w [m/s]. Other


publications (e.g. [8]) later sets the temperature dependence to T0.55,
which is closer to T0.546. Here the original formulation is used, and a
conversion to SI units gives the following expression for the heat
transfer coefficient

(9)

Eriksson et al / SAE Int. J. Engines / Volume 8, Issue 3 (June 2015)

1074

where u is the input heat release and where the following notation for
the numerator and denominator is introduced

where
p
p0
T
up
V
C1
C2
pivc,Vivc,Tivc

cylinder pressure, firing cycle


cylinder pressure, motored cycle
mean gas temperature
mean piston speed
instantaneous cylinder volume
constant 2.28
constant 3.24 103
reference condition at IVC

[Pa]
[Pa]
[K]
[m/s]
[m3]
[-]
[m/(s K)]
[Pa,m3,K]

This gives the following rate of heat transfer

where it is necessary to convert the rate of heat transfer given with


time, t, as independet variable to crank angle as independent variable.
This is achieved with the chain rule that gives.

where e is the engine speed in [rad/s].

(11a)

(11b)

During the simulation the direction of the crevice flow needs to be


determined, which depends on the sign of dp. Noting that the
denominator dpd is strictly positive enables us to only study the sign
of the numerator dpn to determine the flow direction and thus
determine T. Therefore the simulation is implemented so that the
numerator is first calculated which determines the sign of dp and thus
T which in its turn is inserted into denominator and the full
is
calculated. It is also worth to note that in the expressions above the
area A, volume V, and volume derivative
are functions of the
crank angle .

Vibe Heat Release Function

Pressure Simulation Model


The calculation scheme defined in (6) allows us to calculate heat
release from a measured cylinder pressure. However, if the heat
release Qch is specified as input one can now solve (6) for dp to get
an expression for the pressure differential and calculate (i.e. simulate)
the pressure.

In (10) the heat release


is an input and this will be studied for
two cases one when it is prescribed with the well known Vibe
function [9] and another when it is completely free. The Vibe
function is defined as

(12)

This equation is expressed with differentials so that one can select


either time t or crank angle as independent variable, selecting crank
angle as independent variable and inserting the sub-models from
above gives us

(10a)

(10b)

(10c)

(10d)

Figure 5. The burn angles and their definition for the mass fraction burned
trace. Top: Vibe function with the burn angles shown. Bottom: Derivative of
the Vibe function. In the figure (ign, d, b) = (10, 10, 20).

Eriksson et al / SAE Int. J. Engines / Volume 8, Issue 3 (June 2015)


where the parameter ign gives the start of combustion, and a are
related to combustion duration, and m influences the shape of the
burning profile. One must note that the Vibe function is
overparameterized, in particular a and can not be uniquely
determined (while the parameter m can).
Here the combustion placement and process will be characterized
with the ignition angle ign and the burn angles flame development
angle d and rapid burning angle b, defined in Figure 5. If the
burn angles d and b are available the Vibe parameters can be
calculated from them. Due to the over parameterization either or a
must be specified for a unique solution, and by specifying
beforehand (using for example = d + b) the Vibe parameters
can be calculated by

The rate of combustion heat release input is then specified using the
derivative of the Vibe-function as

1075

The wall temperature and initial temperature were then set to, Tw =
440 K, Tivc = 295 K. The initial pressure for the simulation was pivc =
41 103 Pa, and the compression ratio was rc = 10.14 which is near
the 10.1 that follows according to the data sheet for the engine. The
input energy was Qin = 595 J. For motoring cycles there is a good
agreement with C1 = 2.28 that is the same as the original paper, and
this should be the same for the firing cycles as they should have the
same conditions. But for firing cycles C2 had to be significantly
increased to C2 = 5 3.24 103 in order to give a good match
between the measured and modeled pressure. This high value could
be a result of other non-modeled effects, like e.g. charge amplifier
leakage, thermo-shock in the pressure sensor, or piston ring blow-by.
In the end this high parameter value is selected as default parameter
since it is this parameter value gives the best agreement between
modeled and measured pressures.
A validation of the single zone model for a normal cycle is shown in
Figure 6 where it is seen that the agreement between the modeled and
measured pressure is very good. Hence, showing that the model can
describe the work production process in the cylinder very accurately.

4. OPTIMAL COMBUSTION TRACES AND


BURN ANGLES
(13)

where Qin is the total amount of released heat.

Default Parameters and Model Validation


The nominal operating point is selected to be in the mid operating
region where the engine frequently operates during normal driving,
i.e. this is a point of importance for fuel economy of the engine, and
it is also far away from critical conditions like knocking and other
limits coupled to engine design and operation such as maximum
pressures and temperatures.

To simulate the pressure and work production, it is needed to specify


the combustion and this is here done in two steps. First, by looking at
specifying the combustion with the Vibe function [9] and burn angles.
Then, by looking at what happens when one can completely freely
control the combustion process.

Smooth Combustion with Vibe Function


A first study of the optimal combustion is performed using the well
known Vibe function [9] to specify the combustion process. When
specifying the combustion process it is characterized using the burn
angles defined in Figure 5.

The tuning of the model parameters has been performed in two steps,
the first tuning is towards motored cycles from an engine running
with skip fire, then against firing cycles. In the skip fire tests the
engine is run at steady state and 200 consecutive cycles are measured
for the firing engine, then skip firing starts and ignition is skipped
every 10:th cycle, so that in total 5 cycles have skipped of a total of
250 cycles. This stabilizes the engine an gives the motored cycle the
same initial conditions as the firing cycle and the model parameters
can be tuned to these cycles. To fit the model parameters to the
measured data the optimization based approach in [5] is used.
The crevice volume Vcr is selected based on measurements of the
geometry and engineering design considerations, for a cold engine
the total crevice volume (piston top land and behind piston ring), can
reach above 2% of the clearance volume Vc but while operation at
maximum power it can be around 1% of Vc. Based on the low load
point here the crevice volume is set to 1.5% of Vc in the default
model. This crevice volume was fixed and then the parameter tuning
was performed for the other parameters with the method in [5].

Figure 6. Validation of the agreement between the single zone model and
measured combustion traces. As is evident the agreement between the model
and simulation is very good.

1076

Eriksson et al / SAE Int. J. Engines / Volume 8, Issue 3 (June 2015)

Setup of Optimization Problem


In the optimization it is desirable to maximize the work output for a
given amount of input energy, which corresponds to the gross
indicated efficiency which gives the following criterion

Where the model (10) is used to simulate the pressure trace p(; ign,
d, b).
It is important to note that due to numerical round off errors during
the simulation the simulation doesn't necessarily reach an equality in
the amount of energy released

42 crank angle degrees is used (where the flame development angle is


selected to be d = 0.3b) and the optimum ignition timing is
determined for each case. The resulting traces are shown as solid
lines in Figure 7, the corresponding efficiency is shown in Figure 8.
There is a balance between timing losses and heat transfer and
crevice losses, the optimum rapid burning angle is about 12.7 where
42% mfb lies at 6.7 ATDC. A more rapid combustion gives higher
pressures and temperatures which increases the losses to heat transfer
and crevices.
If one starts to decrease the losses the optimal combustion becomes
more rapid and its center (about 42% mfb) moves towards TDC.
Removing the crevice loss, gives the traces in Figures 9 and 10,
where the optimum rapid burning angle 5.8, and the 42% is placed at
5.8 ATDC. In this case the gross indicated efficiency is also higher
now above 40%, since there is one loss mechanism less.

During simulation the differentiated Vibe function is then sampled


and sent as input to the simulation model and as a result of the
differentiation, sampling and integration the total energy released in
simulation does not become equal to Qin which is the motive for
formulating the problem as maximizing the efficiency.

Optimal Combustion Traces


With the problem definition above one can now solve the problem
above using simulation and searching for the optimal combustion
profile as specified by the smooth Vibe function. The model is
implemented in Matlab/Simulink and the problem is solved using
fminsearch in Matlab. The results are shown as a red dotted trace in
Figure 7.
Figure 8. Gross indicated efficiency as a function of rapid burn angle 1 to
42. The red dot is the burn angle for the optmium selection of burn angles
and ignition timing.

Figure 7. Pressure and mass fraction burned traces when the rapid burn angle
is swept from 1 to 42. The red trace is the optimal trace, when burn angles
and igntion timing is allowed to be selected freely.

To gain some insight into it the solution another set problems is


studied, where varying burn rates are specified and where optimum
ignition timing is determined. A sweep of rapid burn angles from 1 to

Figure 9. Pressure and mass fraction burned traces when the rapid burn angle
is swept from 1 to 42, when there is no crevice loss modelled. The red trace
is the optimal trace, when burn angles and igntion timing is allowed to be
selected freely.

Eriksson et al / SAE Int. J. Engines / Volume 8, Issue 3 (June 2015)


Finally in the lossless case the most efficient cycle is the one with the
most rapid combustion and it thus looks like the Otto cycle.

1077

Properties of the Optimal Control Problem


Diving into the optimal control theory one sees that the optimal control
problem is singular which essentially means that it is more difficult to
find an optimal solution to it. The model is also fairly complex so it is
difficult if not impossible to find an optimal solution analytically to the
problem. Therefore the problem is solved with numerical optimal
control software and the software package that is used to solve the
optimal control problem numerically is CasADi [10].
First the problem is discretized using Radau collocation with three
collocation points in each control interval. The resulting nonlinear
program(NLP) is solved using IPOPT, [11], with the MA57 linear
solver from the HSL package, [12]. To avoid making the NLP
excessively large, or the number of controls per crank angle
degree(nu/CAD) too small, the problem is solved as a three phase

Figure 10. Gross indicated efficiency as a function of rapid burn angle 1 to


42, when there is no crevice loss modelled. The red dot is the burn angle for
the optmium selection of burn angles and ignition timing.

Completely Free Combustion


Results from optimal control theory is used to calculate the optimal
solutions for the case when the combustion can be selected
completely freely. The problem is stated as follows. The produced
work is to be maximized

problem with
in phase one and three, and
in
the second phase, where the phase switch locations also are
optimized. The problem is solved on an iteratively finer mesh and all
results shown are with nu/CAD=0.5 in phase one and three, and nu/
CAD=5 in phase two. With numerical optimal control for non-convex
problems one is not guaranteed to find the global optimum.
Therefore, to ensure that the received solution is a good local
minimum the problem is solved using two different initial guesses,
first with the optimal Vibe results, and then with a motored cycle.
Both converge to the same solution. The successful implementation
and solution demonstrates that it is feasible to compute optimal burn
rate solutions to these problems, using optimal control software.

Solutions to the Optimal Control Problem


There are three main results shown below:

where system (pressure and released energy) has to obey the


differential equations in (10), specified here as

where the control signal u must fulfill u > 0 to ensure that there is an
increasing combustion profile (since the combustion is irreversible
and one can not undo combustion). Furthermore the total released
energy at the end point is specified and must be less than the total
available input energy Qch Qin. In this case the total released energy
will reach the limit so one does not need to normalize with the input
energy as was needed above when performing the parametric
optimization above. The denominator in (10) changes depending on
whether the flow is into or out from the crevice volume. This
introduces a potential discontinuity which is undesirable in an
optimization context. Therefore the function is smoothed out, using a
tanh-function as follows.

(14)

Nominal case, compared to optimal Vibe, and the impact of the


different losses, Figure 11.

Sweep in for the lean side, Figure 14.

Sweep in Vcr, Figure 13.

The optimal trajectories for the nominal case are shown in Figure
11-12 and compared to optimal Vibe, a loss-free case, i.e. without
crevice volume and heat transfer, as well as a case with heat transfer
but without crevice volume. The optimal control differ substantially
from the optimal Vibe control.
In the loss-free case the entire heat release occurs in one control
interval at the top dead center, resulting in an ideal Otto cycle like
pressure trace and efficiency. In the nominal case on the other hand,
the heat release consists of two phases, first a sharp pressure increase,
and then constant pressure, resembling a Seiliger cycle. This is even
clearer when looking at the pressure-Volume diagram in Fig. 12.
Even though the trajectories differ the optimal Vibe trajectory offers
close to optimal efficiency, the difference being just 1.15.

1078

Eriksson et al / SAE Int. J. Engines / Volume 8, Issue 3 (June 2015)


So far the discussed results have all been at stoichiometric conditions.
To study if the solution changes as the mixture gets leaner, and the
available energy therefore decreases, the optimal control problem is
solved for varying air-fuel ratio, , and the results are seen in Figure
14. The characteristics of the trajectories remain unaffected, however
the peak pressure, as well as the length of the constant pressure
phase, decreases with increasing . Looking at Fig. 15-top, the
efficiency increases quite drastically as a function of increasing ,
advocating lean combustion-type engines.

5. CONCLUSIONS

Figure 11. Optimal control results for the nominal case.

Looking at Figure 11 it is apparent that the constant pressure control


is a result of the crevice volume, and its associated losses. This raises
the question of how the solution changes with the crevice volume.
Therefore the optimal control problem is solved with varying crevice
volume and the results are shown in Fig. 13. The characteristics of
the solution are seen to be independent of Vcr as long as Vcr > 0,
however the length of the constant pressure phase increases with Vcr
and the peak pressure decreases as the losses to the crevice volume
increases. In Fig. 15-bottom it is seen that there are gains to be made
by making the crevice volume smaller. A 50% decrease in Vcr
increases the efficiency by 5whereas a 50% increase decreases the
efficiency by 4.

Optimal burn profiles for a combustion engine has been studied in the
process the well known Gatowski et al. [1] model has been extended
with three modifications. i) A more advanced gas model that still
keeps the total model within the framework of ideal gases. ii)
Inclusion of the new gas model in the crevice model. ii) explicit
tracking of the masses in the crevice and in the cylinder. The model is
parameterized and validated against measured pressure data, showing
good agreement between model and measurement.

Figure 14. Optimal control trajectories for a sweep in .


Figure 12. p-V plot for the nominal case

Figure 13. Optimal control trajectories for a sweep in Vcr.

Figure 15. Top: Efficiency as a function of . Bottom: Efficiency as a function


of Vcr. Red circle marks the nominal/actual values.

Eriksson et al / SAE Int. J. Engines / Volume 8, Issue 3 (June 2015)


Optimization of the combustion process for maximum efficiency is
then studied, for two cases: first parametric optimization is used to
determine the shape of the Vibe function that provides the highest
efficiency. Parametric studies are performed to see how the optimal
results change with changing model parameters like losses and air to
fuel ratio.
Then optimal control theory and computer tools are used to study
what an optimal combustion would look like if it could be selected
freely. The loss-free case is used as a benchmark to see that the
optimization framework finds the ideal Otto cycle. With losses
included it is shown that the optimal process is very close to the
Seiliger cycle, where the cause for the constant pressure process is a
result of the crevice model. The results from a parametric study show
that whenever there is a crevice there will be a constant pressure part,
even in cases when the heat transfer dominates the losses. It was also
shown that the combustion profile achieved with the Vibe function
can come close to the optimal profile, for the nominal case it was
within 1.
Furthermore when increasing the 42% mass fraction burned goes
towards TDC when doing the parametric optimization with the Vibe
function. For completely free combustion 0-50% mass fraction
burned remains at essentially the same place but the end of
combustion comes earler so the complete combustion is moved closer
to TDC when increases.
To conclude, this demonstrates that modern computer tools for
optimal control can be used to solve the problem of finding the
combustion profile that maximizes the engine work production.

1079

REFERENCES
1.

Gatowski, J., Balles, E., Chun, K., Nelson, F. et al., Heat Release
Analysis of Engine Pressure Data, SAE Technical Paper 841359, 1984,
doi:10.4271/841359.
2. Klein, M. and Eriksson, L., A Specific Heat Ratio Model for SingleZone Heat Release Models, SAE Technical Paper 2004-01-1464, 2004,
doi:10.4271/2004-01-1464.
3. Krieger R.B. and Borman G.L.. The computation of apparent heat
release for internal combustion engines. ASME, 1967.
4. Chang, J., Gralp, O., Filipi, Z., Assanis, D. et al., New Heat Transfer
Correlation for an HCCI Engine Derived from Measurements of
Instantaneous Surface Heat Flux, SAE Technical Paper 2004-01-2996,
2004, doi:10.4271/2004-01-2996.
5. Eriksson, L., Requirements for and a Systematic Method for Identifying
Heat-Release Model Parameters, SAE Technical Paper 980626, 1998,
doi:10.4271/980626.
6. Eriksson, L., CHEPP - A Chemical Equilibrium Program Package for
Matlab, SAE Technical Paper 2004-01-1460, 2004, doi:10.4271/200401-1460.
7. Woschni, G., A Universally Applicable Equation for the Instantaneous
Heat Transfer Coefficient in the Internal Combustion Engine, SAE
Technical Paper 670931, 1967, doi:10.4271/670931.
8. Heywood J. B.. Internal Combustion Engine Fundamentals. McGrawHill series in mechanical engineering. McGraw-Hill, 1988. ISBN 0-07100499-8.
9. Vibe I.I.. Brennverlauf und Kreisprocess von Verbennungsmotoren. VEB
Verlag Technik Berlin, 1970. German translation of the russian original.
10. Andersson Joel. A General-Purpose Software Framework for Dynamic
Optimization. PhD thesis, Arenberg Doctoral School, KU Leuven,
Department of Electrical Engineering (ESAT/SCD) and Optimization in
Engineering Center, Kasteelpark Arenberg 10, 3001-Heverlee, Belgium,
October 2013.
11. Wchter Andreas and Biegler Lorenz T.. On the implementation of
an interior-point filter line-search algorithm for large-scale nonlinear
programming. Mathematical Programming, 106:25-57, 2006.
12. HSL. A collection of fortran codes for large scale scientific computation.
http://www.hsl.rl.ac.uk, 2013. Last visited Oct. 2014.

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or
otherwise, without the prior written permission of SAE International.
Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE International. The author is solely responsible for the content of the paper.

Anda mungkin juga menyukai